Next Article in Journal
Synthesis of Magnetic Biosorbent from Bamboo Powders and Their Application for Methylene Blue Removal from Aqueous Solution: Kinetics, Isotherm, and Regeneration Studies
Next Article in Special Issue
A Facile One-Pot Preparation and Catalytic Application of Tunable Silica-Coated Aqueous Gold Nanoparticles
Previous Article in Journal
Antimicrobial Peptide Databases as the Guiding Resource in New Antimicrobial Agent Identification via Computational Methods
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Luminescent Manganese(II) Iminophosphorane Derivatives

by
Domenico Piccolo
1,2,
Jesús Castro
3,
Daniele Rosa-Gastaldo
1 and
Marco Bortoluzzi
2,4,*
1
Dipartimento di Scienze Chimiche, Università di Padova, Via Marzolo 1, 35131 Padova, Italy
2
Dipartimento di Scienze Molecolari e Nanosistemi, Università Ca’ Foscari Venezia, 30172 Mestre, Italy
3
Departamento de Química Inorgánica, Facultade de Química, Universidade de Vigo, Edificio de Ciencias Experimentais, 36310 Vigo, Spain
4
CIRCC (Consorzio Universitario Reattività Chimica e Catalisi), Via Celso Ulpiani 27, 70126 Bari, Italy
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(6), 1319; https://doi.org/10.3390/molecules30061319
Submission received: 15 February 2025 / Revised: 5 March 2025 / Accepted: 11 March 2025 / Published: 14 March 2025
(This article belongs to the Special Issue 30th Anniversary of Molecules—Recent Advances in Applied Chemistry)

Abstract

The reaction between the iminophosphorane ligand N-phenyl-1,1,1-triphenylphosphanimine (NPh=PPh3) and anhydrous manganese(II) halides allowed the isolation of complexes with the general formula [MnX2(NPh=PPh3)2] (X = Cl, Br, I). The compounds showed luminescence in the green region attributed to the 4T1(4G)6A1(6S) transition of the metal centre in the tetrahedral field, which was superimposed in the cases of X = Cl and X = Br on weak ligand-centred fluorescence. The emission and excitation spectra were compared with those of the free ligand and of the related zinc(II) bromo-complex. DFT calculations on the free ligand and on the manganese(II) bromo-complex helped to rationalise the experimental data. The protonation of NPh=PPh3 led to the formation of the iminium cation [NHPh=PPh3]+, which was used as a building block for the synthesis of organic–inorganic hybrids with the general formula [NHPh=PPh3]2[MnX4] (X = Cl, Br, I). The crystal structure of [NHPh=PPh3]2[MnBr4] was determined by means of X-ray diffraction. Green photoluminescence associated with the metal-centred transition was also observed for the organic–inorganic hybrids, with higher quantum yields with respect to the neutral [MnX2(NPh=PPh3)2] complexes. In the case of X = I, luminescence from the cation was superimposed on that from the tetraiodomanganate anion upon excitation of the compound with near–UV light.

1. Introduction

Luminescent manganese(II) derivatives are of actual interest for the development of devices based on earth-abundant elements and can have applications in several fields of advanced technology [1,2,3,4]. Organic–inorganic hybrids, where the anion is a manganese(II) halide derivative such as [MnX4]2−, [MnX5]3−, [Mn3X12]6− or [MnX3]nn−, constitute a class of widely investigated compounds that sometimes behave as multifunctional materials. Luminescence performance strongly depends on the choice of the organic counter-cation, which is constituted in several cases by nitrogen-based species [5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]. The photoluminescence quantum yield exceeds 95% in some cases.
Another class of extensively studied luminescent manganese(II) compounds is obtained with the introduction of suitable ligands in the coordination sphere, in particular, O-donors based on the {P=O} fragment. The tetrahedral complex [MnX2(O=PPh3)2], reported for the first time in 1961, represents the prototype for this family of manganese(II) derivatives [33,34,35,36,37,38]. The use of polydentate phosphine oxides allowed the preparation of compounds with improved and unusual optical features, and the studies were extended to other {P=O}-donors such as phosphoramides, arylphosphonic diamides, phosphoramidates, phosphonates and H-phosphinates [39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62].
The {P=NR} group of iminophosphoranes, also known as phosphanimines, is almost partly comparable to the {P=O} fragment. The phosphorus–nitrogen bond is highly polarised; thus, it is best described as a resonance hybrid between the two forms {P=NR} and {P+NR} [63]. The nitrogen atom can form coordinative bonds, or it can be protonated to obtain iminium cations. Examples of coordination compounds with iminophosphoranes include Groups 4–6 and 9–12 d-block elements as metal centres [64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79]. The bridging coordination mode was observed for the first time in the dinuclear carbonyl-complex [Mo2(CO)6(μ-NH=PPh3)3] [80].
Bidentate iminophosphoranes with donor moieties connected through either the nitrogen or phosphorus atoms were synthesised and applied for the preparation of d-block metal complexes [81,82,83,84,85]. It is, however, worth noting that several studies concern bis(iminophosphoranyl)methanides, i.e., the conjugate bases obtained from the deprotonation of the methylene bridge in R’N=PR2-CH2-PR2=NR’ species [86,87]. Examples are manganese(II) amide-complexes with the general formula [Mn{N(SiMe3)2}(ArN=PPh2-CH-PPh2=NAr)] (Ar = substituted aryl) [88,89].
The iminophoshorane moiety can also be part of multidentate species together with other donor moieties, and transition metal complexes stabilised by this type of ligands have found applications as catalysts for several organic reactions [90]. Palladium(II), platinum(II) and gold(III) derivatives of the luminescent ligand Ph3P=N–C9H6N (C9H6N = 8-quinolinyl) belong to this category, and they represent rare examples of iminophosphorane complexes investigated from a photophysical point of view [91].
Regarding the chemistry of manganese(II) with neutral iminophosphorane ligands, the complexes [MnCl(μ-Cl)(Me3SiN=PEt3)]2 and [MnI2(Me3SiN=PEt3)2] were prepared through the reaction of Me3SiN=PEt3 with the proper MnX2 precursor and structurally characterised. The thermal treatment of [MnI2(Me3SiN=PEt3)2] led to the chelate complex [MnI2(Et3P=N-SiMe2-N=PEt3)] [92]. No study concerning the luminescence of the manganese(II) complexes was reported. Moreover, to the best of our knowledge, iminophosphorane-based manganese(II) organic–inorganic hybrids were never synthesised. In the present work, we report the synthesis and photophysical characterisation of complexes having the general formula [MnX2(NPh=PPh3)2] and of organic–inorganic hybrids of the type [NHPh=PPh3]2[MnX4] (X = Cl, Br, I), as shown in Scheme 1. The structure of [NHPh=PPh3]2[MnBr4] was determined by means of single-crystal X-ray diffraction.

2. Results and Discussion

2.1. Characterisation of NPh=PPh3 and of Its Zinc(II) Bromide Derivative

N-phenyl-1,1,1-triphenylphosphanimine, NPh=PPh3, was prepared from triphenylphosphine and phenyl azide as reported in the literature [93]. The light yellow solid showed weak photoluminescence in the green region upon UV irradiation. The compound absorbed radiation for wavelengths shorter than 400 nm, with a local maximum at about 342 nm. The excitation (PLE) range at the solid state was wider than the absorption interval in solution, since it extended up to the violet region of the visible spectrum. The emission (PL) spectrum was composed of a single, broad band centred at 496 nm (Figure 1). Time-resolved measurements allowed determination of an average lifetime value (τ) around 2 ns, which is in line with a fluorescent emission. Photoluminescence data are summarised in Table 1. The photoluminescence quantum yield (Φ) of powder samples was too low for a precise determination, and it was estimated below 1%. The weak emission did not allow determination of the Φ value of the compound in solution.
Time-dependent density functional theory (TD-DFT) calculations allowed optimisation of the first-singlet excited state of NPh=PPh3. The predicted value for the S1S0 transition was 483 nm, which is in good agreement with the experimental outcomes. According to the hole–electron distribution depicted in Figure 1 [94], the transition occurred between orbitals mostly localised on the {NPh} fragment and on the P-bonded phenyl rings. The fluorescent emission thus had an intraligand charge transfer nature.
Attempts to further investigate the electronic features of NPh=PPh3 were carried out by means of cyclic voltammetry (Figure S1). The compound was characterised by three irreversible oxidation processes between 0.3 and 0.9 V vs. Fc+/Fc, with two clearly distinguishable peaks at 0.41 and 0.69 V vs. Fc+/Fc. The cathodic region showed an irreversible reduction peak around −2.07 V vs. Fc+/Fc, which was immediately followed by a dominant process attributable to electropolymerisation of the species [95].
In order to shed light on the effects of the interaction with an MX2 fragment on the electronic properties of NPh=PPh3, initial coordination chemistry studies were carried out considering anhydrous zinc(II) bromide as the metal precursor. [ZnBr2(NPh=PPh3)2] was prepared under mild conditions using anhydrous acetonitrile as the solvent and a 2:1 ligand/precursor ratio. All the manipulations were carried out under inert atmosphere given the limited stability of the complex in the presence of moisture. The NMR spectra of the complex showed the resonances expected for the iminophosphorane ligand, which were altered by coordination. In particular, a noticeable high-frequency shift around 29 ppm was observed for the 31P{1H} NMR resonance with respect to the free ligand (δ = 3.1 ppm, CDCl3, 300 K) because of the σ-donation of electron density from the {N=P} moiety to the metal centre. Another consequence of coordination is the broadening of the NMR signals, indicating fluxional behaviour in solution enhanced by the steric hindrance around zinc(II). According to literature data [86], the νP=N stretching of NPh=PPh3 was assigned to a strong band at 1346 cm−1, which was replaced by an absorption at 1240 cm−1 in [ZnBr2(NPh=PPh3)2]. Both NMR and IR data suggest that the charge-separated resonance formula Ph3P+NPh is more prevalent in the coordination compound than in the free ligand. Selected NMR and IR spectra are provided in the Supporting Materials, Figures S2 and S3.
The absorption features of [ZnBr2(NPh=PPh3)2] did not diverge from those of the free ligand. On the other hand, the emission at the solid state was slightly red-shifted, centred around 510 nm (FWHM = 5300 cm−1). The luminescence fell in the green region of the CIE 1931 chromaticity diagram. The PLE spectrum showed a maximum at 347 nm, roughly corresponding to the weak lowest energy transition detected in the corresponding absorption spectrum in solution (Figure 2). Time-correlated single-photon counting (TCSPC) measurements indicated that the excited-state lifetime was in the hundreds of nanoseconds range, approximately two orders of magnitude slower than free NPh=PPh3. Moreover, multi-channel scaling (MSC) measurements revealed the presence of a decay in the hundreds of μs range, which was most likely related to the population of an emitting triplet state (Figure S4). It must be therefore concluded that the luminescence of [ZnBr2(NPh=PPh3)2] is not due to simple fluorescent decay since the experimental data support the idea of exchange between singlet and triplet excited states.
Although not common, the possible involvement of triplet states has already been documented in the recent literature for other zinc(II) derivatives [96,97,98,99]. The previously described well-separated hole and electron distributions for the free iminophosphorane suggest a scarce contribution of the exchange energy in the relative stabilisation of the triplet state, favouring the reverse intersystem crossing (RISC) process. Such a hypothesis was confirmed by calculation of the excited singlet–triplet energy gap for the free ligand, which was carried out at both the excited singlet and triplet state geometries. The values obtained were around 900 cm−1, and it is worth noting that the possibility of RISC is usually invoked for energy gaps below 1000 cm−1 [100]. Unfortunately, the non-radiative decay processes were dominant, with a photoluminescence quantum yield (Φ) around 3%. Moreover, photodecomposition was observed after prolonged irradiation with UV light.
Cyclic voltammetry measurements on [ZnBr2(NPh=PPh3)2] (Figure S1) showed that the oxidation processes described for NPh=PPh3 remained irreversible, but they were shifted at higher potential. Regarding the cathodic region, it is worth noting that the coordination to ZnBr2 blocks the iminophosphorane electropolymerisation, a result highlighting that the complex is quite stable in solution despite the fluxional behaviour observed by means of NMR spectroscopy.

2.2. Synthesis and Luminescence of [MnX2(NPh=PPh3)2]

Complexes having the general formula [MnX2(NPh=PPh3)2] (X = Cl, Br, I) were obtained in good yield by reacting NPh=PPh3 with the proper anhydrous MnX2 salt in acetonitrile at room temperature. As previously reported for the related zinc(II) complex, the stability of the compounds in the presence of water traces was limited. The characterisation data agreed with the proposed formulations. In particular, the molar magnetic susceptibility values were in line with high-spin d5 first-row transition metal centres and corresponded to magnetic moments between 5.8 and 6.2 BM at 293 K. The IR spectra resembled that of [ZnBr2(NPh=PPh3)2], with νP=N stretching assigned around 1220 cm−1, which shifted to lower wavenumbers with respect to the free ligand (1346 cm−1) because of the enhanced role of the ionic ylide resonance form. The spectra showed negligible dependence upon the choice of the coordinated halides, as shown in Figure S5. The 31P{1H} NMR spectra were composed of a single resonance in the 31–39 ppm range, which broadened because of the paramagnetic relaxation (Figure S6). The 1H NMR spectra show signals in the aromatic region between 8.5 and 6.5 ppm (Figure S7). The less broadened resonances are those of [MnI2(NPh=PPh3)2], possibly because of different correlation times (electronic relaxation or ligands exchange) when X = I [101]. The 1H and 31P NMR chemical shifts were similar to those observed for [ZnBr2(NPh=PPh3)2], which is in agreement with the lack of paramagnetic shift expected for Mn(II) derivatives because of the 6S ground state and the consequent isotropic electronic distribution [101]. The UV–VIS of [MnCl2(NPh=PPh3)2] and [MnBr2(NPh=PPh3)2] in dichloromethane were similar to that recorded for [ZnBr2(NPh=PPh3)2] and showed absorptions for wavelengths shorter than 350 nm, with a general blue-shift with respect to the free ligand. A shoulder around 295 nm was enhanced in the iodo-complex, and a further weak band centred at 365 was present when X = I. The spectra are shown in Figure S8 together with that of free NPh=PPh3. Unfortunately, we were unable to obtain single crystals suitable for X-ray diffraction.
The formation of tetrahedral manganese(II) complexes was confirmed by the PL spectra. The three [MnX2(NPh=PPh3)2] compounds showed appreciable yellowish-green emission, and the related PL bands were centred between 528 and 538 nm, with FWHM values in the 2200−2900 cm−1 range (Figure 3). The PL bands appeared to be attributable to the 4T1(4G)6A1(6S) transition of the metal centre in the tetrahedral field. The assignment of the metal-centred emissions was supported by the PLE spectra (Figure 3), where the transitions from the sextet ground state to the lowest-energy quartet term were present between 420 and 530 nm. The shift of the 4A1 + 4E(4G)6A1(6S) transition from 432 to approximately 447 nm moving from X = Cl to X = I indicates a progressive reduction of the Racah B parameter. The average value, approximately 705 cm−1, is in line with previous data collected for other tetrahedral complexes [55]. The value reported for the free ion was 923 cm−1 [102]; therefore, the average nephelauxetic ratio β was around 0.76. The superposition of ligand- and metal-centred processes was present between 300 and 420 nm in the PLE spectra of [MnCl2(NPh=PPh3)2] and [MnBr2(NPh=PPh3)2], while the ligand-centred excitation was blue-shifted for [MnI2(NPh=PPh3)2]; thus, 4P,4D6S transitions centred at 390 and 372 nm were clearly detectable. The luminescence decay curves shown in Figure 3 are in line with the metal-centred nature of the emissions, since the lifetime values range from tens to thousands of μs, with a clear dependence upon the choice of the halide, as commonly observed for luminescent manganese(II) halide complexes. The τ values were meaningfully reduced on increasing the halide atomic number due to acceleration of the 4T1(4G)6A1(6S) radiative decay, which was ascribed to the increased degree of spin–orbit coupling [103].
Excitation with near–UV light of [MnCl2(NPh=PPh3)2] and [MnBr2(NPh=PPh3)2] caused the appearance of a weaker band around 460 nm. The same band was absent for the related iodo-complex. TCSPC measurements on [MnCl2(NPh=PPh3)2] indicated that the shoulder around 460 nm can be attributed to fluorescence decay from the coordinated ligands, with a τ value around 2 ns. The presence of free NPh=PPh3 was ruled out by the 31P{1H} NMR spectra. The ligand-centred emission was thus almost completely replaced by the metal-centred one, probably because of the energy transfer from the ligand-centred excited states to the Mn(II)-emitting level. The presence of only the metal-centred emission in [MnI2(NPh=PPh3)2] was reasonably related to the increased rate of intersystem crossing when X = I, which favoured the energy transfer.
Unfortunately, the luminescence quantum yields of the [MnX2(NPh=PPh3)2] were low at approximately 3% and 7% for the bromo- and iodo-complexes, respectively. The luminescence of [MnCl2(NPh=PPh3)2] was too low to determine the Φ value with the available experimental apparatus. Given a τ value equal to 55 μs, the radiative (kr) and non-radiative (knr) constants for [MnI2(NPh=PPh3)2] were around 1.3 × 103 s−1 and 1.7 × 104 s−1, respectively.
The antenna effect from the coordinated iminophosphoranes on the manganese(II) centre can be proposed on the basis of the PL and PLE spectra. To give insight into this point, the structure of [MnBr2(NPh=PPh3)2] was simulated on the basis of DFT calculations. A tetrahedral mononuclear complex was considered for the input structure, since the metal-centred emission in the green region of the spectrum allowed exclusion of higher coordination numbers and thus the formation of [MnBr2(NPh=PPh3)2]n polynuclear compounds of coordination polymers with bridging ligands. As shown in Figure S10, the spin density plots show that the unpaired electrons were localised on the metal centre for the sextet configuration, while a P-bonded phenyl ring was also involved in the unpaired electron density of the octet configuration. The computed energy difference between the two configurations was around 21,700 cm−1, while the energy of the 4T1(4G) level of MnII can be estimated to be around 20,400 cm−1 from the onset of the PL spectra, that of [MnI2(NPh=PPh3)2] in particular. The limited energy gap between ligand- and metal-centred excited states agrees with the possibility of ligandmetal energy transfer (ET) but also opens the possibility of the reverse process (BET). DFT calculations allowed the speculation that low quantum yields could be almost in part attributable to ligandmetal BET followed by non-radiative decay (Figure S10).

2.3. Synthesis of [NHPh=PPh3]2[MnX4], X-ray Structure Determination of the Bromo-Derivative and Luminescence Measurements

Organic–inorganic hybrids having the general formula [NHPh=PPh3]2[MnX4] (X = Cl, Br) were synthesised by reacting the corresponding MnX2 precursors with two equivalents of the iminium salts [NHPh=PPh3]X, which were obtained in situ from the iminophosphorane and HX in dichloromethane. [NHPh=PPh3]2[MnI4] was prepared by reacting a 1:2 mixture of MnI2 and NaI with two equivalents of [NHPh=PPh3][BF4], which were obtained in situ from NPh=PPh3 and HBF4·Et2O. Because of the progressive decomposition in the presence of water traces, the manipulations under air were avoided when possible. Elemental analysis data were in agreement with the proposed formulations, and the magnetic moments calculated from the molar magnetic susceptibility values at 293 K were close to the 5.9 BM expected for high-spin manganese(II) derivatives. The 31P{1H} NMR spectra were composed by a single resonance in the 36–39 ppm range (Figure S11), while the P- and N-bonded phenyl substituents were associated with broad resonances between 9.0 and 6.5 ppm in the 1H NMR spectra (Figure S12), without meaningful changes when varying the [MnX4]2− anion. The ATR-IR spectra of the three compounds were almost identical, with bands in the 3170–3130 cm−1 range indicating the presence of N-H bonds. The spectra also showed the absence of the strong band at 1346 cm−1 observed for NPh=PPh3, which was replaced by an absorption of medium intensity around 1220 cm−1 (see for instance Figure S13). The spectroscopic data suggest that protonation and coordination to MnII have similar effects on the electronic structure of NPh=PPh3. Such an outcome was confirmed by the UV–VIS spectra since the main absorptions in solution fell at wavelengths shorter than 325 nm, which were blue-shifted compared to free NPh=PPh3 (Figure S14). In the case of [NHPh=PPh3]2[MnI4], a weak band centred at 365 nm was also present.
The formation of organic–inorganic hybrids was confirmed by the single-crystal X-ray structure determination of [NHPh=PPh3]2[MnBr4]. The compound crystallised in the orthorhombic space group Pbca. The asymmetric unit (Figure 4) contained two [NH=PPh3]+ cations and one [MnBr4]2− dianion. Another molecule was present in the asymmetric unit, but it was not modelled, and the contribution to the reflections was eliminated by means of the usual procedure for disordered solvents molecules (see Section 3). Selected bond distances and angles are shown in Table 2.
The bond lengths around the manganese(II) centre ranged from 2.4748(11) to 2.5263(11) Å, and the angles were between 106.94(4) and 113.74(4)° and agreed well with those in similar compounds characterised by [MnBr4]2− tetrahedral units [104]. The coordination polyhedron was studied with the SHAPE program [SHAPE V2.1] [105]. Table S1 displays the output for the geometries expected in a four vertex polyhedron [106] and indicates that the environment around manganese(II) ion is best described as an almost perfect tetrahedron. In Table S1, the classical descriptors τ4 and τ’4 for the four-coordinate geometry are also reported [107,108], and the values agree with a tetrahedral geometry around the metal. In the phosphaniminium cations, the P-N bond distances were 1.633(5) and 1.631(5) Å. These values were slightly longer than that found in the free phosphanimine, 1.603(3) Å and were similarly found in the phosphaniminium cation crystallised with [AuI2] anion, 1.62(2) Å, or with some borate anions such as [PhC≡CB(C6F5)3], 1.6357(17) Å, or [HB(C6F5)3], 1.6415(19) Å [109,110]. The P-N-C angles, 128.6(4) and 128.5(4)°, were between those found for the neutral phosphanimine, 130.4(3)°, and for the borate derivatives, the average value was 126.60(16)°. The angles around the phosphorus atoms ranged between 104.9(3) and 112.2(3)°, a range similar to that found in the above-mentioned phosphaniminium salts.
The presence of bromine atoms resulted in the formation of hydrogen bonds. The classical H-bonds involved the NH fragments, but other interactions with several aromatic CH fragments were also present. The most important are set out in Table 3, and a graphical representation is given in Figure 5. In the supramolecular network, there was also at least one π-π’-stacking interaction between two aniline rings, as shown in Figure S15, with distances between centroids of 4.136(6) Å and ring slippage of 0.898 Å. The manganese atoms in Figure S15 were separated by 11.694(2) Å, but the shortest Mn–Mn’ distance found in the crystal was 10.573(1) Å (sym. op. 3/2 − x, y + 1/2, z), as shown in Figure S16.
The powder X-ray pattern of [NHPh=PPh3]2[MnBr4] was comparable to that derived from the single-crystal X-ray diffraction measurement, confirming the phase purity of the compound (Figure S17).
The three [NHPh=PPh3]2[MnX4] compounds showed appreciable yellowish-green luminescence at the solid state upon excitation with UV light, particularly the bromo- and iodo-derivatives. The PL spectra of [NHPh=PPh3]2[MnCl4] and [NHPh=PPh3]2[MnBr4] were composed of a single band centred between 529 and 532 nm, with FWHM values around 2400 cm−1, as expected for the 4T1(4G)6A1(6S) transition in the tetrahedral field. No dependence upon the excitation wavelength was observed, and the PLE spectra were essentially constituted by 4G6S (420–530 nm), 4P,4D6S (315–415 nm) and 4F6S (<300 nm) transitions. The emission of [NHPh=PPh3]2[MnI4] samples irradiated with near–UV light was markedly different, since the metal-centred transition was superimposed on another band centred at 565 nm (Figure 6). Only the band due to the 4T1(4G)6A1(6S) emission, centred at 537 nm, was present with excitation of the sample with radiation around 280 nm. The dependence of the PL spectrum upon the excitation wavelength is observable in the 3D plot reported in Figure 6, and the PLE spectrum revealed the superposition of the metal-centred transitions on another excitation band. Time-resolved measurements confirmed the dependence of the lifetime values from the choice of the coordinated halides, with monoexponential decays affording τ values equal to 2832 and 313 μs for the chloro- and the bromo-derivatives (Figure 7). Monoexponential decay was observed also for [NHPh=PPh3]2[MnI4] using λexcitation equal to 290 nm, a wavelength causing only metal-centred emission. The τ value obtained from the interpolation of the decay was 49 μs. On the other hand, TCSPC measurements carried out with λexcitation equal to 373 nm showed the curve reported in Figure 7. Using the previously determined τ value for the interpolation of the data, the lifetime of the second decay was estimated at around 10 μs. It is therefore likely to suppose that a phosphorescent process superimposes on the 4T1(4G)6A1(6S) emission of the tetraiodomanganate anion when the excitation is carried out with near–UV light.
In order to shed light into the dual emission of [NHPh=PPh3]2[MnI4], a freshly prepared sample of [NHPh=PPh3][BF4] was synthesised from NPh=PPh3 and HBF4·Et2O in dichloromethane and subjected to photophysical measurements after removal of the solvent. A very weak emission centred at 515 nm (FWHM = 5900 cm−1) was observed upon excitation with near–UV radiation (Figure S18). The lifetime was in the nanoseconds range, which was in line with a fluorescent decay. It can be tentatively supposed that the closeness of the [MnI4]2− anion favours the intersystem crossing in the [NHPh=PPh3]+ cations, with a red-shift of the emission of approximately 50 nm and a noticeable increase in the radiative lifetime. The same phenomenon did not occur for the other organic–inorganic hybrids because of the lower spin–orbit coupling.
As occurred for [MnCl2(NPh=PPh3)2], the emission of [NHPh=PPh3]2[MnCl4] was too low for a precise determination of the photoluminescence quantum yield. On the other hand, the Φ values of the other two organic–inorganic hybrids were between 22% and 23%, which were much higher than those of the neutral complexes. It is possible to suppose that the lack of covalent bonds between the metal centre and the iminophosphorane skeleton likely avoids the back energy transfer, thus reducing the non-radiative decay related to the vibration of the organic fragments. Given a τ value equal to 313 μs, the radiative (kr) and non-radiative (knr) constants for [NHPh=PPh3]2[MnBr4] were around 1.3 × 103 s−1 and 2.5 × 103 s−1, respectively. knr is thus approximately one order of magnitude slower than that estimated for [MnI2(NPh=PPh3)2].

3. Experimental Section

3.1. Materials and Methods

Anhydrous manganese(II) and zinc(II) halides and the other inorganic and organic reactants were purchased from Merck (Darmstadt, Germany). The solvents were purified before use according to established methods [111]. Deuterated chloroform was a Euriso-Top (Saarbrücken, Germany) product, which was used as received. Phenyl azide was synthesised from phenylhydrazine under a fume hood following a reported method. The distillation step was avoided, and the compound was purified by filtration on silica gel using pentane as eluent [112]. Safety note: organic azides are potentially explosive compounds. All the manipulations were carried out using limited quantities of reactants, with careful control of the temperature during all the synthetic steps. N-phenyl-1,1,1-triphenylphosphanimine (NPh=PPh3) was synthesised from phenyl azide and triphenylphosphine as reported in the literature [93]. Gaseous HCl and HBr were prepared by reacting the related sodium salts with concentrated sulfuric acid and they were absorbed in dichloromethane. The concentration was determined by titration immediately before use.
Moisture-sensitive compounds such as NPh=PPh3, its salts and coordination compounds were prepared and stored in a MBraun (Garching bei München, Germany) MB10 glove box filled with nitrogen and equipped for organic and inorganic syntheses.

3.2. Characterisations

Carbon, hydrogen and nitrogen elemental analyses were performed using an Elementar (Langenselbold, Germany) Unicube microanalyzer. The halide contents were determined by means of Mohr’s method [113]. Magnetic susceptibility measurements were carried out on solid samples at 298 K and 3.5 kGauss magnetic field strength using an MK1 magnetic susceptibility balance (Sherwood Scientific Ltd., Cambridge, UK). The measured magnetic susceptibility was corrected for the diamagnetic contribution using tabulated Pascal’s constants [114]. IR spectra of the neutral species dispersed in KBr (spectroscopy grade, Merck) were collected in the 4000–450 cm−1 range using a Perkin–Elmer (Shelton, CT, USA) Spectrum One spectrophotometer. The samples were prepared in a glove box. ATR-IR spectra of the ionic compounds were recorded with a Perkin–Elmer Spectrum Two spectrophotometer equipped with diamond ATR. 1H and 31P{1H} NMR spectra were collected employing a Bruker Avance 400 instrument (Billerica, MA, USA) operating at 400.13 MHz of 1H resonance. 1H NMR spectra were referenced to the partially non-deuterated fraction of the solvent, which is itself quoted to tetramethylsilane. 31P chemical shifts are referred to external 85% H3PO4 in water. Cyclic voltammetry measurements in acetone containing 0.1 M LiClO4 were conducted using an eDAQ (Denistone, Australia) ET014-199 instrument coupled with an eDAQ ET074-1 glassy carbon working electrode (1 mm diameter) and an eDAQ ET078-1 Pt-coated titanium rod auxiliary electrode. All the measurements were carried out under argon at room temperature. Ferrocene was introduced as an internal standard, and a Pt wire was used as pseudo-reference electrode. CV data are reported following the IUPAC convention [115].

3.3. Synthesis of [ZnBr2(NPh=PPh3)2]

Anhydrous ZnBr2 (0.113 g, 0.5 mmol) was dissolved in 25 mL of acetonitrile, and then NPh=PPh3 (0.353 g, 1.0 mmol) was added under stirring. The reaction mixture was kept under stirring at room temperature for 12 h, and then the solvent was removed by evaporation at reduced pressure. Dichloromethane (25 mL) was added, and the solution was purified by centrifugation. After evaporation under reduced pressure, diethyl ether (10 mL) was added, and the solid that separated was collected by filtration, washed with 5 mL of diethyl ether and dried under a vacuum. The yield was 0.303 g (65%).
Characterisation of [ZnBr2(NPh=PPh3)2]. Anal. calcd. for C48H40Br2N2P2Zn (931.99 g mol−1, %): C, 61.86; H, 4.33; N, 3.01; Br, 17.15. Found (%): 61.62; H, 4.35; N, 2.98; Br, 17.08. IR (KBr, cm−1): 1249 sh, 1240 m (νPN). 1H NMR (CDCl3, 300 K) δ 8.00–7.35 (m, br, 15H, P-Ph), 7.15–6.80 (m, br, 5H, N-Ph). 31P{1H} NMR (CDCl3, 300 K) δ 32.0 (s, FWHM = 100 Hz).

3.4. Synthesis of [MnX2(NPh=PPh3)2] (X = Cl, Br, I)

[MnX2(NPh=PPh3)2] complexes (X = Cl, Br, I) were synthesised as described for [ZnBr2(NPh=PPh3)2] using 0.5 mmol of the proper manganese(II) halide (X = Cl, 0.063 g; X = Br, 0.107 g; X = I, 0.154 g). The solubilisation of MnCl2 before the addition of the ligand was favoured by heating to reflux the reaction mixture for 30 min. Yields: X = Cl, 0.246 g (59%); X = Br, 0.327 g (71%); X = I, 0.355 g (70%).
Characterisation of [MnCl2(NPh=PPh3)2]. Anal. calcd. for C48H40Cl2N2P2Mn (832.64 g mol−1, %): C, 69.24; H, 4.84; N, 3.36; Cl, 8.52. Found (%): 68.95; H, 4.87; N, 3.30; Cl, 8.48. χMcorr (cgsu): 1.65 × 10−2. μ (293 K, BM): 6.2. IR (KBr, cm−1): 1220 m (νPN). 31P{1H} NMR (CDCl3, 300 K) δ 38.0 (s, FWHM = 370 Hz).
Characterisation of [MnBr2(NPh=PPh3)2]. Anal. calcd. for C48H40Br2N2P2Mn (921.54 g mol−1, %): C, 62.56; H, 4.38; N, 3.04; Br, 17.34. Found (%): 62.32; H, 4.41; N, 3.00; Br, 17.27. χMcorr (cgsu): 1.60 × 10−2. μ (293 K, BM): 6.1. IR (KBr, cm−1): 1226 m, 1222 m (νPN). 31P{1H} NMR (CDCl3, 300 K) δ 36.2 (s, FWHM = 120 Hz).
Characterisation of [MnI2(NPh=PPh3)2]. Anal. calcd. for C48H40I2N2P2Mn (1015.54 g mol−1, %): C, 56.77; H, 3.97; N, 2.76; I, 24.99. Found (%): 56.55; H, 4.00; N, 2.73; I, 25.05. χMcorr (cgsu): 1.46 × 10−2. μ (293 K, BM): 5.8. IR (KBr, cm−1): 1217 m (νPN). 31P{1H} NMR (CDCl3, 300 K) δ 31.8 (s, FWHM = 150 Hz).

3.5. Synthesis of [NHPh=PPh3)]2[MnX4] (X = Cl, Br)

A freshly prepared 2.0 M solution of HX (X = Cl, Br) in dichloromethane (0.5 mL) was added to 0.353 g (1.0 mmol) of NPh=PPh3 dissolved in 10 mL of dichloromethane. After 30 min under stirring at room temperature, the solution was added to an acetonitrile solution (20 mL) containing 0.5 mmol of anhydrous MnX2 (X = Cl, 0.063 g; X = Br, 0.107 g). The resulting reaction mixture was left under stirring for 12 h, then the solvents were removed by evaporation under reduced pressure and diethyl ether (10 mL) was added. The solid products were collected by filtration, washed with 5 mL of diethyl ether and dried under vacuum. Yields: X = Cl, 0.403 g (89%); X = Br, 0.493 g (91%). Crystals of [NHPh=PPh3)]2[MnBr4] suitable for X-ray diffraction were obtained by slow diffusion of diethyl ether in dichloromethane solutions.
Characterisation of [NHPh=PPh3)]2[MnCl4]. Anal. calcd. for C48H42Cl4MnN2P2 (905.56 g mol−1, %): C, 63.66; H, 4.67; N, 3.09; Cl, 15.66. Found (%): 63.40; H, 4.70; N, 3.05; Cl, 15.72. χMcorr (cgsu): 1.45 × 10−2. μ (293 K, BM): 5.8. ATR-IR (cm−1): 3134 (νNH), 1220 (νPN). 31P{1H} NMR (CDCl3, 300 K) δ 38.8 (FWHM = 190 Hz).
Characterisation of [NHPh=PPh3)]2[MnBr4]. Anal. calcd. for C48H42Br4MnN2P2 (1083.36 g mol−1, %): C, 53.22; H, 3.91; N, 2.59; Br, 29.50. Found (%): 53.11; H, 3.88; N, 2.55; Br, 29.61. χMcorr (cgsu): 1.48 × 10−2. μ (293 K, BM): 5.9. ATR-IR (cm−1): 3134 (νNH), 1219 (νPN). 31P{1H} NMR (CDCl3, 300 K) δ 37.8 (FWHM = 260 Hz).

3.6. Synthesis of [NHPh=PPh3)]2[MnI4]

Anhydrous MnI2 (0.309 g, 1.0 mL) was dissolved in 30 mL of acetonitrile together with 0.300 g (2.0 mmol) of NaI. In another flask, NPh=PPh3 (0.707 g, 2.0 mmol) was dissolved in 20 mL of dichloromethane, and HBF4·Et2O (275 μL, 2.0 mmol) was added. After 30 min under stirring at room temperature, the two solutions were mixed and left under stirring at room temperature for 12 h. The reaction mixture was purified by centrifugation, and then the solvents were evaporated under reduced pressure. The addition of diethyl ether (10 mL) caused the separation of a solid that was filtered, washed with 5 mL of diethyl ether and dried under a vacuum. The yield was 0.439 g (69%).
Characterisation of [NHPh=PPh3)]2[MnI4]. Anal. calcd. for C48H42I4MnN2P2 (1271.36 g mol−1, %): C, 45.35; H, 3.33; N, 2.20; I, 39.93. Found (%): 45.15; H, 3.29; N, 2.15; I, 40.08. χMcorr (cgsu): 1.42 × 10−2. μ (293 K, BM): 5.8. ATR-IR (cm−1): 3169 (νNH), 1225 (νPN). 31P{1H} NMR (CDCl3, 300 K) δ 36.4 (FWHM = 80 Hz).

3.7. Absorption and Luminescence Measurements

Absorption spectra in dichloromethane at room temperature were recorded in a 10 mm × 10 mm quartz cuvette (Hellma GmbH, Müllheim, Germany) with a Yoke 6000Plus double-beam spectrophotometer (Fengxian, China). Photoluminescence emission (PL) and excitation (PLE) spectra as well as lifetime decay curves were registered on solid samples at room temperature using a Horiba Jobin Yvon (Kyoto, Japan) Fluorolog-3 spectrofluorometer. Air-tight quartz sample holders were used and filled in the glove box to avoid interactions of the air-sensitive complexes with moisture. A continuous wave xenon arc lamp was used as the source, and the excitation wavelength was selected using a double Czerny–Turner monochromator. Suitable long-pass filters were placed in front of the acquisition systems. The detector was composed of a single Horiba (Kyoto, Japan) iHR 320 monochromator and a Hamamatsu (Shizuoka, Japan) R928 photomultiplier tube. The excitation and emission spectra were corrected for the instrumental functions. Time-resolved analyses were performed in multi-channel scaling mode (MCS) or time-correlated single photon counting mode (TCSPC) employing Horiba SpectraLED and NanoLED pulsed sources. The room-temperature photoluminescence quantum yields (Φ) at the solid state were measured employing an OceanOptics (Orlando, FL, USA) HR4000CG UV–vis–NIR detector fibre-coupled to an integrating sphere connected to OceanOptics UV LED continuous sources (λmax = 310 and 365 nm). The values are reported as the average of three measurements.

3.8. Single Crystal X-ray Structure Determination and Powder X-ray Measurements

Crystallographic data relative to [NHPh=PPh3]2[MnBr4] were collected at CACTI (Universidade de Vigo) with a Bruker (Billerica, MA, USA) D8 Venture Photon II CMOS detector. The Mo-Kα radiation (λ = 0.71073 Å) was generated by an Incoatec (Geesthacht, Germany) Microfocus Source IµS. The software APEX version 2022.1-1 was used for collecting frames of data, indexing reflections and determination of lattice parameters, SAINT version 8.40B was used for integration of the intensity of reflections, and SADABS version 2016/2 was used for scaling and empirical absorption correction [116]. The crystallographic treatment was performed with the Oscail version 4.7.1 program [117] and solved using the SHELXT version 2018/2 program [118]. The structure was subsequently refined by a full-matrix least-squares based on F2 using the SHELXL version 2019/2 program [119]. Non-hydrogen atoms were refined with anisotropic displacement parameters. A disorder solvent was found in the final density map, but its contribution to the reflections was eliminated by using the usual SQUEEZE procedure of PLATON version 110423 [120]. The same PLATON suite was also used to obtain some geometrical parameters from the cif file. Further details concerning crystal data and structural refinement are given in Table 4. CCDC 2421580 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from the Cambridge Crystallographic Data Centre via https://www.ccdc.cam.ac.uk/structures/ (accessed on 10 March 2025). Powder X-ray measurements were carried out at room temperature with a Malvern Panalytical (Malvern, UK) Empirean diffractometer.

3.9. Computational Details

Computational geometry optimisations were carried out using the PBEh-3c method, which is a reparametrised version of the hybrid-GGA PBE0 functional (with 42% HF exchange) that uses a split-valence double-zeta basis set (def2-mSVP) and adds three corrections considering dispersion, basis set superposition and other basis set incompleteness effects [121,122,123,124]. The stationary points were characterised as local minima my means of IR simulations (harmonic approximation). The same method was applied for TD-DFT calculations, which were carried out considering 16 states. The unrestricted approach was used for compounds containing unpaired electrons, and the absence of meaningful spin contamination was verified by comparing the computed <S2> values with the theoretical ones. The software used was ORCA version 5.0.3 [125]. The output was elaborated using MultiWFN, version 3.8 [94,126,127]. Cartesian coordinates of the DFT-optimised structures are provided in the Supporting Materials, Lists S1–S4.

4. Conclusions

Manganese(II) iminophosphorane derivatives were subjected to an investigation of their photophysical features for the first time. The ligand N-phenyl-1,1,1-triphenylphosphanimine was revealed to be suitable for the preparation of luminescent tetrahedral manganese(II) halide complexes, with sensitisation of the metal-centred luminescence. The photoluminescence quantum yields of the neutral complexes resulted, however limited, most likely because of metal-to-ligand back energy transfer followed by non-radiative decay. The protonation of N-phenyl-1,1,1-triphenylphosphanimine led to the formation of the corresponding iminium cation, from which luminescent organic–inorganic hybrids containing tetrahedral tetrahalomanganate anions were prepared. The photoluminescence quantum yield values were higher than those of the corresponding neutral complexes but were unfortunately lower than the best results reported in the recent literature. The present work, however, highlighted the versatility of iminophosphoranes for the preparation of manganese(II) luminescent derivatives, and future research will be focused on the replacement of the N- and P-bonded substituents in the NR’=PR3 skeleton to improve the luminescent features. Bidentate iminophosphoranes will be also subjected to investigation in this field.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules30061319/s1: Figure S1: Cyclic voltammograms of [ZnBr2(NPh=PPh3)2] and NPh=PPh3; Figure S2: 1H NMR and 31P{1H} NMR spectra of [ZnBr2(NPh=PPh3)2]; Figure S3: IR spectra of [ZnBr2(NPh=PPh3)2] and NPh=PPh3; Figure S4: Luminescence decay curves of [ZnBr2(NPh=PPh3)2]; Figure S5: IR spectra (KBr) of [MnX2(NPh=PPh3)2] (X = Cl, Br, I) and of NPh=PPh3; Figure S6: 31P{1H} NMR spectra of [MnX2(NPh=PPh3)2]; Figure S7: 1H NMR spectra of [MnX2(NPh=PPh3)2]; Figure S8: UV-VIS spectra of [MnX2(NPh=PPh3)2] (X = Cl, Br, I) and of NPh=PPh3; Figure S9: PL spectra of [MnCl2(NPh=PPh3)2] recorded at different excitation wavelengths; Figure S10: DFT-optimised structures of [MnBr2(NPh=PPh3)2], sextet and octet configurations, with spin density surfaces and energy differences between the two structures compared to the 4T1(4G)–6A1(6S) experimental energy gap; Figure S11: 31P{1H} NMR spectra of [NHPh=PPh3]2[MnX4]; Figure S12: 1H NMR spectra of [NHPh=PPh3]2[MnX4]; Figure S13 ATR-IR spectrum of [NHPh=PPh3]2[MnI4]; Figure S14: UV–VIS spectra of [NHPh=PPh3]2[MnX4] (X = Cl, Br, I) and of NPh=PPh3; Table S1: Output of the SHAPE software and four-coordinate geometry indexes τ4 and τ’4; Figure S15: π,π’-stacking interaction; Figure S16: Unit cell content showing the shortest distance between two Mn atoms; Figure S17: Comparison of the X-ray patterns of [NHPh=PPh3]2[MnBr4] obtained from powder and single crystal diffraction measurements; Figure S18: Normalised PL and PLE spectra of [NHPh=PPh3][BF4]; List S1: Cartesian coordinates of NPh=PPh3, ground state; List S2: Cartesian coordinates of NPh=PPh3, first singlet excited state; List S3: Cartesian coordinates of [MnBr2(NPh=PPh3)2], sextet state; List S4: Cartesian coordinates of [MnBr2(NPh=PPh3)2], octet state.

Author Contributions

Conceptualisation, M.B.; methodology, M.B., D.P. and J.C.; validation, D.P., J.C., D.R.-G. and M.B.; formal analysis, M.B., D.P. and J.C.; investigation, M.B., D.P. and J.C.; resources, J.C., D.R.-G. and M.B.; data curation, M.B., J.C. and D.P.; writing—original draft preparation, M.B., J.C. and D.P.; writing—review and editing, D.P., J.C., D.R.-G. and M.B.; visualisation, M.B., D.P. and J.C.; supervision, D.R.-G. and M.B.; project administration, D.R.-G. and M.B.; funding acquisition, D.R.-G. All authors have read and agreed to the published version of the manuscript.

Funding

Project funded under the National Recovery and Resilience Plan (NRRP), Mission 4 Component 2 Investment 1.3—Call for tender No. 1561 of 11.10.2022 of Ministero dell’Università e della Ricerca (MUR); funded by the European Union—NextGenerationEU, project code PE0000021, Concession Decree No. 1561 of 11.10.2022 adopted by Ministero dell’Università e della Ricerca (MUR), CUP C93C22005230007, project title “Network 4 Energy Sustainable Transition—NEST.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

CACTI (University of Vigo) and CINECA (Bologna) are respectively acknowledged for X-ray data collection and the availability of high-performance computing resources (class C project INLIGHT). We thank Tiziano Finotto (Ca’ Foscari University of Venice) for the powder X-ray measurements.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses or interpretation of the data; in the writing of the manuscript or in the decision to publish the results.

References

  1. Qin, Y.; She, P.; Huang, X.; Huang, W.; Zhao, Q. Luminescent manganese(II) complexes: Synthesis, properties and optoelectronic applications. Coord. Chem. Rev. 2020, 416, 213331. [Google Scholar] [CrossRef]
  2. Tao, P.; Liu, S.-J.; Wong, W.-J. Phosphorescent manganese(II) complexes and their emerging applications. Adv. Opt. Mater. 2020, 8, 2000985. [Google Scholar] [CrossRef]
  3. Liang, D.; Xiao, H.; Cai, W.; Lu, S.; Zhao, S.; Zang, Z.; Xie, L. Mn2+-Based Luminescent Metal Halides: Syntheses, Properties, and Applications. Adv. Opt. Mater. 2023, 11, 2202997. [Google Scholar] [CrossRef]
  4. Zhang, W.; Zheng, W.; Li, L.; Huang, P.; Xu, J.; Zhang, W.; Shao, Z.; Chen, X. Unlocking the Potential of Organic-Inorganic Hybrid Manganese Halides for Advanced Optoelectronic Applications. Adv. Mater. 2024, 36, 2408777. [Google Scholar] [CrossRef]
  5. Zhang, Y.; Liao, W.-Q.; Fu, D.-W.; Ye, H.-Y.; Liu, C.-M.; Chen, Z.-N.; Xiong, R.-G. The first organic-inorganic hybrid luminescent multiferroic: (pyrrolidinium)MnBr3. Adv. Mater. 2015, 27, 3942–3946. [Google Scholar] [CrossRef]
  6. Zhang, Y.; Liao, W.-Q.; Fu, D.-W.; Ye, H.-Y.; Chen, Z.-N.; Xiong, R.-G. Highly efficient red-light emission in an organic-inorganic hybrid ferroelectric: (pyrrolidinium)MnCl3. J. Am. Chem. Soc. 2015, 137, 4928–4931. [Google Scholar] [CrossRef]
  7. Ye, H.-Y.; Zhou, Q.; Niu, X.; Liao, W.-Q.; Fu, D.-W.; Zhang, Y.; You, Y.-M.; Wang, J.; Chen, Z.-N.; Xiong, R.-G. High-temperature ferroelectricity and photoluminescence in a hybrid organic-inorganic compound: (3-Pyrrolinium)MnCl3. J. Am. Chem. Soc. 2015, 137, 13148–13154. [Google Scholar] [CrossRef]
  8. Jiang, C.; Zhong, N.; Luo, C.; Lin, H.; Zhang, Y.; Peng, H.; Duan, C.-G. (Diisopropylammonium)2MnBr4: A multifunctional ferroelectric with efficient green-emission and excellent gas sensing properties. Chem. Commun. 2017, 53, 5954–5957. [Google Scholar] [CrossRef]
  9. You, Y.-M.; Liao, W.-Q.; Zhao, D.; Ye, H.-Y.; Zhang, Y.; Zhou, Q.; Niu, X.; Wang, J.; Li, P.-F.; Fu, D.-W.; et al. An organic-inorganic perovskite ferroelectric with large piezoelectric response. Science 2017, 357, 306–309. [Google Scholar] [CrossRef]
  10. Mei, Y.; Yu, H.; Wei, Z.; Mei, G.; Cai, H. Two coordinated geometries of Mn2+ ions in one single molecule: Organic–inorganic hybrids constructed with tris(2-aminoethyl)amine and manganese halide and fluorescent properties. Polyhedron 2017, 127, 458–463. [Google Scholar] [CrossRef]
  11. Chen, S.; Gao, J.; Chang, J.; Zhang, Y.; Feng, L. Organic-inorganic manganese (II) halide hybrids based paper sensor for the fluorometric determination of pesticide ferbam. Sens. Actuators B 2019, 297, 126701. [Google Scholar] [CrossRef]
  12. Mao, L.; Guo, P.; Wang, S.; Cheetham, A.K.; Seshadri, R. Design principles for enhancing photoluminescence quantum yield in hybrid manganese bromides. J. Am. Chem. Soc. 2020, 142, 13582–13589. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, X.; Ge, C.; Yang, Z.; Song, Y.; Wang, A.; Kang, Y.; Li, B.; Dong, Q. Guanidinetemplated manganese halides single crystals toward efficient mechanoluminescence and photoluminescence by supramolecular interactions modulation. Adv. Opt. Mater. 2021, 9, 2100862. [Google Scholar] [CrossRef]
  14. Zhang, S.; Zhao, Y.; Zhou, J.; Ming, H.; Wang, C.-H.; Jing, X.; Ye, S.; Zhang, Q. Structural design enables highly-efficient green emission with preferable blue light excitation from zero-dimensional manganese(II) hybrids. Chem. Eng. J. 2021, 421, 129886. [Google Scholar] [CrossRef]
  15. Zhang, S.; Zhao, Y.; Zhou, Y.; Li, M.; Wang, W.; Ming, H.; Jing, X.; Ye, S. DipoleOrientation-dependent Förster resonance energy transfer from aromatic head groups to MnBr42− blocks in organic-inorganic hybrids. J. Phys. Chem. Lett. 2021, 12, 8692–8698. [Google Scholar] [CrossRef]
  16. Jana, A.; Sree, V.G.; Ba, Q.; Cho, S.C.; Lee, S.U.; Cho, S.; Jo, Y.; Meena, A.; Kim, H.; Im, H. Efficient organic manganese(II) bromide green-light-emitting diodes enabled by manipulating the hole and electron transport layer. J. Mater. Chem. C 2021, 9, 11314–11323. [Google Scholar] [CrossRef]
  17. Wang, S.; Han, X.; Kou, T.; Zhou, Y.; Liang, Y.; Wu, Z.; Huang, J.; Chang, T.; Peng, C.; Wei, Q.; et al. Lead-free MnII-based red-emitting hybrid halide (CH6N3)2MnCl4 toward high performance warm WLEDs. J. Mater. Chem. C 2021, 9, 4895–4902. [Google Scholar] [CrossRef]
  18. He, J.; Zhao, H.; Hu, X.; Fangg, Z.; Wang, J.; Zhang, R.; Zheng, G.; Zhou, B.; Long, F. Synthesis, structure, and photoelectric properties of novel 0-dimensional organicinorganic hybrid perovskite (2-5-py)2MnBr4. J. Phys. Chem. C 2021, 125, 22898–22906. [Google Scholar] [CrossRef]
  19. Yan, S.; Tian, W.; Chen, H.; Tang, K.; Lin, T.; Zhong, G.; Qiu, L.; Pan, X.; Wang, W. Synthesis of 0D Manganese-Based Organic–Inorganic Hybrid Perovskite and Its Application in Lead-Free Red Light-Emitting Diode. Adv. Funct. Mater. 2021, 31, 2100855. [Google Scholar] [CrossRef]
  20. Berezin, A.S.; Davydova, M.P.; Samsonenko, D.G.; Sukhikh, T.S.; Artem’ev, A.V. A family of brightly emissive homo- and mixed-halomanganates(II): The effect of halide on optical and magnetic properties. J. Lumin. 2021, 236, 118069. [Google Scholar] [CrossRef]
  21. Liu, H.-L.; Ru, H.-Y.; Sun, M.-E.; Wang, Z.-Y.; Zang, S.-Q. Organic−Inorganic Manganese Bromide Hybrids with Water-Triggered Luminescence for Rewritable Paper. Adv. Opt. Mater. 2022, 10, 2101700. [Google Scholar] [CrossRef]
  22. Zhang, H.; Tan, Y.-H.; Tang, Y.-Z.; Fan, X.-W.; Peng, X.-L.; Han, R.-R.; Li, Y.-K.; Wang, F.-X. Two Manganese(II)-Based Hybrid Multifunctional Phase Transition Materials with Strong Photoluminescence, High Quantum Yield, and Switchable Dielectric Properties: (C6NH16)2MnBr4 and (C7NH18)2MnBr4. Inorg. Chem. 2022, 61, 10454–10460. [Google Scholar] [CrossRef] [PubMed]
  23. You, X.; Wang, L.; Peng, Z.; Wei, Z. Two manganese chloride organic–inorganic hybrid compounds showing strong luminescence and switchable phase transition and dielectric anomaly. New J. Chem. 2022, 46, 20005–20009. [Google Scholar] [CrossRef]
  24. Davydova, M.P.; Meng, L.; Rakhmanova, M.I.; Bagryanskaya, I.Y.; Sulyaeva, V.S.; Meng, H.; Artem’ev, A.V. Highly Emissive Chiral Mn(II) Bromide Hybrids for UV-Pumped Circularly Polarized LEDs and Scintillator Image Applications. Adv. Opt. Mater. 2023, 11, 2202811. [Google Scholar] [CrossRef]
  25. Wei, X.; He, J.; Zhu, Y.; Qin, Z.; Zheng, G.; Zhang, R.; Mo, S.; Wang, J.; Yao, D.; Lin, B.; et al. Synthesis, structure, and photoelectric properties of a novel zero-dimensional organic-inorganic hybrid perovskite (C6H9N2)2MnI4. Opt. Mater. 2023, 136, 113360. [Google Scholar] [CrossRef]
  26. Wu, L.-K.; Zou, Q.-H.; Yao, H.-Q.; Ye, H.-Y.; Li, J.-R. Zero-dimensional organic–inorganic hybrid manganese bromide with coexistence of dielectric–thermal double switches and efficient photoluminescence. Dalton Trans. 2023, 52, 11558–11564. [Google Scholar] [CrossRef]
  27. Wang, M.; Wang, X.; Zhang, B.; Li, F.; Meng, H.; Liu, S.; Zhao, Q. Chiral hybrid manganese(II) halide clusters with circularly polarized luminescence for X-ray imaging. J. Mater. Chem. C 2023, 11, 3206–3212. [Google Scholar] [CrossRef]
  28. Li, W.; Zhou, Z.; Wang, C.; Li, Y.; Kurosawa, S.; Ren, G.; OuYang, X.; Wu, Y. Red-Emitting Organic–Inorganic Hybrid Manganese(II)Halides for X-Ray Imaging. Adv. Sensor Res. 2023, 2, 2200083. [Google Scholar] [CrossRef]
  29. Dong, G.; Hu, B.; Chen, C.; Yu, H.; Han, Q.; Wu, W. Two Organic–Inorganic Hybrid Manganese Bromides with Highly Efficient Emission toward White LEDs. Inorg. Chem. 2024, 63, 20830–20839. [Google Scholar] [CrossRef]
  30. Wu, H.; Zeng, R.; Tang, Z.; Bai, Y.; Zhang, S.; Dai, Y. Efficient green emission in organic-inorganic hybrid (C13H22N)2MnCl4 metal halides with large Mn–Mn spacing distance. J. Lumin. 2024, 267, 120403. [Google Scholar] [CrossRef]
  31. Cheng, H.; Hu, X.; Cao, C.; Xu, H.; Liu, X.; Li, X.; Wang, H.; Huang, K.; Yang, W.; Wang, D.; et al. Manganese-manganese distance modulation in zero-dimensional hybrid manganese halides using pyridine quaternary ammonium salts as templates. Mater. Today Chem. 2024, 42, 102343. [Google Scholar] [CrossRef]
  32. Yu, X.; Zhong, S.; Guo, Z.; Guan, J.; Tang, H.; He, X.; Chen, Y.; Pan, S. Switchable circularly polarized luminescent Mn-based hybrid metal halides. J. Mater. Chem. C 2025, 13, 2190–2197. [Google Scholar] [CrossRef]
  33. Goodgame, D.M.L.; Cotton, F.A. Phosphine oxide complexes. Part V. Tetrahedral complexes of manganese(II) containing triphenylphosphine oxide, and triphenylarsine oxide as ligands. J. Chem. Soc. 1961, 3735–3741. [Google Scholar] [CrossRef]
  34. Chandra, B.P.; Kaza, B.R. Mechanoluminescence, electroluminescence and high-pressure photoluminescence of Mn(Ph3PO)2Br2. J. Lumin. 1982, 27, 101–107. [Google Scholar] [CrossRef]
  35. Cotton, F.A.; Daniels, L.M.; Huang, P. Correlation of structure and triboluminescence for tetrahedral manganese(II) compounds. Inorg. Chem. 2001, 40, 3576–3578. [Google Scholar] [CrossRef]
  36. Tang, Y.-Y.; Wang, Z.-X.; Li, P.-F.; You, Y.-M.; Stroppa, A.; Xiong, R.-G. Brilliant triboluminescence in a potential organic–inorganic hybrid ferroelectric: (Ph3PO)2MnBr2. Inorg. Chem. Front. 2017, 4, 154–159. [Google Scholar] [CrossRef]
  37. Huang, X.; Qin, Y.; She, P.; Meng, H.; Liu, S.; Zhao, Q. Functionalized triphenylphosphine oxide-based manganese(II) complexes for luminescent printing. Dalton Trans. 2021, 50, 8831–8836. [Google Scholar] [CrossRef]
  38. Berezin, A. Birefringence and polarized luminescence of a manganese(II) chloride–triphenylphosphine oxide compound: Application in LEDs and photolithography. Mater. Chem. Front. 2023, 7, 2475–2483. [Google Scholar] [CrossRef]
  39. Chen, J.; Zhang, Q.; Zheng, F.-K.; Liu, Z.-F.; Wang, S.-H.; Wu, A.-Q.; Guo, G.-C. Intense photo- and tribo-luminescence of three tetrahedral manganese(II) dihalides with chelating bidentate phosphine oxide ligand. Dalton Trans. 2015, 44, 3289–3294. [Google Scholar] [CrossRef]
  40. Wu, Y.; Zhang, X.; Zhang, Y.-Q.; Yang, M.; Chen, Z.-N. Achievement of ligand-field induced thermochromic luminescence via two-step single-crystal to single-crystal transformations. Chem. Commun. 2018, 54, 13961–13964. [Google Scholar] [CrossRef]
  41. Wu, Y.; Zhang, X.; Xu, L.-J.; Yang, M.; Chen, Z.-N. Luminescent vapochromism due to a change of the ligand field in a one-dimensional manganese(II) coordination polymer. Inorg. Chem. 2018, 57, 9175–9181. [Google Scholar] [CrossRef]
  42. Bortoluzzi, M.; Castro, J.; Enrichi, F.; Vomiero, A.; Busato, M.; Huang, W. Green-emitting manganese(II) complexes with phosphoramide and phenylphosphonic diamide ligands. Inorg. Chem. Commun. 2018, 92, 145–150. [Google Scholar] [CrossRef]
  43. Bortoluzzi, M.; Castro, J.; Trave, E.; Dallan, D.; Favaretto, S. Orange-emitting manganese(II) complexes with chelating phosphine oxides. Inorg. Chem. Commun. 2018, 90, 105–107. [Google Scholar] [CrossRef]
  44. Berezin, A.S.; Samsonenko, D.G.; Brel, V.K.; Artem’ev, A.V. “Two-in-one” organic–inorganic hybrid MnII complexes exhibiting dual-emissive phosphorescence. Dalton Trans. 2018, 47, 7306–7315. [Google Scholar] [CrossRef]
  45. Bortoluzzi, M.; Castro, J. Dibromomanganese(II) complexes with hexamethylphosphoramide and phenylphosphonic bis(amide) ligands. J. Coord. Chem. 2019, 72, 309–327. [Google Scholar] [CrossRef]
  46. Berezin, A.S.; Davydova, M.P.; Bagryanskaya, I.Y.; Artyushin, O.I.; Brel, V.K.; Artem’ev, A.V. A red-emitting Mn(II)-based coordination polymer build on 1,2,4,5- tetrakis(diphenylphosphinyl)benzene, Inorg. Chem. Commun. 2019, 107, 107473. [Google Scholar] [CrossRef]
  47. Qin, Y.; Tao, P.; Gao, L.; She, P.; Liu, S.; Li, X.; Li, F.; Wang, H.; Zhao, Q.; Miao, Y.; et al. Designing Highly Efficient Phosphorescent Neutral Tetrahedral Manganese(II) Complexes for Organic Light-Emitting Diodes. Adv. Opt. Mater. 2019, 7, 1801160. [Google Scholar] [CrossRef]
  48. Artem’ev, A.V.; Davydova, M.P.; Berezin, A.S.; Brel, V.K.; Morgalyuk, V.P.; Bagryanskaya, I.Y.; Samsonenko, D.G. Luminescence of the Mn2+ ion in non-Oh and Td coordination environments: The missing case of square pyramid. Dalton Trans. 2019, 48, 16448–16456. [Google Scholar] [CrossRef]
  49. Bortoluzzi, M.; Castro, J.; Gobbo, A.; Ferraro, V.; Pietrobon, L.; Antoniutti, S. Tetrahedral photoluminescent manganese(II) halide complexes with 1,3-dimethyl-2-phenyl-1,3-diazaphospholidine-2-oxide as a ligand. New J. Chem. 2020, 44, 571–579. [Google Scholar] [CrossRef]
  50. Bortoluzzi, M.; Castro, J.; Gobbo, A.; Ferraro, V.; Pietrobon, L. Light harvesting indolyl-substituted phosphoramide ligand for the enhancement of Mn(II) luminescence. Dalton Trans. 2020, 49, 7525–7534. [Google Scholar] [CrossRef]
  51. Davydova, M.P.; Bauer, I.A.; Brel, V.K.; Rakhmanova, M.I.; Bagryanskaya, I.Y.; Artem’ev, A.V. Manganese(II) thiocyanate complexes with bis(phosphine oxide) ligands: Synthesis and excitation wavelength-dependent multicolor luminescence. Eur. J. Inorg. Chem. 2020, 2020, 695–703. [Google Scholar] [CrossRef]
  52. Artem’ev, A.V.; Davydova, M.P.; Berezin, A.S.; Sukhikh, T.S.; Samsonenko, D.G. Photo- and triboluminescent robust 1D polymers made of Mn(II) halides and meta-carborane based bis(phosphine oxide). Inorg. Chem. Front. 2021, 8, 2261–2270. [Google Scholar] [CrossRef]
  53. Artem’ev, A.V.; Davydova, M.P.; Rakhmanova, M.I.; Bagryanskaya, I.Y.; Pishchur, D.P. A family of Mn(II) complexes exhibiting strong photo- and triboluminescence as well as polymorphic luminescence. Inorg. Chem. Front. 2021, 8, 3767–3774. [Google Scholar] [CrossRef]
  54. Meng, H.; Zhu, W.; Li, F.; Huang, X.; Qin, Y.; Liu, S.; Yang, Y.; Huang, W.; Zhao, Q. Highly emissive and stable five-coordinated manganese(II) complex for X-ray imaging. Laser Photonics Rev. 2021, 15, 2100309. [Google Scholar] [CrossRef]
  55. Bortoluzzi, M.; Castro, J.; Di Vera, A.; Palù, A.; Ferraro, V. Manganese(II) bromo- and iodo-complexes with phosphoramidate and phosphonate ligands: Synthesis, characterization and photoluminescence. New J. Chem. 2021, 45, 12871–12878. [Google Scholar] [CrossRef]
  56. Bortoluzzi, M.; Ferraro, V.; Castro, J. Synthesis and photoluminescence of manganese(II) naphtylphosphonic diamide complexes. Dalton Trans. 2021, 50, 3132–3136. [Google Scholar] [CrossRef]
  57. Bortoluzzi, M.; Castro, J.; Ferraro, V. Dual emission from Mn(II) complexes with carbazolyl-substituted phosphoramides. Inorg. Chim. Acta 2022, 536, 120896. [Google Scholar] [CrossRef]
  58. Ferraro, V.; Castro, J.; Agostinis, L.; Bortoluzzi, M. Dual-emitting Mn(II) and Zn(II) halide complexes with 9,10-dihydro-9-oxa-10-phosphaphenanthrene-10-oxide as ligand. Inorg. Chim. Acta 2023, 545, 121285. [Google Scholar] [CrossRef]
  59. Tan, G.-H.; Chen, Y.-N.; Chuang, Y.-T.; Lin, H.-C.; Hsieh, C.-A.; Chen, Y.-S.; Lee, T.-Y.; Miao, W.-C.; Kuo, H.-C.; Chen, L.-Y.; et al. Highly Luminescent Earth-Benign Organometallic Manganese Halide Crystals with Ultrahigh Thermal Stability of Emission from 4 to 623 K. Small 2023, 19, 2205981. [Google Scholar] [CrossRef]
  60. Davydova, M.P.; Meng, L.; Rakhmanova, M.I.; Jia, Z.; Berezin, A.S.; Bagryanskaya, I.Y.; Lin, Q.; Meng, H.; Artem’ev, A.V. Strong Magnetically-Responsive Circularly Polarized Phosphorescence and X-Ray Scintillation in Ultrarobust Mn(II)–Organic Helical Chains. Adv. Mater. 2023, 35, 2303611. [Google Scholar] [CrossRef]
  61. Zhou, Z.; Meng, H.; Li, F.; Jiang, T.; Yang, Y.; Liu, S.; Zhao, Q. Highly Luminescent Nonclassical Binuclear Manganese(II) Complex Scintillators for Efficient X-ray Imaging. Inorg. Chem. 2023, 62, 5279–5736. [Google Scholar] [CrossRef] [PubMed]
  62. Ferraro, V.; Castro, J.; Bortoluzzi, M. Luminescent Behavior of Zn(II) and Mn(II) Halide Derivatives of 4-Phenyldinaphtho[2,1-d:1′,2′-f][1,3,2]dioxaphosphepine 4-Oxide and Single-Crystal X-ray Structure Determination of the Ligand. Molecules 2024, 29, 239. [Google Scholar] [CrossRef] [PubMed]
  63. Abel, E.W.; Mucklejohn, S.A. The chemistry of phosphinimines. Phosphorus Sulfur Silicon Relat. Elem. 1981, 9, 235–266. [Google Scholar] [CrossRef]
  64. Appel, R.; Schaaff, R. Triphenylphosphinimin-Komplexe Von Übergangselementen. Z. Naturforsch. B 1961, 16, 405. [Google Scholar] [CrossRef]
  65. Hieber, W.; Winter, E.; Schubert, E. Reaktionen des Vanadinhexacarbonyls mit verschiedenartigen Basen und die Säurefunktion von Vanadincarbonylwasserstoff-Verbindungen, II. Chem. Ber. 1962, 95, 3070–3076. [Google Scholar] [CrossRef]
  66. Seidel, W. Complexes of Iminophosphoranes with Metal Salts. Angew. Chem. Int. Ed. 1965, 4, 785. [Google Scholar] [CrossRef]
  67. Bock, H.; Dieck, H.T. Triphenylphosphin-imin, -oxid und -äthylen als Liganden in Molybdän[0]- und Wolfram[0]-Carbonyl-Komplexen. Z. Naturforsch. B 1966, 21, 739–746. [Google Scholar] [CrossRef]
  68. Fukui, M.; Itoh, K.; Ishii, Y. Iminophosphorane Complexes of Palladium(II). Bull. Chem. Soc. Jpn. 1975, 48, 2044–2046. [Google Scholar] [CrossRef]
  69. Abel, E.W.; Mucklejohn, S.A.; Cameron, T.S.; Cordes, R.E. The Dimeric Phosphinimine Complex [CdI2(HN:PPh3)2]2 Containing N-H···I Hydrogen Bonds. Z. Naturforsch. B 1978, 33, 339–340. [Google Scholar] [CrossRef]
  70. Dilworth, J.R.; de Liefde Meijer, H.J.; Teuben, J.H. Cyclopentadienyl titanium complexes with aryldiazenido- or phosphiniminato-ligands. J. Organomet. Chem. 1978, 159, 47–52. [Google Scholar] [CrossRef]
  71. Abel, E.W.; Mucklejohn, S.A. Some triphenylphosphinimine complexes of zinc, cadmium, mercury, rhodium and palladium. Inorg. Chim. Acta 1979, 37, 107–111. [Google Scholar] [CrossRef]
  72. Fenske, D.; Böhm, E.; Dehnicke, K.; Strähle, J. N-Trimethylsilyl-iminotriphenylphosphoran-Kupfer(II)-chlorid, [Me3SiNPPh3 · CuCl2]2 Synthese und Kristallstruktur. Z. Naturforsch. B 1988, 43, 1–4. [Google Scholar] [CrossRef]
  73. Maurer, A.; Fenske, D.; Beck, J.; Strähle, J.; Böhm, E.; Dehnicke, K. Kupfer(I)-chlorid-Addukte von Phosphoraniminen Die Kristallstrukturen von Ph3PNPh·CuCl und von 2,3-Bis(triphenylphosphoranylidenamino)maIeinsäure-N-methylimid- Kupfer(I)-chlorid. Z. Naturforsch. B 1988, 43, 5–11. [Google Scholar] [CrossRef]
  74. Imhoff, P.; Elsevier, C.J.; Stam, C.H. Iminophosphorane complexes of rhodium(I) and X-ray crystal structure of [Rh(COD)Cl(Et3P=N-p-tolyl)]. Inorg. Chim. Acta 1990, 175, 209–216. [Google Scholar] [CrossRef]
  75. Meyer Zu Köcker, R.; Frenzen, G.; Neumüller, B.; Dehnicke, K.; Magull, J. Synthese und Kristallstrukturen der Phosphanimin-Komplexe MCl2(Me3SiNPMe3)2 mit M = Zn und Co, und CoCl2(HNPMe3)2. Z. Anorg. Allg. Chem. 1994, 620, 431–437. [Google Scholar] [CrossRef]
  76. Miekisch, T.; Mai, H.J.; Meyer zu Köcker, R.; Dehnicke, K.; Magull, J.; Goesmann, H. Silylierte Phosphanimin-Komplexe von Chrom(II), Palladium(II) und Kupfer(II). Die Kristallstrukturen von [CrCl2(Me3SiNPMe3)2], [PdCl2(Me3SiNPEt3)2] und [CuCl2(Me3SiNPMe3)]2. Z. Anorg. Allg. Chem. 1996, 622, 583–588. [Google Scholar] [CrossRef]
  77. Braun, T.P.; Gutsch, P.A.; Zimmer, H. Complexes of N-Aryltriphenylphosphinimines with Mercury(II) Halides. Z. Naturforsch. B 1999, 54, 858–862. [Google Scholar] [CrossRef]
  78. Wheaton, C.A.; Ireland, B.J.; Hayes, P.G. Synthesis and Structural Diversity of ZnCl2 and Zn(C6F5)2 Complexes of a Phosphinimine Ligand. Z. Anorg. Allg. Chem. 2011, 637, 2111–2119. [Google Scholar] [CrossRef]
  79. Venderbosch, B.; Oudsen, J.-P.H.; van der Vlugt, J.I.; Korstanje, T.J.; Tromp, M. Cationic Copper Iminophosphorane Complexes as CuAAC Catalysts: A Mechanistic Study. Organometallics 2020, 39, 3480–3489. [Google Scholar] [CrossRef]
  80. Miller, J.S.; Visscher, M.O.; Caulton, K.G. Spectra and Structure of Mo2(CO)6(μ-Ph3PNH)3. A New Bonding Mode for Phosphinimines. Inorg. Chem. 1974, 13, 1632–1639. [Google Scholar] [CrossRef]
  81. Appel, R.; Volz, P. 1,2-Bis-triphenylphosphoranylidenamino-äthan als zweizähniger Ligand in Übergangsmetallkomplexen. Z. Anorg. Allg. Chem. 1975, 413, 45–50. [Google Scholar] [CrossRef]
  82. Imhoff, P.; van Asselt, R.; Elsevier, C.J.; Zoutberg, M.C.; Stam, C.H. Reactions of bis(iminophosphoranyl)methanes with chloro-bridged rhodium or iridium dimers giving complexes in which the ligand is coordinated either as a σ-N,σ-N′ or as a σ-N, σ-C chelate. X-ray crystal structure of the σ-N, σ-N′ Rh(I) complex Rh{(4-CH3-C6H4-N=PPh2)2CH2}(COD)]PF6. Inorg. Chim. Acta 1991, 184, 73–87. [Google Scholar] [CrossRef]
  83. Avis, M.W.; Vrieze, K.; Kooijman, H.; Veldman, N.; Spek, A.L.; Elsevier, C.J. Selective Formation of Four-Membered Metallacyclic Pt-N-P-C Compounds from Reactions of Bis((N-arylimino)phosphoranyl)methanes with Halide-Bridged Platinum(II) Phosphine Dimers. X-ray Crystal Structures of [PtCl(PMe2Ph){CH(PPh2=NC6H4-4-CH3)(PPh2NHC6H4-4-CH3)}]+Cl and [PtCl(PEt3){CH(PPh2=NC6H4-4-CH3)(PPh2NHC6H4-4-CH3)}]+[PtCl3(PEt3)]. Inorg. Chem. 1995, 34, 4092–4105. [Google Scholar] [CrossRef]
  84. Reetz, M.T.; Bohres, E.; Goddard, R. Chiral diiminophosphoranes: A new class of ligands for enantioselective transition metal catalysis. Chem. Commun. 1998, 935–936. [Google Scholar] [CrossRef]
  85. Al-Benna, S.; Sarsfield, M.J.; Thornton-Pett, M.; Ormsby, D.L.; Maddox, P.J.; Brès, P.; Bochmann, M. Sterically hindered iminophosphorane complexes of vanadium, iron, cobalt and nickel: A synthetic, structural and catalytic study. J. Chem. Soc. Dalton Trans. 2000, 4247–4257. [Google Scholar] [CrossRef]
  86. Steiner, A.; Zacchini, S.; Richards, P.I. From neutral iminophosphoranes to multianionic phosphazenates. The coordination chemistry of imino–aza-P(V) ligands. Coord. Chem. Rev. 2002, 227, 193–216. [Google Scholar] [CrossRef]
  87. Jain, A.; Karmakar, H.; Roesky, P.W.; Panda, T.K. Role of Bis(phosphinimino)methanides as Universal Ligands in the Coordination Sphere of Metals across the Periodic Table. Chem. Rev. 2023, 123, 13323–13373. [Google Scholar] [CrossRef]
  88. Evans, D.J.; Hill, M.S.; Hitchcock, P.B. Tuning low-coordinate metal environments: High spin d5–d7 complexes supported by bis(phosphinimino)methyl ligation. Dalton Trans. 2003, 570–574. [Google Scholar] [CrossRef]
  89. Hill, M.S.; Hitchcock, P.B. Three-coordinate bis(phosphinimino)methanide derivatives of ‘open shell’ [M(II)] (M = Mn, Fe, Co) transition metals. J. Organomet. Chem. 2004, 689, 3163–3167. [Google Scholar] [CrossRef]
  90. García-Álvarez, J.; García-Garrido, S.E.; Cadierno, V. Iminophosphorane–phosphines: Versatile ligands for homogeneous catalysis. J. Organomet. Chem. 2014, 751, 792–808. [Google Scholar] [CrossRef]
  91. Frik, M.; Jiménez, J.; Vasilevski, V.; Carreira, M.; de Almeida, A.; Gascón, E.; Benoit, F.; Sanaú, M.; Casini, A.; Contel, M. Luminescent iminophosphorane gold, palladium and platinum complexes as potential anticancer agents. Inorg. Chem. Front. 2014, 1, 231–241. [Google Scholar] [CrossRef] [PubMed]
  92. Mai, H.-J.; Wocadlo, S.; Massa, W.; Weller, F.; Dehnicke, K.; Maichle-Mössmer, C.; Strähle, J. Phosphanimin-Komplexe von Mangan(II)halogeniden. Die Kristallstrukturen von [MnCl2(Me3SiNPEt3)]2, [MnI2(Me3SiNPEt3)2] und [MnI2(Me2Si(NPEt3)2)]. Z. Naturforsch. B 1995, 50, 1215–1221. [Google Scholar] [CrossRef]
  93. Foster, R.S.; Jakobi, H.; Harrity, J.P.A. A General and Regioselective Synthesis of 5-Trifluoromethyl-pyrazoles. Org. Lett. 2012, 14, 4858–4861. [Google Scholar] [CrossRef] [PubMed]
  94. Liu, Z.; Lu, T.; Chen, Q. An sp-hybridized all-carboatomic ring, cyclo[18]carbon: Electronic structure, electronic spectrum, and optical nonlinearity. Carbon 2020, 165, 461–467. [Google Scholar] [CrossRef]
  95. Yang, H.; Bard, A.J. The application of fast scan cyclic voltammetry. Mechanistic study of the initial stage of electropolymerization of aniline in aqueous solutions. J. Electroanal. Chem. 1992, 339, 423–449. [Google Scholar] [CrossRef]
  96. Sakai, Y.; Sagara, Y.; Nomura, H.; Nakamura, N.; Suzuki, Y.; Miyazaki, H.; Adachi, C. Zinc complexes exhibiting highly efficient thermally activated delayed fluorescence and their application to organic light-emitting diodes. Chem. Commun. 2015, 51, 3181–3184. [Google Scholar] [CrossRef]
  97. Berezin, A.S.; Vinogradova, K.A.; Krivopalov, V.P.; Nikolaenkova, E.B.; Plyusnin, V.F.; Kupryakov, A.S.; Pervukhina, N.V.; Naumov, D.Y.; Bushuev, M.B. Excitation-Wavelength-Dependent Emission and Delayed Fluorescence in a Proton-Transfer System. Chem. Eur. J. 2018, 24, 12790–12795. [Google Scholar] [CrossRef]
  98. Xiong, J.; Li, K.; Teng, T.; Chang, X.; Wei, Y.; Wu, C.; Yang, C. Dinuclear ZnII Complexes Exhibiting Thermally Activated Delayed Fluorescence and Luminescence Polymorphism. Chem. Eur. J. 2020, 26, 6887–6893. [Google Scholar] [CrossRef]
  99. Ferraro, V.; Baggio, F.; Castro, J.; Bortoluzzi, M. Green Phosphorescent Zn(II) Halide Complexes with N,N,N′,N′-tetramethyl-P-indol-1-ylphosphonic Diamide as Ligand. Eur. J. Inorg. Chem. 2022, 2022, e202200119. [Google Scholar] [CrossRef]
  100. Yersin, H. Highly Efficient OLEDs: Materials Based on Thermally Activated Delayed Fluorescence; Wiley-VCH: Weinheim, Germany, 2018; pp. 4–46. [Google Scholar]
  101. Bertini, I.; Luchinat, C. The hyperfine shift. Coord. Chem. Rev. 1996, 150, 29–75. [Google Scholar] [CrossRef]
  102. Liu, X.; Dronskowski, R.; Glaum, R.; Tchougréeff, A.L. Experimental and Quantum-Chemical Investigations of the UV/Vis Absorption Spectrum of Manganese Carbodiimide, MnNCN. Z. Anorg. Allg. Chem. 2010, 636, 343–348. [Google Scholar] [CrossRef]
  103. Wrighton, M.; Ginley, D. Excited state decay of tetrahalomanganese(II) complexes. Chem. Phys. 1974, 4, 295–299. [Google Scholar] [CrossRef]
  104. Gusev, A.; Kiskin, M.; Braga, E.; Zamnius, E.; Lisichnikova, A.; Linert, W. New highly efficient phosphorescent manganese halides as green and blue emitters. Dalton Trans. 2025, 54, 3335–3345. [Google Scholar] [CrossRef] [PubMed]
  105. Alvarez, S.; Alemany, P.; Casanova, D.; Cirera, J.; Llunell, M.; Avnir, D. Shape maps and polyhedral interconversion paths in transition metal chemistry. Coord. Chem. Rev. 2005, 249, 1693–1708. [Google Scholar] [CrossRef]
  106. Cirera, J.; Alemany, P.; Alvarez, S. Mapping the Stereochemistry and Symmetry of Tetracoordinate Transition-Metal Complexes. Chem. Eur. J. 2004, 10, 190–207. [Google Scholar] [CrossRef]
  107. Yang, L.; Powell, D.R.; Houser, R.P. Structural variation in copper(I) complexes with pyridylmethylamide ligands: Structural analysis with a new four-coordinate geometry index, τ4. Dalton Trans. 2007, 955–964. [Google Scholar] [CrossRef]
  108. Okuniewski, A.; Rosiak, D.; Chojnacki, J.; Becker, B. Coordination polymers and molecular structures among complexes of mercury(II) halides with selected 1-benzoylthioureas. Polyhedron 2015, 90, 47–57. [Google Scholar] [CrossRef]
  109. Böhm, E.; Dehnicke, K.; Beck, J.; Hiller, W.; Strähle, J.; Maurer, A.; Fenske, D. Die Kristallstrukturen von Ph3PNPh, [Ph3PN(H)Ph][AuI2] und von 2,3-Bis(triphenylphosphoranimino)maleinsäure-N-methylimid. Z. Naturforsch. B 1988, 43, 138–144. [Google Scholar] [CrossRef]
  110. Jiang, C.; Stephan, D.W. Phosphinimine–borane combinations in frustrated Lewis pair chemistry. Dalton Trans. 2013, 42, 630–637. [Google Scholar] [CrossRef]
  111. Armarego, W.L.F.; Chai, C.L.L. Purification of Laboratory Chemicals, 4th ed.; Butterworth-Heinemann: Oxford, UK, 1996; pp. 67–68, 215–216. [Google Scholar]
  112. Lindsay, R.O.; Allen, C.F.H. Phenyl azide. Org. Synth. 1942, 22, 96–97. [Google Scholar] [CrossRef]
  113. Deucher, N.C.; Silva de Souza, R., Jr.; Borges, E.M. Teaching Precipitation Titration Methods: A Statistical Comparison of Mohr, Fajans, and Volhard Techniques. J. Chem. Educ. 2025, 102, 364–371. [Google Scholar] [CrossRef]
  114. Bain, G.A.; Berry, J.F. Diamagnetic corrections and Pascal’s constants. J. Chem. Educ. 2008, 85, 532–536. [Google Scholar] [CrossRef]
  115. Elgrishi, N.; Rountree, K.J.; McCarthy, B.D.; Rountree, E.S.; Eisenhart, T.T.; Dempsey, J.L. A Pratical Beginner’s Guide to Cyclic Voltammetry. J. Chem. Educ. 2018, 95, 197–206. [Google Scholar] [CrossRef]
  116. APEX4 v.2022.1-1, SAINT v.8.40B, XPREP v.2014/2, SADABS-2016/2, Bruker AXS Inc.: Madison, WI, USA, 2022.
  117. McArdle, P. Oscail, a program package for small-molecule single-crystal crystallography with crystal morphology prediction and molecular modelling. J. Appl. Crystallogr. 2017, 50, 320–326. [Google Scholar] [CrossRef]
  118. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A 2015, 71, 3–8. [Google Scholar] [CrossRef]
  119. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C 2015, 71, 3–8. [Google Scholar] [CrossRef]
  120. Spek, A.L. checkCIF validation ALERTS: What they mean and how to respond. Acta Crystallogr. Sect. E 2020, 76, 1–11. [Google Scholar] [CrossRef]
  121. Grimme, S.; Brandenburg, J.G.; Bannwarth, C.; Hansen, A. Consistent structures and interactions by density functional theory with small atomic orbital basis sets. J. Chem. Phys. 2015, 143, 054107. [Google Scholar] [CrossRef]
  122. Kruse, H.; Grimme, S. A geometrical correction for the inter- and intra-molecular basis set superposition error in Hartree-Fock and density functional theory calculations for large systems. J. Chem. Phys. 2012, 136, 154101. [Google Scholar] [CrossRef]
  123. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  124. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  125. Neese, F. Software update: The ORCA program system -Version 5.0. WIREs Comput. Mol. Sci. 2022, 12, e1616. [Google Scholar] [CrossRef]
  126. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  127. Lu, T. A comprehensive electron wavefunction analysis toolbox for chemists, Multiwfn. J. Chem. Phys. 2024, 161, 082503. [Google Scholar] [CrossRef]
Scheme 1. Manganese(II) iminophosphorane derivatives described in this work.
Scheme 1. Manganese(II) iminophosphorane derivatives described in this work.
Molecules 30 01319 sch001
Figure 1. Normalised UV–VIS spectrum (violet dashed line, CH2Cl2 solution, r.t.) and normalised PL (green line, λex = 300 nm, solid, r.t.) and PLE (violet line, λex = 510 nm, solid, r.t.) spectra of NPh=PPh3. CIE 1931 chromaticity coordinates: x = 0.261, y = 0.393. First-singlet excited state geometry of NPh=PPh3 (P, orange; N, blue; C, grey; H, white) with hole (yellow) and electron (green) isosurfaces for the S1S0 transition. Emission is the reverse process.
Figure 1. Normalised UV–VIS spectrum (violet dashed line, CH2Cl2 solution, r.t.) and normalised PL (green line, λex = 300 nm, solid, r.t.) and PLE (violet line, λex = 510 nm, solid, r.t.) spectra of NPh=PPh3. CIE 1931 chromaticity coordinates: x = 0.261, y = 0.393. First-singlet excited state geometry of NPh=PPh3 (P, orange; N, blue; C, grey; H, white) with hole (yellow) and electron (green) isosurfaces for the S1S0 transition. Emission is the reverse process.
Molecules 30 01319 g001
Figure 2. Normalised UV–VIS (violet dashed line, CH2Cl2 solution, r.t.), PL (green line, λex = 350 nm, solid, r.t.) and PLE (violet line, λex = 510 nm, solid, r.t.) spectra of [ZnBr2(NPh=PPh3)2]. CIE 1931 chromaticity diagram; chromaticity coordinates: x = 0.294, y = 0.418.
Figure 2. Normalised UV–VIS (violet dashed line, CH2Cl2 solution, r.t.), PL (green line, λex = 350 nm, solid, r.t.) and PLE (violet line, λex = 510 nm, solid, r.t.) spectra of [ZnBr2(NPh=PPh3)2]. CIE 1931 chromaticity diagram; chromaticity coordinates: x = 0.294, y = 0.418.
Molecules 30 01319 g002
Figure 3. Normalised PL (solid lines, λex = 358 nm, solid, r.t.) and PLE (dashed lines, λem = 560 nm) spectra of [MnX2(NPh=PPh3)2] (solid, r.t.). Luminescence decay curves (MCS, solid state, r.t., λex = 290 nm, λem = 535 nm). CIE 1931 chromaticity coordinates: x = 0.294, y = 0.572 (X = Cl); x = 0.317, y = 0.550 (X = Br); x = 0.356, y = 0.620 (X = I). PL spectra of [MnCl2(NPh=PPh3)2] recorded at different excitation wavelengths are shown in Figure S9.
Figure 3. Normalised PL (solid lines, λex = 358 nm, solid, r.t.) and PLE (dashed lines, λem = 560 nm) spectra of [MnX2(NPh=PPh3)2] (solid, r.t.). Luminescence decay curves (MCS, solid state, r.t., λex = 290 nm, λem = 535 nm). CIE 1931 chromaticity coordinates: x = 0.294, y = 0.572 (X = Cl); x = 0.317, y = 0.550 (X = Br); x = 0.356, y = 0.620 (X = I). PL spectra of [MnCl2(NPh=PPh3)2] recorded at different excitation wavelengths are shown in Figure S9.
Molecules 30 01319 g003
Figure 4. Molecular structure of [NHPh=PPh3]2[MnBr4]. Hydrogen atoms are not drawn, except for those on the iminium nitrogen atoms.
Figure 4. Molecular structure of [NHPh=PPh3]2[MnBr4]. Hydrogen atoms are not drawn, except for those on the iminium nitrogen atoms.
Molecules 30 01319 g004
Figure 5. Some of the hydrogen bonds found in [NHPh=PPh3]2[MnBr4]. Symmetry operations: i, 3/2 − x, y + 1/2, z; ii, 1 − x, y + 1/2, 1/2 - z; iii, x + 1/2, y, 1/2 − z.
Figure 5. Some of the hydrogen bonds found in [NHPh=PPh3]2[MnBr4]. Symmetry operations: i, 3/2 − x, y + 1/2, z; ii, 1 − x, y + 1/2, 1/2 - z; iii, x + 1/2, y, 1/2 − z.
Molecules 30 01319 g005
Figure 6. Normalised PL (solid lines, λex = 370 nm, solid, r.t.) and PLE (dashed lines, λem = 560 nm) spectra of [NHPh=PPh3]2[MnX4] (solid, r.t.). Three-dimensional plot of the excitation-dependent emission of [NHPh=PPh3]2[MnI4] (solid, r.t.). CIE 1931 chromaticity coordinates: x = 0.309, y = 0.620 (X = Cl); x = 0.314, y = 0.631 (X = Br); x = 0.407, y = 0.558 (X = I).
Figure 6. Normalised PL (solid lines, λex = 370 nm, solid, r.t.) and PLE (dashed lines, λem = 560 nm) spectra of [NHPh=PPh3]2[MnX4] (solid, r.t.). Three-dimensional plot of the excitation-dependent emission of [NHPh=PPh3]2[MnI4] (solid, r.t.). CIE 1931 chromaticity coordinates: x = 0.309, y = 0.620 (X = Cl); x = 0.314, y = 0.631 (X = Br); x = 0.407, y = 0.558 (X = I).
Molecules 30 01319 g006
Figure 7. Luminescence decay curves of [NHPh=PPh3]2[MnX4] (MCS, λex = 290 nm, λem = 520–535 nm; TCSPC, λex = 373 nm, λem = 565 nm). Solid state, r.t.
Figure 7. Luminescence decay curves of [NHPh=PPh3]2[MnX4] (MCS, λex = 290 nm, λem = 520–535 nm; TCSPC, λex = 373 nm, λem = 565 nm). Solid state, r.t.
Molecules 30 01319 g007
Table 1. Room-temperature photophysical data. UV–VIS data for CH2Cl2 solutions; PL, PLE, τ and Φ data for solid samples.
Table 1. Room-temperature photophysical data. UV–VIS data for CH2Cl2 solutions; PL, PLE, τ and Φ data for solid samples.
CompoundPhotoluminescence Data
NPh=PPh3UV-VIS (nm): <400, 342 sh, 298 sh, 256.
PLex = 300 nm, nm): 496 (FWHM = 4800 cm−1).
PLEem = 510 nm, nm): <450.
τex = 373 nm, λem = 415 nm, ns, TCSPC): 2.
[ZnBr2(NPh=PPh3)2]UV-VIS (nm): <400, 335 sh, 262 sh.
PLex = 350 nm, nm): 510 (FWHM = 5300 cm−1).
PLEem = 510 nm, nm): <470, max 347.
τex = 373 nm, λem = 510 nm, ns, TCSPC): 222 (av.).
τex = 290 nm, λem = 510 nm, μs, MCS): 303 (av.).
Φ: 3%.
[MnCl2(NPh=PPh3)2]UV-VIS (nm): <350, 295 sh, 276 sh, 269, 264.
PLex = 358 nm, nm): 460 sh, 530 (FWHM = 2400 cm−1).
PLEem = 560 nm, nm): 470, 446, 432, 420–300 (358 max), <295.
τex = 373 nm, λem = 410 nm, ns, TCPSC): 2.
τex = 290 nm, λem = 535 nm, μs, MCS): 2478.
[MnBr2(NPh=PPh3)2]UV-VIS (nm): <350, 295 sh, 276 sh, 268, 261.
PLex = 358 nm, nm): 528 (FWHM = 2700 cm−1).
PLEem = 560 nm, nm): 470 sh, 454, 435, 420–300 (359 max), <295.
τex = 373 nm, λem = 410 nm, ns, TCSPC): 2.
τex = 290 nm, λem = 535 nm, μs, MCS): 151.
Φ: 3%.
[MnI2(NPh=PPh3)2]UV-VIS (nm): <400, 364, 292, 276, 268.
PLex = 358 nm, nm): 538 (FWHM = 2200 cm−1).
PLEem = 560 nm, nm): 472, 447 sh, 390, 372, <360.
τex = 290 nm, λem = 535 nm, μs, MCS): 55.
Φ: 7%.
[NHPh=PPh3]2[MnCl4]UV-VIS (nm): 299 sh, 276, 269, 262.
PLex = 370 nm, nm): 532 (FWHM = 2300 cm−1).
PLEem = 560 nm, nm): 467 sh, 448, 432, 382 sh, 371 sh, 358, <295.
τex = 290 nm, λem = 530 nm, μs, MCS): 2832.
[NHPh=PPh3]2[MnBr4]UV-VIS (nm): 299 sh, 276, 269, 262.
PLex = 370 nm, nm): 529 (FWHM = 2500 cm−1).
PLEem = 560 nm, nm): 471 sh, 454, 436, 375 sh, 362, 343, 312, <295.
τex = 290 nm, λem = 535 nm, μs, MCS): 313.
Φ: 22%.
[NHPh=PPh3]2[MnI4]UV-VIS (nm): 365, 299 sh, 276, 269, 262, <255.
PLex = 370 nm, nm): 565, 525 sh.
PLex = 280 nm, nm): 537 (FWHM = 2200 cm−1).
PLEem = 560 nm, nm): 471, 446 sh, 390, 373, <355.
τex = 290 nm, λem = 520 nm, μs, MCS): 49.
τex = 373 nm, λem = 565 nm, μs, MCS): 10 (55%), 49 a (45%).
Φ: 23%.
a Fixed value in the biexponential fit.
Table 2. Selected bond lengths [Å] and angles [°] for [NHPh=PPh3]2[MnBr4].
Table 2. Selected bond lengths [Å] and angles [°] for [NHPh=PPh3]2[MnBr4].
Mn(1)-Br(1)2.5055(10)Mn(1)-Br(3)2.4916(11)
Mn(1)-Br(2)2.5263(11)Mn(1)-Br(4)2.4748(11)
P(1)-N(1)1.633(5)P(2)-N(2)1.631(5)
P(1)-C(31)1.785(5)P(2)-C(61)1.773(6)
P(1)-C(21)1.786(6)P(2)-C(71)1.783(7)
P(1)-C(41)1.795(6)P(2)-C(81)1.798(7)
N(1)-C(11)1.423(7)N(2)-C(51)1.421(8)
Br(1)-Mn(1)-Br(2)106.94(4)Br(3)-Mn(1)-Br(1)109.01(4)
Br(4)-Mn(1)-Br(1)108.17(4)Br(3)-Mn(1)-Br(2)113.74(4)
Br(4)-Mn(1)-Br(2)108.77(4)Br(4)-Mn(1)-Br(3)110.03(4)
N(1)-P(1)-C(31)111.2(2)N(2)-P(2)-C(61)104.9(3)
N(1)-P(1)-C(21)106.5(2)N(2)-P(2)-C(71)112.2(3)
N(1)-P(1)-C(41)111.8(3)N(2)-P(2)-C(81)111.6(3)
C(11)-N(1)-P(1)128.6(4)C(51)-N(2)-P(2)128.5(4)
Table 3. Hydrogen bonds for [NHPh=PPh3]2[MnBr4] [Å and °].
Table 3. Hydrogen bonds for [NHPh=PPh3]2[MnBr4] [Å and °].
D-H…A(D-H)d(H…A)d(D…A)<(DHA)
N(1)-H(1)…Br(2)0.862.763.586(5)161.9
C(12)-H(12)…Br(2)0.932.963.757(7)144.5
N(2)-H(2)…Br(3)0.862.683.520(5)164.8
C(23)-H(23)…Br(4)0.932.893.720(7)149.6
C(64i)-H(64i)…Br(4)0.932.973.658(8)132.3
C(34ii)-H(34ii)…Br(4)0.932.963.627(7)130.0
C(36iii)-H(36iii)…Br(4)0.932.913.769(6)154.1
Symmetry operations used: i, 3/2 − x, y + 1/2, z; ii, 1 − x, y + 1/2, 1/2 − z; iii, x + 1/2, y, 1/2 − z.
Table 4. Crystal data and structure refinement for [NHPh=PPh3]2[MnBr4].
Table 4. Crystal data and structure refinement for [NHPh=PPh3]2[MnBr4].
Empirical formulaC48 H42 Br4 Mn N2 P2Index ranges−18 ≤ h ≤ 21
Moiety formula(C24 H21 N P)2(Mn Br4) −18 ≤ k ≤ 18
Formula weight1083.35 −47 ≤ l ≤ 44
Temperature296(2) KReflections collected95,854
Wavelength0.71073 ÅIndependent reflections9780 [Rint = 0.0581 Rσ = 0.0382]
Crystal systemOrthorhombicReflections observed (>2σ)6622
Space groupPbcaData completeness0.999
Unit cell dimensionsa = 17.7894(8) ÅAbsorption correctionSemi-empirical from equivalents
b = 15.3223(6) ÅMax. and min. transmission0.8620 and 0.6668
c = 39.1289(17) ÅRefinement methodFull-matrix least-squares on F2
Volume10,665.5(8) Å3Data/restraints/parameters9780/0/514
Z8Goodness-of-fit on F21.048
Density (calculated)1.349 Mg/m3Final R indices [I > 2σ(I)]R1 = 0.0548
Absorption coefficient3.335 mm−1 wR2 = 0.1363
F(000)4312R indices (all data)R1 = 0.0862
Crystal size0.141 mm × 0.092 mm × 0.048 mm wR2 = 0.1508
Theta range for data collection2.082 to 25.368°Largest diff. peak and hole0.516 and −0.521 e.Å−3
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Piccolo, D.; Castro, J.; Rosa-Gastaldo, D.; Bortoluzzi, M. Luminescent Manganese(II) Iminophosphorane Derivatives. Molecules 2025, 30, 1319. https://doi.org/10.3390/molecules30061319

AMA Style

Piccolo D, Castro J, Rosa-Gastaldo D, Bortoluzzi M. Luminescent Manganese(II) Iminophosphorane Derivatives. Molecules. 2025; 30(6):1319. https://doi.org/10.3390/molecules30061319

Chicago/Turabian Style

Piccolo, Domenico, Jesús Castro, Daniele Rosa-Gastaldo, and Marco Bortoluzzi. 2025. "Luminescent Manganese(II) Iminophosphorane Derivatives" Molecules 30, no. 6: 1319. https://doi.org/10.3390/molecules30061319

APA Style

Piccolo, D., Castro, J., Rosa-Gastaldo, D., & Bortoluzzi, M. (2025). Luminescent Manganese(II) Iminophosphorane Derivatives. Molecules, 30(6), 1319. https://doi.org/10.3390/molecules30061319

Article Metrics

Back to TopTop