Next Article in Journal
High-Energy-Low-Temperature Technologies for the Synthesis of Nanoparticles: Microwaves and High Pressure
Next Article in Special Issue
Experimental and Theoretical Studies of the Factors Affecting the Cycloplatination of the Chiral Ferrocenylaldimine (SC)-[(η5-C5H5)Fe{(η5-C5H4)–C(H)=N–CH(Me)(C6H5)}]
Previous Article in Journal / Special Issue
Attachment of Luminescent Neutral “Pt(pq)(C≡CtBu)” Units to Di and Tri N-Donor Connecting Ligands: Solution Behavior and Photophysical Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Hydrido η1-Alkynyl Diplatinum Complex Obtained from a Phosphinito Phosphanido Complex and Trimethylsilylacetylene

1
Dipartimento di Ingegneria Civile, Ambientale, del Territorio, Edile e di Chimica (DICATECh), Politecnico di Bari, via Orabona 4, I-70125 Bari, Italy
2
Consiglio Nazionale delle Ricerche, Istituto di Chimica dei Composti Organometallici (ICCOM-CNR), Via Orabona 4, 70125 Bari, Italy
3
Departamento de Química, University of La Rioja, Madre de Dios 51, 26006 Logroño, Spain
*
Author to whom correspondence should be addressed.
Inorganics 2014, 2(4), 591-605; https://doi.org/10.3390/inorganics2040591
Submission received: 11 September 2014 / Revised: 13 October 2014 / Accepted: 14 October 2014 / Published: 31 October 2014
(This article belongs to the Special Issue Organoplatinum Complexes)

Abstract

:
The reaction of (trimethylsilyl)acetylene with the phosphinito phosphanido Pt(I) complex [(PHCy2)Pt(μ-PCy2){κ2P,O-μ-P(O)Cy2}Pt(PHCy2)](Pt-Pt) (1) results in the protonation of the Pt-Pt bond with the formation of the bridging hydride complex [(PHCy2)(Me3SiC≡C)Pt(μ-PCy2)(μ-H) Pt(PHCy2){κP-P(O)Cy2}](Pt-Pt) (2), which was characterized by spectroscopic, spectrometric and XRD analyses. Complex 2 exhibits in the solid state at 77 K a long-lived, weak, orange emission assigned as metal-metal to ligand charge transfer (MMLCT) (L = alkynyl) due to the presence of a very short Pt···Pt distance [2.8209(2) Å]. Reaction of 2 with etherate HBF4 results in the selective protonation of the phosphinito ligand to afford the species [(PHCy2)(Me3SiC≡C)Pt(μ-PCy2)(μ-H)Pt(PHCy2){κP-P(OH)Cy2}](Pt-Pt)[BF4] ([3]BF4).

1. Introduction

It is now well established that the mixed bridge phosphinito-phosphanido Pt(I) complex [(PHCy2)Pt1(μ-PCy2){κP,O-μ-P(O)Cy2}Pt2(PHCy2)](Pt–Pt) (1) [1] exhibits an amphiphilic behavior, in that it reacts with both nucleophiles (with initial attack at the Pt1 center) [2] and electrophiles (with initial attack at the O atom) [3,4]. In particular, reaction of 1 with HX Brønsted acids affords, with the exception of HF, [4] dinuclear compounds endowed with a bridging hydride, a bridging dicyclohexylphosphanide and a covalent bond between Pt1 and the X residue (Scheme 1).
Scheme 1. Reactivity of Complex 1 with Brønsted acids.
Scheme 1. Reactivity of Complex 1 with Brønsted acids.
Inorganics 02 00591 g006
In the framework of platinum chemistry, the interaction of terminal alkynes with mono- or bi-nuclear complexes has been an intense area of research in connection with their involvement in many insertion, coupling and polymerizations reactions [5,6,7,8,9,10]. On the other hand, platinum alkynyl complexes have been attracting widespread interest for several years, mainly due to their structural variety, remarkable reactivity and photophysical properties [11,12,13,14,15,16,17,18].
Therefore, we decided to explore the reactivity of the Pt(I) Complex 1 with substituted acetylenes possessing the C≡C–H moiety able to protonate the diplatinum complex and to form new Pt-C bonds. In this paper, we describe for the first time the reaction between a platinum(I) complex and alkynes, which led to the synthesis of the acetylide complex [(PHCy2)(Me3Si-C≡C)Pt(μ-PCy2){κP-P(O)Cy2}Pt(PHCy2)](Pt–Pt) (2).

2. Results and Discussion

2.1. Synthesis and Spectroscopic Features of the Hydride/Acetylide Complex 2

The Pt(I) Complex 1 transformed quantitatively into the bridging hydride complex [(PHCy2)(η1-Me3SiC≡C)Pt(μ-PCy2)(μ-H) Pt(PHCy2){κP-P(O)Cy2}](Pt-Pt) (2) by reaction in n-hexane with an equimolar quantity of (trimethylsilyl)acetylene (Scheme 2). This reactivity is similar to that observed with phenol, trifluoroethanol or dicyclohexylphosphane, [2] and the product is analogous to that obtained in the first stage of reaction of 1 with HCl or HBr (Scheme 1). [3] However, differently from the analogous chlorido or bromido complexes [(PHCy2)(X)Pt(μ-PCy2)(μ-H) Pt(PHCy2){κP-P(O)Cy2}](Pt-Pt) (X = Cl, Br), treatment of 2 with NaOH in dichloromethane did not restore the parent Complex 1.
Scheme 2. Reactivity of 1 with (trimethylsilyl)acetylene.
Scheme 2. Reactivity of 1 with (trimethylsilyl)acetylene.
Inorganics 02 00591 g007
Complex 2 is a pale yellow solid soluble in alkanes, aromatic solvents and tetrahydrofuran, which gave, from the HR ESI-MS analysis, an intense peak at m/z 1295.5852 (Figure 1) corresponding to the cation obtained by protonation of 2, plus a peak at 2591.1691 ascribable to a protonated dimer (exact mass = 2591.1657 Da) [19] formed in the ionization chamber, presumably featuring a hydrogen ion linking two neutral molecules of 2 through the P=O moieties. The MS/MS spectrogram of the peak at m/z 2591.1691 showed only two peaks at 1197.5257 and 1295.5765 ascribable to [1 + H]+ (exact mass = 1197.5303) and to [2 + H]+ (exact mass = 1295.5855), respectively. On the other hand, the MS/MS spectrogram of the peak at m/z 1295.5852 showed an intense peak at m/z 1197.5210 due to the loss of (trimethylsilyl)acetylene from 2, plus a weak peak at 981.3602, due to the contemporary loss of (trimethylsilyl)acetylene and PH(O)Cy2 from 2.
Figure 1. HR ESI-MS spectrograms of 2 showing the peaks due to [2 + H]+ (A) and [2·2 + H]+ (B). The errors between simulated and observed isotopic patterns are 0.7 ppm for [2 + H]+ and −1.0 ppm for [2·2 + H]+.
Figure 1. HR ESI-MS spectrograms of 2 showing the peaks due to [2 + H]+ (A) and [2·2 + H]+ (B). The errors between simulated and observed isotopic patterns are 0.7 ppm for [2 + H]+ and −1.0 ppm for [2·2 + H]+.
Inorganics 02 00591 g001
The molecular structure of Complex 2 was deduced from IR, NMR and XRD analyses.
In the IR spectrum (KBr disk), Complex 2 showed, beside the bands of P–H and Pt–H–Pt stretchings at 2329 and 1632 cm−1, respectively, a very strong band at 2042 cm−1 attributable to a C≡C stretching of the (trimethylsilyl)acetylide. This value is in the range commonly observed for platinum complexes featuring η1-C≡CSiMe3 (e.g., trans-Pt(C≡CSiMe3)2(PPh3)2) 2041 cm−1 [20]; (NBu4)2[Pt(C≡CSiMe3)4] 2015 cm−1) [21]. In fact, the blue-shift of 4 cm−1 observed for the C≡C stretching on passing from free Me3SiC≡C–H (ν = 2038 cm−1 in n-hexane) to 2 indicates that the C≡C bond order in the coordinated acetylide and in the free (trimethylsilyl)acetylene is nearly the same, thus ruling out a possibleη2-coordination of the C≡C triple bond to Pt.
The position and the multiplicity of the 31P, 1H and 195Pt NMR spectral features are similar to those reported for the analogous compounds having Cl, Br, CF3CH2O or PhO in place of SiMe3–C≡C and support the structure in which the (trimethylsilyl)acetylide is η1-bonded to Pt1 and the P(O)Cy2 ligand is κ-P-bonded to Pt2. For Complex 2, the 31P{1H} NMR spectrum showed four mutually-coupled signals at δ 120.9, δ 77.0, δ 13.7 and δ 2.8 (Figure 2). The signal at δ 120.9 is ascribable to a bridging phosphanide subtending a Pt-Pt bond [22], while the signal at δ 77.0 is diagnostic of a dicyclohexylphosphinite P(O)Cy2 ligand bonded to Pt through the P atom. In fact, Pt-P(O)Cy2 31P NMR resonances fall in the range of 75–90 ppm, while their protonated analogues Pt-P(OH)Cy2 give 31P NMR signals in the range of 115–135 ppm [3,4,23]. The remaining signals (δ 13.7 and δ 2.8) are ascribed to terminal PHCy2 ligands. As a consequence of a higher trans influence of the Me3SiC≡C ligand compared to halogens or alkoxides [24], the value of the Pt1-μP1 direct coupling constant is significantly lower for 2 (1JPt,P = 1,680 Hz) than for analogous complexes having chlorine (1JPt,P = 2,484 Hz), bromine (1JPt,P = 2,465 Hz) or trifluoroethoxide (1JPt,P = 2,520 Hz)3 in place of trimethylsilylacetylide.
Figure 2. 31P{1H} NMR spectrum of 2 (C6D6, 298 K).
Figure 2. 31P{1H} NMR spectrum of 2 (C6D6, 298 K).
Inorganics 02 00591 g002
In the 1H NMR spectrum of 2, the presence of a bridging hydride is attested by a multiplet centered at δ −4.46 flanked by two sets of 195Pt satellites (Figure 3a). The multiplicity of the central signal is due only to 1H-31P couplings, as indicated by the 1H{31P} spectrum shown in Figure 3b.
The 195Pt resonances (δ −5,582 and δ −5,562) were attributed to Pt1 and Pt2, respectively, by combined analysis of the 195Pt{1H} NMR,31P{1H} NMR and 1H-195Pt HMQC spectra.
Figure 3. Portion of the 1H (a) and 1H{31P} (b) spectra of 2 showing the hydride signal (C6D6, 298 K).
Figure 3. Portion of the 1H (a) and 1H{31P} (b) spectra of 2 showing the hydride signal (C6D6, 298 K).
Inorganics 02 00591 g003
The 13C{1H} NMR spectrum in n-hexane showed, beside the cyclohexyl signals, a sharp singlet at δ 1.5 attributable to SiMe3 and two low-field signals flanked by 195Pt satellites at δ 129.9 (doublet of doublets) and at δ 116.6 (doublet) attributable to the quaternary carbons C1 and C2 (Scheme 2). The chemical shift and the multiplicity of these signals confirm the σ-coordination of the (trimethylsilyl)acetylide to Pt. In fact, C1 showed coupling constants JC,P of 82 Hz (coupling with the bridging phosphanide) and 19 Hz (coupling with one of the PHCy2), while C2 showed only a coupling constant JC,P of 16 Hz, due to the coupling with the bridging phosphanide. Accordingly, the 13C-195Pt coupling constants are JC,Pt = 1011 Hz for C1 and JC,Pt = 253 Hz for C2, which are comparable to those obtained for platinum complexes featuring terminal bonded alkynyl ligands (e.g., data for (NBu4)2[PtC≡CSiMe3)4] δCα 140.4, 1JPt-Cα = 924 Hz; δCβ 104.4, 2JPt-Cβ = 250.4 Hz [25].

2.2. X-ray Structure of 2

Crystals suitable for XRD analysis were obtained from a saturated solution of 2 in n-hexane. Not unexpectedly, due to the presence of phosphinito ligands, Complex 2 incorporates a water molecule and crystallizes as 2·2 n-hexane·H2O. The molecular structure is shown in Figure 4. Selected bond parameters (angles and lengths) and crystallographic data are summarized in Table 1 and Table 2. As is shown in Figure 4b, each water molecule is simultaneously hydrogen bonded to two phosphinite ligands of two neighboring diplatinum complexes, giving rise to a dimer generated at equivalent positions (−x, −y, −z). A similar feature has been observed in the mononuclear phosphinito complex trans-[Pt(Cl)(PHCy2)2 {κP-P(O)Cy2}] [25] and in the dinuclear species [(PHCy2){κP-P(O)Cy2} Pt(μ-PCy2)(μ-H)Pt(PHCy2){κP-P(O)Cy2}](Pt-Pt).2 The O1···O2, O1···O2’ distances of 2.726(5) and 2.775(6) Å found for 2·2 n-hexane·H2O lie in the range of other hydrated systems [26,27]. The central dimetallic core Pt(μ-PCy2)(μ-H)Pt of 2·2 n-hexane·H2O is rather similar to that of [(PHCy2) {κP-P(O)Cy2}Pt(μ-PCy2)(μ-H)Pt(PHCy2){κP-P(O)Cy2}](Pt-Pt)2 with Pt–H distances (Pt(1)-H 1.80(5) Å; Pt(2)-H 1.65(5) Å) essentially identical within experimental error. These values and the Pt···Pt distance (2.8209(2) Å) compare to those found in other 30 e diplatinum hydride complexes containing mixed (μ-X)(μ-H) bridges [22,28,29,30,31] for which the existence of a through-ring Pt-Pt bonding interaction has been suggested on the basis of theoretical calculations [32]. The Pt-C(1) (2.036(4) Å) and Pt-P distances (2.2409(10)–2.3255(10) Å) and structural details of the terminal C≡CSiMe3 (C≡C 1.192(6) Å; Pt(1)-C(1)-C(2) 177.1(4)º; C(1)-C(2)-Si(1) 177.2(4)º) are unexceptional for these ligands. Although many different types of dimeric hydride bridged platinum complexes have been reported, including systems containing mixed (μ-C≡CR)(μ-H), complexes containing terminal alkynyl as auxiliary ligands are very rare. To the best of our knowledge, only the related complexes [(η1-C≡CR)(But2HP) Pt(μ-PBut2)(μ-H)Pt(CO)(PBut2H)](TfO) (R = But, Ph) and [Pt2(C≡Ph)2(µ-H)(µ-dppm)2]C12 characterized by spectroscopic means have been reported [33,34].
Table 1. Selected bond lengths and angles for 2·2 n-hexane·H2O. The prime (') character in the atoms labels indicates that these atoms are at equivalent positions (−x, −y, −z).
Table 1. Selected bond lengths and angles for 2·2 n-hexane·H2O. The prime (') character in the atoms labels indicates that these atoms are at equivalent positions (−x, −y, −z).
Distances (Å)
Pt(1)-C(1)2.036(4)Pt(2)-P(3)2.3255(10)
Pt(1)-P(1)2.2409(10)Pt(2)-P(4)2.2690(10)
Pt(1)-P(2)2.2864(10)Pt(2)-P(2)2.3089(9)
Pt(1)-H(100)1.80(5)Pt(2)-H(100)1.65(5)
Pt(1)-P(2)2.8209(2)C(1)-C(2)1.192(6)
O(1)-O(2)2.726(5)O(1)-O(2)’2.775(6)
Angles (°)
C(1)-Pt(1)-P(1)90.76(11)P(3)-Pt(2)-P(4)93.84(4)
P(1)-Pt(1)-P(2)104.46(4)P(4)-Pt(2)-P(2)102.67(4)
P(2)-Pt(1)-H(100)84.1(15)P(2)-Pt(2)-H(100)86.7(16)
C(1)-Pt(1)-H(100)80.7(15)P(3)-Pt(2)-H(100)77.1(16)
H(100)-Pt(1)-Pt(2)33.5(15)H(100)-Pt(2)-Pt(1)37.0(16)
P(2)-Pt(1)-Pt(2)52.49(2)P(2)-Pt(2)-Pt(1)51.77(2)
Pt(1)-P(2)-Pt(2)75.74(3)Pt(1)-H(100)-Pt(2)109.5(28)
Pt(1)-C(1)-C(2)177.1(4)C(1)-C(2)-Si(1)177.2(4)
Figure 4. View of the molecular structure of 2·2 n-hexane·H2O; (a) the asymmetric unit showing the presence of the oxygen from the H2O and (b) the dimer generated through four O···H···O bonds and at equivalent positions (–x, –y, –z).
Figure 4. View of the molecular structure of 2·2 n-hexane·H2O; (a) the asymmetric unit showing the presence of the oxygen from the H2O and (b) the dimer generated through four O···H···O bonds and at equivalent positions (–x, –y, –z).
Inorganics 02 00591 g004

2.3. Photoluminescence of 2

Due to the presence of the very short Pt···Pt distance, we decided to investigate the photoluminescent properties of Complex 2 in the solid state. The complex is not emissive at room temperature. However, it displays a long-lived, weak emission at 77 K (Figure 5), which fits to a double exponential decay (τ1 5.6 µs 55%; τ2 13.5, µs 45%). According to previous luminescence studies in alkynyl dinuclear complexes with short Pt··Pt distances [35,36], this emission is tentatively ascribed to an excited state having a metal-metal to alkynyl charge transfer character (3MMLCT).
The profile of the emission band is asymmetric, and the excitation spectra monitored at two different wavelengths are somewhat different, suggesting the presence of heterogeneity in the bulk solid.
Figure 5. Excitation and emission spectra of 2 in solid state at 77 K. (first line).
Figure 5. Excitation and emission spectra of 2 in solid state at 77 K. (first line).
Inorganics 02 00591 g005

2.4. Reactivity of 2 with HBF4

Complex 2 possesses several sites susceptible to protonation. In fact, beside the bridging phosphanide, the bridging hydride, the terminal (trimethylsilyl)acetylide (that could give dihydrogen and trimethylsilylacetylene, respectively) and the dicyclohexylphosphinite represent possible protonation sites.
In order to study the site selectivity of the protonation reaction, Complex 2 was treated with HBF4·Me2O in n-hexane (Scheme 3). The reaction gave smoothly the corresponding dicyclohexylphosphinic acid complex [3]BF4 as the only product, indicating that only the POCy2 ligand is protonated. The most striking evidence revealing the transformation of the POCy2 ligand into P(OH)Cy2 is the presence, in the 1H NMR spectrum of [3]BF4, of a broad peak at δ 6.71 ascribable to the POH proton. Moreover, as reported in related hydrido bridged diplatinum phosphinito species, [3,4] the protonation of the POCy2 ligand provokes a high field shift of the 1H NMR hydride resonance (from δ −4.46 (in 2) to δ −5.83 (in [3]BF4)) and a low field shift of the phosphinito 31P NMR resonance (from δ 77.0 (in 2) to δ 124.3 (in [3]BF4)). IR and 13C NMR data confirmed the presence of the η1-acetylide bonded to Pt1 in [3]BF4. In fact, the C≡C stretching band was found in the IR spectrum at 2045 cm−1, while the 13C{1H} NMR signals of the coordinated (trimethylsilyl)acetylide were found at δ 126.7 (C1), δ 118.1 (C2) and 1.2 (CH3).
Scheme 3. Protonation of 2 by tetrafluoroboric acid.
Scheme 3. Protonation of 2 by tetrafluoroboric acid.
Inorganics 02 00591 g008

3. Experimental Section

Complex 1 was prepared as described [2]. (Trimethylsilyl)acetylene and etherate fluoroboric acid were from commercial suppliers and were used without further purification. All manipulations were carried out under a pure nitrogen atmosphere, using freshly distilled and oxygen-free solvents. C, H elemental analyses were carried out on a Vario Micro-CHNSO elemental analyzer. Infrared spectra were recorded on a Jasco 4200-FTIR spectrometer. NMR spectra were recorded on a Bruker Avance 400 spectrometer; frequencies are referenced to Me4Si (1H and 13C), 85% H3PO4 (31P) and H2PtCl6 (195Pt). The signal attributions (see Scheme 4 and Scheme 5 for atom numbering) and coupling constant assessment was made on the basis of a multinuclear NMR analysis, including 1H-31P HMQC, 1H-195Pt HMQC, COSY and NOESY experiments. The coupling constants not directly extractable from the monodimensional spectra were obtained and attributed by the tilts of the multiplets due to the “passive” nuclei [37] in the aforementioned 2D spectra. High resolution mass spectrometry (HR-MS) analyses were performed using a time-of-flight mass spectrometer equipped with an electrospray ion source (Bruker micrOTOF-Q II). The sample solutions were introduced by continuous infusion with the aid of a syringe pump at a flow-rate of 180 μL/h. The instrument was operated at end plate offset −500 V and capillary −4500 V. Nebulizer pressure was 0.3 bar (N2) and the drying gas (N2) flow 4.0 L/min. Capillary exit was 170 V. The drying gas temperature was set at 180 °C. The software used for the simulations is Bruker Daltonics Data Analysis (version 4.0, Bruker Daltonik GmbH, Bremen, Germany). All experimental HRMS signals had an isotope pattern superimposable to that calculated for the relevant formula, on the basis of the natural abundances. Emission and excitation spectra were obtained on a Jobin-Yvon Horia Fluorog 3-11 Tau-3 spectrofluorometer (Horiba, Kyoto, Japan), with the lifetime measured in the phosphorimeter mode.

3.1. Synthesis of 2

To a n-hexane solution of 1 (164 mg, 0.137 mmol in 4.0 mL), 21 μL of (trimethylsilyl)acetylene (14.8 mg, 0.150 mmol) were added at 298 K and stirred for 12 h. After the reaction, the mixture was taken to dryness under vacuum to eliminate the solvent and the excess (trimethylsilyl)acetylene. The residue was redissolved in 1.0 mL of n-hexane, and the procedure was repeated twice. The final pale yellow solid was dried under high vacuum overnight.
Yield: 163 mg (92%).
The complex is air stable in the solid state and is very soluble in halogenated solvents, aromatic solvents, diethyl ether and n-hexane.
C53H100OP4Pt2Si: calcd. C, 49.14; H, 7.78; found C 49.09, H 7.59.
HRMS(+) (n-hexane): exact mass calcd. for C53H100OP4Pt2Si: 1294.5786 da; found: 1295.5852 da [M + H]+.
IR (KBr): = 2329 (m, P-H), 2042 (vs, C≡C), 1632 (w, Pt-H-Pt), 1446 (vs), 1342 (m), 1293 (m), 1266 (m), 1241 (s), 1191 (s), 1175 (s), 1117 (s), 1083 (s, P=O), 1044 (m), 1003 (vs, P=O), 916 (s), 888 (s), 858 (vs), 819 (s), 754 (m), 728 (s), 689 (m), 540 (s), 518 (s), 462 (s), 414 (m) cm−1.
1H NMR (400 MHz, C6D6, 298 K): δ = 5.36 [m, H(4), 1JH(4),P(4) = 344 Hz, 2JH(4),Pt(2) = 46 Hz], 4.70 [m, H(2), 1JH(2),P(2) = 376 Hz, 2JH(2),Pt(1) = 83 Hz], 0.44 [s, H(3)], −4.46 [m, H(1), 2JH(1),P(4) = 76 Hz, 2JH(1),P(2) = 68 Hz, 2JH(1),P(1) = 13 Hz, 2JH(1),P(3) = 7 Hz] ppm.
13C{1H} NMR (100 MHz, n-hexane, 298 K, δ): 129.9 (dd, C1, 2JC,P(1) = 82 Hz, 2JC,P = 19 Hz, 1JC,Pt(1) = 1011 Hz), 116.6 (d, C2, 3JC,P(1) = 16 Hz, 2JC,Pt(1) = 253 Hz), 1.5 (s, C3) ppm.
31P{1H} NMR (161 MHz, C6D6, 298 K): δ = 120.8 [d, P(1), 2JP(1),P(3) = 245 Hz, 1JP(1),Pt(1) = 1680 Hz, 1JP(1),Pt(2) = 1170 Hz], 77.0 [d, P(3), 2JP(3),P(1) = 245 Hz, 1JP(3),Pt(2) = 2333 Hz], 13.7 [broad, P(4), 3JP(4),P(2) = 55 Hz, 1JP(4),Pt(2) = 3946 Hz], 2.8 [d, P(2), 3JP(2),P(4) = 55 Hz, 1JP(2),Pt(1) = 3726 Hz, 2JP(2),Pt(2) = 209 Hz] ppm.
195Pt{1H} NMR (86 MHz, C6D6, 298 K): δ = −5582 [ddd, Pt(1), 1JPt(1),P(1) = 1680 Hz, 1JPt(1),P(2) = 3726 Hz, 2JPt(1),P(4) = 224 Hz], −5562 [m, Pt(2), 1JPt(2),P(1) = 1170 Hz, 2JPt(2),P(2) = 209 Hz, 1JPt(2),P(3) = 2333 Hz, 1JPt(2),P(4) = 3946 Hz] ppm.
Scheme 4. Atom numbering for Complex 2.
Scheme 4. Atom numbering for Complex 2.
Inorganics 02 00591 g009

3.2. Synthesis of [3]BF4

To an n-hexane solution of 2 (104 mg, 0.080 mmol in 2.0 mL), 10 μL of HBF4·Me2O (11 mg, 0.080 mmol) were added at 298 K, causing the precipitation of a brownish solid. The solid was filtered off, washed with n-hexane (3 × 1 mL) and dried under high vacuum.
Yield: 77 mg (70%).
Complex [3]BF4 is stable in the solid state, but slowly undergoes a substitution of a chloride for the trimethylsilylacetylide, when it is dissolved in chlorinated solvents. It is soluble in halogenated solvents and insoluble in diethyl ether, in alkanes and in aromatic solvents.
C53H101BF4OP4Pt2Si: calcd. C, 46.02; H, 7.36; found C 46.09, H 7.49.
HRMS(+) (i-propanol): exact mass calcd. for the cation C53H101OP4Pt2Si: 1295.5855 da; found: 1295.5808 da [M]+.
IR (Nujol mull): = 3358 (broad m, P–O–H), 2662 (w), 2333 (m, P-H), 2045 (vs, C≡C), 1628 (w, Pt-H-Pt), 1452 (vs), 1378 (s), 1295 (m), 1270 (m), 1058 (broad s, BF4 +PO) 854 (vs), 733 (vs), 518 (s), 467 (s), cm−1.
1H NMR (400 MHz, CD2Cl2, 298 K): δ = 6.71 [broad, H(5)], 5.07 [m, H(4), 1JH(4),P(4) = 372 Hz], 4.93 [m, H(2), 1JH(2),P(2) = 353 Hz, 2JH(2),Pt(1) = 72 Hz], 0.12 [s, H(3)], −5.83 [m, H(1), 2JH(1),P(4) = 72 Hz, 2JH(1),P(2) = 63 Hz, 2JH(1),P(1) = 18 Hz, 2JH(1),P(3) = 12 Hz] ppm.
13C{1H} NMR (100 MHz, CD2Cl2, 298 K, δ): 126.7 (dd, C1, 2JC,P(1) = 86 Hz, 2JC,P = 15 Hz), 118.1 (d, C2, 3JC,P(1) = 20 Hz, 2JC,Pt(1) = 243 Hz), 1.2 (s, C3) ppm.
31P{1H} NMR (161 MHz, CD2Cl2, 298 K): δ = 144.4 [d, P(1), 2JP(1),P(3) = 262 Hz, 1JP(1),Pt(1) = 1742 Hz, 1JP(1),Pt(2) = 1471 Hz], 124.3 [d, P(3), 2JP(3),P(1) = 262 Hz, 1JP(3),Pt(2) = 2710 Hz], 2.5 [broad, P(4), 1JP(4),Pt(2) = 3588 Hz], 0.6 [d, P(2), 3JP(2),P(4) = 49 Hz, 1JP(2),Pt(1) = 3786 Hz, 2JP(2),Pt(2) = 218 Hz] ppm.
195Pt{1H} NMR (86 MHz, CD2Cl2, 298 K): δ = −5622 [dd, Pt(1), 1JPt(1),P(1) = 1742 Hz, 1JPt(1),P(2) = 3786 Hz], −5673 [m, Pt(2), 1JPt(2),P(1) = 1471 Hz, 2JPt(2),P(2) = 262 Hz, 1JPt(2),P(3) = 2710 Hz, 1JPt(2),P(4) = 3588 Hz] ppm.
Scheme 5. Atom numbering for Complex 3+.
Scheme 5. Atom numbering for Complex 3+.
Inorganics 02 00591 g009

3.3. X-ray Structure Determination

Details of the structural analysis for Complex 2 are summarized in Table 2. Pale yellow crystals were obtained by slow evaporation at 4 °C of a concentrated solution of the complex in n-hexane. Two molecules of n-hexane and one of water were found in the asymmetric unit. Graphite-monochromatic Mo-Kα radiation was used. X-ray intensity data were collected with a NONIUS-κCCD area-detector diffractometer (CAMCOR, Eugene, OR, USA) and images processed using the DENZO (Academic Press: New York, NY, USA) and SCALEPACK suite of programs [38], carrying out the absorption correction at this point. The structure was solved by direct and Patterson methods using SIR2004 [39] and refined by full-matrix least squares on F2 with SHELXL-97 [40]. All non-hydrogen atoms were assigned anisotropic displacement parameters. The hydride ligand, H(100), has been located from difference maps and assigned isotropic parameters. The rest of the hydrogen atoms (except the H2O hydrogens) were constrained to idealized geometries fixing isotropic displacement parameters 1.2-times the Uiso value of their attached carbon. Several restraints have been used in order to model the n-hexane molecules. The hydrogen atoms of the H2O molecule have been omitted from the density map, but have been included in the empirical formula and in the molecular weight. Finally, the structure showed one residual peaks greater than 1 eA−3 in the vicinity of the Pt(2) atom, but with no chemical meaning.

4. Conclusions

Reaction of the phosphinito-phosphanido Pt(I) complex [(PHCy2)Pt(μ-PCy2){κ2P,O-μ-P(O)Cy2} Pt(PHCy2)](Pt-Pt) (1) with (trimethylsilyl)acetylene results in the C-H bond activation of the alkyne with the formation of the bridging hydride complex [(PHCy2)(η1-Me3SiC≡C)Pt(μ-PCy2){κP-P(O)Cy2} Pt(PHCy2)](Pt–Pt) (2), which displays a σ-bonded acetylide ligand and a bridging hydride. The reactivity of 1 with (trimethylsilyl)acetylene parallels that already found with PhOH or CF3CH2OH, two weak Brønsted acids endowed with a coordinating anion. Complex 2 is stable both in the solid state and in solution, and it is resistant to deprotonation by NaOH. It crystallizes from n-hexane as 2·2 n-hexane·H2O. The water molecule is simultaneously hydrogen bonded to two phosphinite ligands of two neighboring diplatinum complexes, giving rise to a dimer. Due to the presence of the short Pt-Pt bond, Complex 2 exhibits a long-lived weak emission at 77 K, which is ascribed to an excited state having a metal-metal to alkynyl charge transfer character (3MMLCT). Despite the number of protonation sites present in Complex 2, its reaction with HBF4·Me2O in toluene gave the phosphinic acid cationic Complex 3+ selectively, still embodying the trimethylsilylacetylide ligand.
Table 2. Selected crystal data and structure refinement details for 2·2C6H14·H2O.
Table 2. Selected crystal data and structure refinement details for 2·2C6H14·H2O.
Parameters2·2C6H14·H2O
empirical formulaC65H130O2P4Pt2Si
Fw1485.84
T (K)173(1)
crystal system; space grouptriclinic; P–1
a (Å); α (°)13.1300(3); 67.8780(10)
b (Å); β(°)16.4810(6); 88.243(2).
c (Å); γ (°)19.1700(7); 73.174(2).
V3); Z3664.0(2); 2
Dcalcd (Mg/m3)1.345
absorption coefficient (mm−1)3.954
F(000)1528
crystal size (mm)0.6 × 0.2 × 0.05
θ range for data collection (°)3.19 to 27.48
index ranges0 ≤ h ≤ 16,
−20 ≤ k ≤ 21
−24 ≤ l ≤ 24
No of data/restraints/params16,496/12/611
goodness of fit on F2 a1.014
final R indexes (I > 2σ(I)) aR1 = 0.0329, wR2 = 0.0777
R indices (all data) aR1 = 0.0441, wR2 = 0.082
largest diff. peak and hole (e·Å−3)1.580 and −1.078
a R1 = Σ(|Fo| − |Fc|)/Σ|Fo|; wR2 = [Σw(Fo2Fc2)2wFo2] 1/2; goodness of fit = {Σ[w(Fo2Fc2)2]/(NobsNparam)}1/2; w = [σ2(Fo) + (g1P)2 + g2P]−1; P = [max(Fo2;0 + 2Fc2]/3.

Acknowledgments

We wish to thank Politecnico of Bari for FRA funding and the Spanish MINECO projects (CTQ 2008-06669-C02-02 and CTQ2013-45518-P) for financial support.

Author Contributions

Piero Mastrorilli and Elena Lalinde: Responsibles for research publication. Mario Latronico: Inorganic synthesis. Stefano Todisco: MS and IR analysis. Vito Gallo: NMR. Santiago Ruiz: Crystallographer.

Conflicts of Interest

The authors declare no conflict of interest.

References and Notes

  1. Gallo, V.; Latronico, M.; Mastrorilli, P.; Nobile, C.F.; Suranna, G.P.; Ciccarella, G.; Englert, U. Synthesis of phosphido-bridged phosphinito platinum(I) complexes by reaction of cis-PtCl2(PHCy2)2 with oxygenated bases—Crystal structure of [(PCy2OMe)Pt(μ-PCy2)]2(Pt-Pt). Eur. J. Inorg. Chem. 2005, 22, 4607–4616. [Google Scholar]
  2. Gallo, V.; Mastrorilli, P.; Polini, F.; Latronico, M.; Re, N.; Englert, U. Reactivity of a phosphinito-bridged PtI-PtI complex with nucleophiles: Substitution versus addition. Inorg. Chem. 2008, 47, 4785–4795. [Google Scholar]
  3. Latronico, M.; Polini, F.; Gallo, V.; Mastrorilli, P.; Calmuschi-Cula, B.; Englert, U.; Re, N.; Repo, T.; Raisanen, M. Site selectivity in the protonation of a phosphinito bridged PtI-PtI complex: A combined NMR and density-functional theory mechanistic study. Inorg. Chem. 2008, 47, 9779–9796. [Google Scholar]
  4. Latronico, M.; Mastrorilli, P.; Gallo, V.; Dell’Anna, M.M.; Creati, F.; Re, N.; Englert, U. Hydrido phosphanido bridged polynuclear complexes obtained by protonation of a phosphinito bridged Pt(I) complex with HBF4 and HF. Inorg. Chem. 2011, 50, 3539–3858. [Google Scholar]
  5. Forniés, J.; Lalinde, E. Platinum–Carbon π-Bonded Complexes. In Comprehensive Organometallic Chemistry III; Carbtree, R.H., Mingos, D.M.P., Eds.; Elsevier: Oxford, UK, 2007; pp. 611–673. [Google Scholar]
  6. Belluco, U.; Bertani, R.; Michelin, R.A.; Mozzon, M. Platinum-alkynyl and -alkyne complexes: Old systems with new chemical and physical perspectives. J. Organomet. Chem. 2000, 600, 37–55. [Google Scholar]
  7. Hahn, C.; Miranda, M.; Chittineni, N.P.B.; Pinion, T.A.; Perez, R. Mechanistic Studies on Platinum(II) Catalyzed Hydroarylation of Alkynes. Organometallics 2014, 33, 3040–3050. [Google Scholar]
  8. Jourdain, I.; Knorr, M.; Strohmann, C.; Unkelbach, C.; Rojo, S.; Gómez-Iglesias, P.; Villafañe, F. Reactivity of Silyl-Substituted Iron–Platinum Hydride Complexes toward Unsaturated Molecules: 4. Insertion of Fluorinated Aromatic Alkynes into the Platinum-Hydride Bond. Synthesis and Reactivity of Heterobimetallic Dimetallacylopentenone, Dimetallacyclobutene, μ-Vinylidene, and μ2-σ-Alkenyl Complexes. Organometallics 2013, 32, 5343–5359. [Google Scholar]
  9. Berenguer, J.R.; Fernández, J.; Giménez, N.; Lalinde, E.; Moreno, M.T.; Sánchez, S. Unexpected Formation of Ferrocenyl(vinyl)benzoquinoline Ligands by Oxidation of an Alkyne Benzoquinolate Platinum(II) Complex. Organometallics 2013, 32, 3943–3953. [Google Scholar]
  10. West, N.M.; White, P.S.; Templeton, J.L. Alkyne Insertion into the Pt-H Bond of Pt(H)(1-pentene)(β-diiminate) Initiates a Reaction Cascade That Results in C-H Activation or C-C Coupling. Organometallics 2008, 27, 5252–5262. [Google Scholar]
  11. Wang, W.; Yang, H.B. Linear neutral platinum-acetylide moiety: Beyond the links. Chem. Commun. 2014, 50, 5171–5186. [Google Scholar]
  12. Muro, M.L.; Rachford, A.A.; Wang, X.; Castellano, F.N. PlatinumII Acetylide Photophysics. Top. Organomet. Chem. 2010, 29, 159–191. [Google Scholar]
  13. Berenguer, J.R.; Lalinde, E.; Moreno, M.T. An overview of the chemistry of homo and heteropolynuclear platinum complexes containing bridging acetylide (μ-C≡CR) ligands. Coord. Chem. Rev. 2010, 254, 832–875. [Google Scholar]
  14. Canty, A.J.; Sharma, M. η1-Alkynyl Chemistry for the Higher Oxidation State of Palladium and Platinum. Top. Organomet. Chem. 2011, 35, 111–128. [Google Scholar]
  15. Long, N.J.; Willians, C.K. Metal Alkynyl σ Complexes: Synthesis and Materials. Angew. Chem. Int. Ed. 2003, 42, 2586–2617. [Google Scholar]
  16. Tao, C.H.; Yam, V.W.W. Branched carbon-rich luminescent multinuclear platinum(II) and palladium(II) alkynyl complexes with phosphine ligands. J. Photochem. Photobio. C 2009, 10, 130–140. [Google Scholar]
  17. Wong, K.M.; Yam, V.W.W. Self-Assembly of Luminescent Alkynylplatinum(II) Terpyridyl Complexes: Modulation of Photophysical Properties through Aggregation Behavior. Acc. Chem. Res. 2011, 44, 424–434. [Google Scholar]
  18. Wong, W.Y. Luminescent organometallic poly(aryleneethynylene)s: Functional properties towards implications in molecular optoelectronics. Dalton Trans. 2007, 4495–4510. [Google Scholar]
  19. Throughout the paper, the calculated (exact mass) and the experimental (accurate) m/z values have been compared considering the principal ion (which gives the most intense peak) of the isotope pattern.
  20. Berenguer, J.R.; Forniés, J.; Martínez, F.; Cubero, J.C.; Lalinde, E.; Moreno, M.T.; Welch, A.J. Synthesis and reactivity of bimetallic acetylide-bridged Pt-Pt complexes. Crystal and molecular structure of [(PPh3)(C6F5)Pt(μ-C≡CPh)2Pt(C6F5)(PPh3)]. Polyhedron 1993, 12, 1797–1804. [Google Scholar]
  21. Benito, J.; Berenguer, J.R.; Forniés, J.; Gil, B.; Gómez, J.; Lalinde, E. Synthesis, characterization and luminescence properties of homoleptic platinum(II) acetylide complexes. Dalton Trans. 2003. [Google Scholar] [CrossRef]
  22. Mastrorilli, P. Bridging and Terminal (Phosphanido)platinum Complexes. Eur. J. Inorg. Chem. 2008, 31, 4835–4850. [Google Scholar]
  23. Mastrorilli, P.; Latronico, M.; Gallo, V.; Polini, F.; Re, N.; Marrone, A.; Gobetto, R.; Ellena, S. Facile Activation of Dihydrogen by a Phosphinito-Bridged Pt(I)-Pt(I) Complex. J. Am. Chem. Soc. 2010, 132, 4752–4765. [Google Scholar]
  24. Furlani, A.; Licoccia, S.; Russo, M.V.; Chiesi, V.A.; Guastini, C. New platinum hydrido acetylides. Crystal and molecular structure of [PtH{C≡C–C(OH)MeEt}(PPh3)2]. Dalton Trans. 1982. [Google Scholar] [CrossRef]
  25. Mastrorilli, P.; Nobile, C.F.; Latronico, M.; Gallo, V.; Englert, U.; Fanizzi, F.P.; Sciacovelli, O. Mltinuclear and dynamic NMR study of trans-[Pt(Cl)(PHCy2)2(PCy2)], [Pt(Cl)(PHCy2)3][BF4], [Pt(Cl)(PHCy2)3][Cl], trans-[Pt(Cl) (PHCy2)2{P(S)Cy2}], and trans-[Pt(Cl)(PHCy2)2{P(O)Cy2}]. Inorg. Chem. 2005, 44, 9097–9104. [Google Scholar]
  26. Desiraju, G.R.; Steiner, T. The Weak Hydrogen Bond in Structural Chemistry and Biology; Oxford University Press: Oxford, UK, 1999. [Google Scholar]
  27. Nishio, M. Weak Hydrogen Bonds in Encyclopedia of Supramolecular Chemistry; Atwood, J.L., Steed, J.L., Eds.; Marcel Dekker Inc.: New York, NY, USA, 2004. [Google Scholar]
  28. Ara, I.; Falvello, L.R.; Forniés, J.; Lalinde, E.; Martínez, F.; Moreno, M.T. Synthesis and Structure of New Neutral Bimetallic Platinum Hydride Complexes. Organometallics 1997, 16, 5392–5405. [Google Scholar]
  29. Berenguer, J.R.; Bernechea, M.; Lalinde, E. Rearrangement or C-H Activation Processes Promoted by Reaction with the Solvate [cis-Pt(C6F5)2(thf)2]. Organometallics 2007, 26, 1161–1172. [Google Scholar]
  30. Leoni, P.; Manetti, S.; Pasquali, M. Synthesis of New Neutral and Cationic Phosphido-Bridged Dinuclear Platinum(II) Hydride. Inorg. Chem. 1995, 34, 749–752. [Google Scholar]
  31. Berenguer, J.R.; Bernechea, M.; Lalinde, E. C-H and P-C(Ph) activation competitive processes caused by interaction with the solvate [cis-Pt(C6F5)2(thf)2]. Dalton Trans. 2007, 2384–2393. [Google Scholar]
  32. Aullón, G.; Alemany, P.; Alvarez, S. Through-ring bonding in edge-sharing dimers of square planar complexes. J. Organomet. Chem. 1994, 478, 75–83. [Google Scholar]
  33. Crementieri, S.; Leoni, P.; Marchetti, F.; Marchetti, L.; Pasquali, M. Stable η1-Alkynyl−μ,η12-Alkenyl Complexes from the Reaction of Terminal Alkynes with Encumbered Dinuclear Platinum Compounds and Their Formyl, Methoxycarbonyl, and Hydride Derivatives. Organometallics 2002, 21, 2575–2577. [Google Scholar]
  34. Langrick, C.R.; Pringle, P.G.; Shaw, B.L. Bimetallic systems. Part 10. Synthesis of complexes of type [(RC≡C)Pt(µ-dppm)2Pt(C≡CR)] (dppm = Ph2PCH2PPh2, R = Ph or p-tolyl) and their corresponding “A frames” [(RC≡C)Pt(µ-dppm)2(µ-H)Pt(C≡CR)]Cl or [(RC≡C)Pt(µ-dppm)2(µ-X)Pt(C≡CR)] with X = CS2 or MeOOCC≡CCOOMe. Dalton Trans. 1985. [Google Scholar] [CrossRef]
  35. Lam, W.H.; Yam, V.W.W. Computational Studies on the Photophysical Properties and NMR Fluxionality of Dinuclear Platinum(II) A-Frame Alkynyl Diphosphine Complexes. Inorg. Chem. 2010, 49, 10930–10939. [Google Scholar]
  36. Wong, K.M.C.; Hui, K.C.; Yu, L.K.; Yam, V.W.W. Luminescence studies of dinuclear platinum(II) alkynyl complexes and their mixed-metal platinum(II) -copper(I) and -silver(I) complexes. Coord. Chem. Rev. 2002, 229, 123–132. [Google Scholar]
  37. Carlton, L. Relative Sign of J Couplings from Two-Dimensional NMR Spectra. Bruker Rep. 2000, 148, 28–29. [Google Scholar]
  38. Otwinowski, Z.; Minor, W. Macromolecular Crystallography, Part A; Carter, C.W., Jr., Sweet, R.M., Eds.; Academic Press: New York, NY, USA, 1997; Volume 276, pp. 307–326. [Google Scholar]
  39. Burla, M.C.; Caliandro, R.; Camalli, M.; Carrozzini, B.; Cascarano, G.L.; De Caro, L.; Giacovazzo, C.; Polidori, G.; Spagna, R. SIR2004: An improved tool for crystal structure determination and refinement. J. Appl. Crystallogr. 2005, 38, 381–388. [Google Scholar]
  40. Sheldrick, G.M. SHELXL-97; University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]

Share and Cite

MDPI and ACS Style

Latronico, M.; Gallo, V.; Lalinde, E.; Ruiz, S.; Todisco, S.; Mastrorilli, P. A Hydrido η1-Alkynyl Diplatinum Complex Obtained from a Phosphinito Phosphanido Complex and Trimethylsilylacetylene. Inorganics 2014, 2, 591-605. https://doi.org/10.3390/inorganics2040591

AMA Style

Latronico M, Gallo V, Lalinde E, Ruiz S, Todisco S, Mastrorilli P. A Hydrido η1-Alkynyl Diplatinum Complex Obtained from a Phosphinito Phosphanido Complex and Trimethylsilylacetylene. Inorganics. 2014; 2(4):591-605. https://doi.org/10.3390/inorganics2040591

Chicago/Turabian Style

Latronico, Mario, Vito Gallo, Elena Lalinde, Santiago Ruiz, Stefano Todisco, and Piero Mastrorilli. 2014. "A Hydrido η1-Alkynyl Diplatinum Complex Obtained from a Phosphinito Phosphanido Complex and Trimethylsilylacetylene" Inorganics 2, no. 4: 591-605. https://doi.org/10.3390/inorganics2040591

Article Metrics

Back to TopTop