Next Article in Journal
Synthesis and Characterization of NaCd0.92Sn1.08, Na(Cd0.28Sn0.72)2 and Na2CdSn5 with Three-Dimensional Cd-Sn Frameworks
Previous Article in Journal
Misinterpretations in Evaluating Interactions of Vanadium Complexes with Proteins and Other Biological Targets
Previous Article in Special Issue
Preparation and DFT Studies of κ2C,N-Hypercoordinated Oxazoline Organotins: Monomer Constructs for Stable Polystannanes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

CO2 Derivatives of Molecular Tin Compounds. Part 2: Carbamato, Formato, Phosphinoformato and Metallocarboxylato Complexes †

by
Laurent Plasseraud
Institut de Chimie Moléculaire de l’Université de Bourgogne (ICMUB), UMR-CNRS 6302, Université de Bourgogne Franche-Comté, 9 Avenue A. Savary, F-21078 Dijon, France
For Part 1, Please Refer to: CO2 Derivatives of Molecular Tin Compounds. Part 1: Hemicarbonato and Carbonato Complexes. Inorganics 2020, 8, 31; https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8050031
Submission received: 16 December 2020 / Revised: 4 February 2021 / Accepted: 8 February 2021 / Published: 24 February 2021
(This article belongs to the Special Issue Hypercoordinated Organotin Compounds)

Abstract

:
Single-crystal X-ray diffraction structures of organotin compounds bearing hemicarbonate and carbonate ligands were recently reviewed by us—“CO2 Derivatives of Molecular Tin Compounds. Part 1: Hemicarbonato and Carbonato Complexes”, Inorganics 2020, 8, 31—based on crystallographic data available from the Cambridge Structural Database. Interestingly, this first collection revealed that most of the compounds listed were isolated in the context of studies devoted to the reactivity of tin precursors towards carbon dioxide, at atmospheric pressure or under pressure, thus highlighting the suitable disposition of Sn to fix CO2. In the frame of a second part, the present review carries on to explore CO2 derivatives of molecular tin compounds by describing successively the complexes with carbamato, formato, and phosphinoformato ligands, and obtained from insertion reactions of carbon dioxide into Sn–X bonds (X = N, H, P, respectively). The last chapter is devoted to X-ray structures of transition metal/tin CO2 complexes exhibiting metallocarboxylato ligands. As in Part 1, for each tin compound reported and when described in the original study, the structural descriptions are supplemented by synthetic conditions and spectroscopic data.

Graphical Abstract

1. Introduction

More than ever and everywhere in the world, the recovery and utilization of carbon dioxide have become a real priority, mobilizing many actors, from academic research to industry, involved in various domains and specialties. Long considered simply as abundant and chemically inert waste, but with devastating effects, carbon dioxide is now differently viewed by chemists, considered and accepted as a promising C1 building block, useful for a wide range of reactions and leading to a wide diversity of organic compounds. In this quest to transform CO2 into chemicals with higher added value, coordination and organometallic chemistry played—and continue to play—an important and determining role in facilitating its activation [1]. Since the first resolution of an X-ray crystal structure highlighted the metal coordination of a CO2 molecule—it was in 1975 that M. Aresta’s group published the characterization of a (carbon dioxide)bis(tricyclohexylphosphine) nickel complex [2]—advances and knowledges relative to carbon dioxide binding modes have continued to progress, and have proved particularly beneficial in catalysis and organic synthesis. Much later in 2019, for instance, Galindo et al. reported the X-ray structures of cis-M(C2H4)2(CO2)(PNP) compounds (M = Mo and W; PNP = 2,6-bis(diphenylphosphinomethyl)pyridine), which are the first examples of stable metal complexes with coordinated ligands of ethylene and carbon dioxide [3]. In conclusion of this recent work, the authors aim, as perspectives, for the conversion of C2H4 and CO2 into acrylate derivatives.
Since the 1990s, the chemical transformation of CO2 and its coordination on metal centers have been regularly reviewed and updated [4,5,6,7,8,9,10,11,12,13,14]. During the last decade, the reactivity of main group elements, and in particular metals and metalloids from group 14 (Si, Ge, Sn, Pb), toward small molecules also aroused a growing interest [15,16]. In this context and based on our previous work in this field, we have recently reviewed molecular organotin compounds bearing hemi-carbonate and carbonate ligands and which result from the reactivity of tin precursors towards carbon dioxide [1]. In this second part, we continue to explore from a structural point of view the molecular derivatives of tin which result from the insertion of carbon dioxide into the Sn–X bonds (X = N, H, P). Thus, the formation and structures of carbamato, formato and phosphinoformato tin derivatives are successively described and discussed. The last chapter is devoted to metallocarboxylato tin complexes exhibiting a bridging CO2 ligand. Although these latter compounds are not derived from the direct reactivity of tin precursors with carbon dioxide, we found appropriate to include them in this review. Regarding the general method used to carry out this inventory, the compounds described below were identified from the data available on the online portal of the Cambridge Structural Database (CSD) web interface (2017) [17]. In addition to the structural description, comparison tables including selections of structural and spectroscopic data are also displayed at the end of each section (Table 1, Table 2, Table 3, Table 4, Table 5, Table 6, Table 7 and Table 8).
Among the useful chemicals with higher added-value which can be produced by the transformation of CO2, the derivatives involving C–N bonds occupy an important place, historically and economically. Thus, from 1922, the German Bosch-Meiser process allowed the industrial use of carbon dioxide (together with ammonia) as raw material for the synthesis of urea [(NH2)2CO]. It is still today the most important industrial chemical process in terms of CO2 amounts converted (115 Mt) [18]. Since then, the use of carbon dioxide as C1 synthon has been extended to the synthesis of other types of nitrogen compounds such as oxazolidinones, quinazolinediones, ureas, carbamates, isocyanates and polyurethanes (Scheme 1). These molecules constitute intermediates and derivatives of great importance for the chemical, agrochemical and pharmaceutical industry. The domain was firstly reviewed by He et al. in 2012 [19], and recently updated by Song et al. [20] and then by Dalpazzo et al. [21], both in 2019. In most cases, the intervention of a catalytic precursor—organic or organometallic, is required to achieve the targeted product.

Tin-Promoted Insertion Reaction of CO2 into C–N Bonds

Regarding more specifically the use of main group metals as catalytic species for the insertion of CO2 into the C–N bond, the beneficial role of organotin compounds was regularly emphasized in the past, being the subject of several studies. In 2002, Tominaga and Sasaki found that di-n-butyltin oxide, n-Bu2SnO, was an effective catalyst for the dehydrating condensation of 1,2-aminoethanols with carbon dioxide to lead to the formation of 2-oxazolidinones [22] (Scheme 2). Reactions were carried out under 5 MPa of CO2, at 180 °C, for 16 h, in N-methyl-2-pyrrolidone (NMP) as solvent and gave satisfactory yields in the range of 53 to 94%. The use of NMP has been found to be more effective than the use of protic and non-polar solvents. From a mechanistic point of view, the authors proposed the formation of a seven-membered cyclic tin carbamate complex (Scheme 3) by the insertion of CO2 into the Sn–N bond of a stannaoxazolidine species and whose in-situ formation was supported by electrospray mass spectrometry. However, to our knowledge, such a species has not been yet structurally isolated. From this intermediate, the intramolecular elimination of one molecule of 2-oxazolidinone was suggested to explain the regeneration of the starting complex, n-Bu2SnO.
In 2013, Ghosh et al. revisited the previous reaction using chlorostannoxane derivatives, 1,3-dichloro-1,1,3,3-tetraalkyldistannoxanes (Scheme 4), as catalytic precursor, in methanol under 1.72 MPa of CO2 pressure and at 150 °C [23]. Chlorostannoxanes were selected because of (i) the presence of multi-active catalytic centers, and (ii) the ability to control Lewis acidity by changing the substituents on the metal centers. In comparison to the previous Tominaga’s study using n-Bu2SnO [22], an improvement of turnover numbers was recorded using chlorostannoxanes, from 9 to 138. Furthermore, it has been demonstrated that the activity of the catalytic precursors depends on the nature of the R and R′ substituents, the n-butyl derivative (R and R’ = n-Bu) being the most efficient.
The direct synthesis of carbamates from carbon dioxide has also aroused renewed interest in the past two decades [24]. The state of the art was well established in 2003 by Calderazzo et al. [25] and recently updated by Marchetti et al. [26]. Indeed, carbamates can be viewed as a potential alternative to access to isocyanates, which are historically employed for the production of polyurethanes (PU). Industrially, isocyanates are produced by reacting primary amines with phosgene, but handling of this reactant is not without risks and requires drastic precautions. Therefore, the recent route based on the thermal decomposition of carbamates into isocyanates can be considered as a safer and greener synthetic pathway [27,28].
In the past, several groups investigated the action of organotins as catalyst precursors for the direct conversion of CO2 into carbamates. In 2001, Sakakura et al. first published the synthesis of carbamates by reaction of amines and alcohols catalyzed by tin complexes (n-Bu2SnO and Me2SnCl2) [29] (Scheme 5). Optimal reaction conditions require a pressure of 30 MPa (300 bar) at 200 °C for 24 h. Moreover, the addition of acetals (2,2-diethoxy- or 2,2-dimethoxypropane) acting as a dehydrating agent significantly improves the conversion, and the increase in pressure promotes the urethane selectivity.
In 2016, Choi et al. reported the synthesis of N-phenylcarbamates from CO2 and aniline promoted by several di-n-butyltin dialkoxides: n-Bu2Sn(OMe)2, n-Bu2Sn(OEt)2, n-Bu2Sn(OPr-n)2, and n-Bu2Sn(OBu-n)2 (Scheme 6) [30]. The best yield of N-phenylcarbamate (41%) was obtained under 5 MPa (5 bar) of CO2 for 20 min at 150 °C using n-Bu2Sn(OMe)2 as tin precursor. The catalytic activity changes depending on the nature of the alkoxide substituents linked to the tin atom [n-Bu2Sn(OMe)2 > n-Bu2Sn(OPr-n)2 > n-Bu2Sn(OEt)2 > n-Bu2Sn(OBu-n)2]. N,N′-diphenylurea is formed as byproduct (4%). From a mechanistic point of view, the hemicarbonato tin complex, n-Bu2Sn(OMe)(OCO2Me), resulting from the reactivity of CO2 with n-Bu2Sn(OMe)2 [1,31], is claimed as the key intermediate of the reaction. In the case of reactions involving n-Bu2Sn(OBu-n)2, the authors also showed that the starting tin complex could be regenerated after the first catalytic run (by reacting with n-BuOH) and could thus be recycled without loss of activity.
Concomitantly and supported by density functional theory (DFT) calculations, Schaub et al. confirmed that dialkyltin(IV) dialkoxides can be used as reagents for the synthesis of aromatic carbamates from aromatic amines and carbon dioxide [32]. The process takes place between 60 and 70 bar of pressure, at 135 °C and in pentane as solvent. A NMR yield of 92% was recorded for the synthesis of methyl phenylcarbamate from aniline using three molar equivalent of n-Bu2Sn(OMe)2. Neither methylaniline nor diphenylurea was chromatographically detected. Extension of the reaction conditions to the synthesis of diurethanes from 4,4′-methylenedianiline and 2,4-diaminotoluene was successful with yields of 60 and 77%, respectively. In the perspective of an industrial development, the authors have demonstrated the possibility of regenerating and re-using the dialkyltin(IV) dialkoxides, by using an excess of dimethylcarbonate. Shortly after, in a new study, the same group improved the reaction yield (up to 97%) by combining with dialkyltin oxide, used as a tin precursor, tetraalkyl orthosilicates, which promote the generation of dialkyltin(IV) dialkoxides [33]. This positive effect had been initially shown for the tin-promoted synthesis of organic carbonates from carbon dioxide and alcohols [34].

2. Carbamato Tin Complexes

2.1. Early Works

Historically, the pioneering works on carbamates of tin derivatives dates back to the 1960s and are to be credited to A. J. Bloodworth and A. G. Davies who conducted extensive and fundamental investigations applying a systematic approach. From 1965, they described the synthesis of N-organo-N-trialkylstannylcarbamates resulting from the addition of trialkyltin alkoxides to isocyanates [35]. The reactions were reported as rapid and exothermic and taken place at room temperature (Scheme 7). Alkyl and aryl isocyanates exhibit a comparable reactivity.
Additionally in 1965, the study was then extended to bis(trialkyltin) oxides, which react exothermically with isocyanates to lead to trialkyltin N-trialkylstannylcarbamates (Scheme 8) [36].
From a coordination point of view, the authors preferentially recommended the representation of the carbamato ligand as N-coordinated to tin, in amido form, but did not exclude an O-coordination, in imido form, via an interconversion process. Subsequently, this structural consideration was the subject of several subsequent works, involving other research groups, and based on studies carried out by 119Sn Mössbauer spectroscopy [37], as well as more recently by computational calculations [38].

2.2. Reactivity

In terms of reactivity, N-organo-N-trialkylstannylcarbamates have been also involved in various organic reactions. In their initial article [35], Bloodworth and Davies showed in particular their reactivity towards acetic anhydride, ethylamine and ethanol, which lead, according to a metathetical process, to the corresponding urethanes. In 1968, the same authors reported the decarboxylation of trialkyltin esters of N-organo-N-trialkylstannylcarbamic acids by isocyanates and isothiocyanates providing an alternative route to the synthesis of carbodi-imides [39]. In 1992 and then in 1993, Shibata et al. underlined the positive and efficient role of methyl ester of N-ethyl-N-tributylstannylcarbamic acid as promoter for the Darzens reaction [40] and the addition of Michael [41] (Scheme 9).

2.3. X-Ray Crystal Structures

Regarding the solid-state characterization of metal carbamates by single-crystal X-ray diffraction, the area was first reviewed by Calderazzo et al. [25] and then recently updated by Marchetti et al. [26]. To date, numerous examples of metal carbamates have thus been isolated and characterized, demonstrating a rich coordination chemistry and highlighting a great diversity of structures. Herein, we will focus exclusively on X-ray structures involving tin moieties. To the best of our knowledge, six CSD depositions of carbamato tin complexes have been identified up to now (CSD entries and deposition numbers are reported in Table 1).
The first X-ray crystallographic investigation on a tin complex bearing a carbamato ligand was reported by Zakharov et al. in 1980, as part of studies on organotin isocyanate derivatives [42]. Thus, the Russian group characterized at the solid-state the structure of the trimethyltin(IV) ester of N-trimethylstannylcarbamic acids, Me3SnNHC(=O)–OSnMe3 (1), which describes a polymeric arrangement displaying a zigzag one-dimensional infinite chain along the c-axis (Figure 1). The organization of 1 consists of SnMe3 moieties bridged by carbamate dianions. Two distinct sites of tin atoms are distinguished: (i) those located in the main chain exhibiting a trigonal bipyramidal geometry (TBP) whose equatorial plane is occupied by three methyl substituents and the apical positions by two oxygen atoms; (ii) SnMe3 pendant groups, connected to the main chain by nitrogen atoms (Sn–N = 2.04(2) Å) and positioned in a syndiotactic arrangement, and whose tin atoms exhibit a tetrahedral geometry. This type of architecture is comparable to the chain-like structures reported for the organotin(IV) selenite complexes, (Me3Sn)2SeO3·H2O and (Ph3Sn)2SeO3 [43], as well as for the organotin(IV) carbonato complexes, (Me3Sn)2CO3 and (i-Bu3Sn)2CO3 [44]. However, the authors pointed out that, despite the presence of a hydrogen atom on the nitrogen of carbamato groups, no hydrogen bonds are observed between the chains, which is explained by the proximity of bulky Me3Sn groups. Moreover, we consider that compound 1 is the only example of an N-coordinated tin carbamate, structurally characterized to date.
Thereafter, with the aim of applications in heterogeneous catalysis, Calderazzo et al. reported the preparation of two N,N-dialkylcarbamato tin(IV) complexes, Sn(O2CNi-Pr2)4 (2) and Sn(O2CNEt2)4 (3), which were used as precursors for the chemical implantation of metal cations on a silica support [45]. The grafting method consisted of promoting reactivity of N,N-dialkylcarbamates with silanol groups of silica. Compounds 2 and 3 were synthesized by treating anhydrous Sn(IV) chloride with diethylamine and di-iso-propylamine, respectively, in toluene at room temperature, under atmospheric pressure of carbon dioxide (Scheme 10). Yields higher than 90% are reported.
Recrystallized from heptane, suitable single crystals of 2 were analyzed by X-ray diffraction leading to the resolution of the crystallographic structure. In 2002, in the frame of metal-organic chemical vapor deposition (MOCVD) investigations, Molloy et al. described a new synthetic route to 3, starting from a solution of tetrakis(N,N-diethylamino)tin(IV) complex in hexane that was bubbled with carbon dioxide (Scheme 11) [46].
X-ray structures of 2 and 3 revealed comparable mononuclear molecules. In both compounds, the tin atom is eight-coordinated by four chelating N,N-dialkylcarbamato ligands, which are exclusively O-donor (Figure 2). For 3, the authors suggest that the geometry can be viewed as a distorted square antiprism. From a spectroscopic point of view, similar 119Sn NMR chemical shifts were also recorded for 2 and 3, at −920.8 and −930.0 ppm, respectively (Table 2). In 13C NMR, the carbamato carbons show one resonance at 164.9 ppm for 2, and 166.2 ppm for 3. In the case of 2, CP/MAS 13C NMR measurements confirm the chemical shift recorded in solution. A thermogravimetric analysis was also carried out on 3 showing a continuous weight loss between room temperature and 300 °C (and resulting in the recovery of SnO2 as final residue). Beyond their interest as precursors for the deposition of tin oxide films [4,5], tin tetracarbamates were also used as versatile reactants for various organometallic and organic syntheses leading to alkoxystannanes [47], tetraalkynylstannanes [48] (Scheme 12), and O-silylurethanes [49].
In 2010, Kemp et al. reported the reactivity of bis(2,2,5,5-tetramethyl-2,5-disila-1-azacyclopent-1-yl)tin toward CO2 leading to a silyl-substituted carbamate of tin(II) [50]. The exposure of {[(CH2)Me2Si]2N}2Sn in solution in pentane under 60 psig of CO2 (4 bar) at room temperature causes a rapid reaction involving a color change from red-orange to yellow. In parallel, GC–MS investigations on the reaction medium have shown the co-formation of carbon monoxide and the cyclic ether 2,2,5,5-tetramethyl-2,5-disila-1-oxacyclopentane, [(CH2)Me2Si]2O (Scheme 13). Sn4(μ4-O){μ2-O2CN[SiMe2(CH2)]2}4(μ2-N=C=O)2 (4) was isolated as single crystals and analyzed by X-ray diffraction analysis. The unprecedented structure of 4 can be described as a tetranuclear tin cluster based on a [μ4-O]Sn4 core and including in periphery two bridging –N=C=O isocyanates and four bridging –O2CN[SiMe2(CH2)]2 carbamato ligands (Figure 3). The presence of the central atom μ4-O was explained as resulting from the cleavage of CO2 and not from O2 or H2O. Furthermore, the reaction of {[(CH2)Me2Si]2N}2Sn was also envisaged with OCS and CS2 leading, respectively, to an insoluble polymeric material and to a trinuclear cluster based on the coordination of three {[(CH2)Me2Si]2N}2Sn moieties by one CS2 molecule acting as μ3 ligand. Previously, during the 1990s, Sita et al. explored already the reactivity of Ge and Sn complexes of silyl-containing amides toward carbon dioxide [51]. In particular, they focused on the reaction involving [(Me3Si)2N]2Sn with CO2, which led to the isolation of the dimeric bisalkoxide complex, [Sn(OSiMe3)2]2. However, in this case and compared to the Kemp’s report, the tin carbamate is not stable but leads to the formation of [Sn(OSiMe3)2]2, trimethylsilyl isocyanate, and 1,3-bis(trimethylsilyl)carbodiimide, respectively, according to a ligand metathesis process.
In 2011, continuing their study on the reactivity of silyl-substituted tin(II) amides towards heteroallenes, Kemp et al. reported the interaction of CO2, OCS and CS2 with (Me2N)2Sn leading to the formation of insertion products characterized as bis-(N,N-dimethylcarbamato)tin(II), bis(N,N-dimethylthiocarbamato)tin(II), and bis(N,N-dimethyldithiocarbamato)tin(II), respectively [52]. Thus, (Me2N)2Sn in hexane solution reacts quickly with carbon dioxide, at room temperature and under atmospheric pressure, to give with a yield almost quantitative, the new complex [(Me2NCO2)2Sn]2 (5) (Scheme 14). The authors describe the reaction as exothermic. In the solid-state, compound 5 is organized as a dimer involving two types of carbamato ligands: bridging and terminally (η2-chelating) coordinated to Sn. From a supramolecular point of view, the existence of intermolecular interactions Sn···O [2.8499(18) Å] between neighboring dimers and involving one oxygen atom of terminal carbamato ligands, leads to the propagation of a polymer chain (Figure 4). The coordination geometry of tin atoms can be viewed as a highly distorted trigonal bipyramid. In solution, the 119Sn NMR spectrum exhibits only one resonance at δ –613 ppm. In the 13C {1H} NMR spectrum, the carbamato carbon also displays only one signal, not making it possible to differentiate the bridging and terminal modes highlighted in crystals of 5.
To our knowledge, the latest tin carbamate structure resolved to date by X-ray diffraction is to be credited to Fulton et al., who published in 2011, the reactivity of β-diketiminate tin derivatives with carbon dioxide [53]. [HC{(Me)CN(2,6-i-Pr2C6H3)}2]SnNi-Pr2 and [HC{(Me)CN(2,6-i-Pr2C6H3)}2]SnNH(2,6-i-Pr2C6H3), dissolved in toluene, react with CO2 (1 atm) at room temperature yielding to the corresponding carbamato tin(II) complexes, [HC{(Me)CN(2,6-i-Pr2C6H3)}2]SnOC(O)Ni-Pr2 (6) and [HC{(Me)CN(2,6-i-Pr2C6H3)}2]SnOC(O)NH(2,6-i-Pr2C6H3) (7), respectively (Scheme 15). In the solid-state, 6 consists of a mononuclear complex describing a trigonal pyramidal arrangement of the ligands around the tin atom and in an endo configuration. The carbamate moiety is terminally O-bonded to Sn [Sn−O = 2.1346(16) Å] and its orientation can be viewed almost perpendicular to the NCCCN plane of the β-diketiminate ligand (Figure 5). This mode of coordination is unusual for carbamates and is similar to that observed in the case of hemicarbonato derivatives [1]. The X-ray structure of complex 7 has not been reported but the spectroscopic data are very similar to 6 (Table 2) and the authors mention that for both compounds, the insertion reaction is irreversible, even under reduced pressure.
Table 1. Comparison of selected structural parameters found in carbamates of organotin complexes.
Table 1. Comparison of selected structural parameters found in carbamates of organotin complexes.
CompoundsSn−O(C)
(Å)
Sn−N(C)
(Å)
C−O(Sn)
(Å)
C=O
(Å)
N−C(O2)
(Å)
O−C−O
(deg)
N−C−O
(deg)
CSD Entry
Deposition Number
Ref.
Me3SnNHC(=O)–OSnMe3 (1)2.23(2)
2.24(2)
2.04(2)1.27(4)
1.30(3)
na1.35(4)126(3)120(2)
114(3)
MESNCB1211377[42]
Sn(O2CNi-Pr2)4 (2)2.146(5)
2.217(4)
2.136(5)
2.195(4)
2.198(4)
2.135(5)
2.219(4)
2.126(5)
na1.283(7)
1.276(8)
1.279(7)
1.260(8)
1.268(8)
1.269(7)
1.275(8)
1.282(7)
na1.33(1)
1.352(9)
1.34(1)
1.333(8)
117.7(6)
117.3(6)
116.1(6)
117.2(6)
119.9(6)
122.4(6)
122.9(6)
119.7(6)
120.6(6)
122.2(6)
116.1(6)
121.2(6)
MULPUE
196069
[45]
Sn(O2CNEt2)4 (3)2.156(1)
2.209(1)
2.163(1)
2.196(1)
na1.292(2)
1.277(2)
1.298(2)
1.276(2)
na1.338(2)
1.331(2)
117.0(1)
116.6(1)
121.5(2)
121.7(1)
121.7(2)
UGOZUL
196788
[46]
Sn4(μ4-O){μ2-O2CN[SiMe2(CH2)]2}4(μ2-N=C=O)2 (4)2.331(2)
2.164(3)
2.226(3)
2.429(3)
2.716(4)
2.165(3)
2.379(2)
2.225(3)
2.458(2)
na1.274(58)
1.282(5)
1.272(5)
1.291(5)
1.273(6)
1.289(5)
1.301(6)
1.257(5)
na1.349(5)
1.345(5)
1.355(5)
121.0(4)
121.1(4)
121.6(4)
121.3(4)
117.6(4)
118.2(4)
120.3(4)
116.9(4)
121.5(4)
IFOXEH
664586
[50]
[(Me2NCO2)2Sn]2 (5)2.332(1)
2.235(2)
2.432(2)
2.115(2)
na1.272(3)
1.280(3)
1.263(3)
1.295(3)
na1.338(3)
1.336(3)
118.8(2)
121.8(2)
120.2(2)
121.0(2)
121.3(2)
116.9(2)
UYOBIU
806650
[52]
[HC{(Me)CN(2,6-i-Pr2C6H3)}2]SnOC(O)Ni-Pr2 (6)2.1346(16)na1.304(3)1.238(3)1.366(3)121.7(2)116.6(2)
121.7(2)
MIYHIN
791226
[53]
Table 2. Selection of spectroscopic data (NMR and IR) assigned to carbamates of tin complexes.
Table 2. Selection of spectroscopic data (NMR and IR) assigned to carbamates of tin complexes.
Compounds119Sn{1H} NMR
(δ, ppm)
13C{1H} NMR
-NC(O)O-
(δ, ppm)
IR
ν(C=O)
(cm−1)
Ref.
Sn(O2CNi-Pr2)4 (2)−920.8 a164.4 ana[45]
Sn(O2CNEt2)4 (3)−930.0 b166.2 b1612[46]
Sn4(μ4-O){μ2-O2CN[SiMe2(CH2)]2}4(μ2-N=C=O)2 (4)−316.0 b167.9bna[50]
[(Me2NCO2)2Sn]2 (5)−613.0 a164.5a1651[52]
[HC{(Me)CN(2,6-i-Pr2C6H3)}2]SnOC(O)Ni-Pr2 (6)−394.0 b161.75 b1595, 1575, 1552, and 1524[53]
[HC{(Me)CN(2,6-i-Pr2C6H3)}2]SnOC(O)NH(2,6-i-Pr2C6H3) (7)−398.0 b161.59 b1624, 1554, 1526, and 1517[53]
a Measured in chloroform-d; b measured in benzene-d6.

3. Formato Tin Complexes

3.1. Early Works

To our knowledge, pioneering work on formato tin complexes dates back to 1927. Elöd and Kolbach described a general method to prepare a series of Sn(II) compounds of formula M2Sn(O2CH)4·5H2O (M = Na, K, and NH4) by dissolving tin(II) chloride into solutions of alkali metal or ammonium formate, in formic acid (35% w/w) [54]. In 1964, Donaldson and Knifton revisited the synthetic protocol by reporting the preparation and analysis of alkali metal (K, Rb, and Cs) and ammonium derivatives of the triformatostannate(II) anion [Sn(O2CH)3] [55]. In 1969, Jelen and Lindqvist solved the crystal structure of potassium triformatostannate(II), KSn(HCO2)3 in which the tin atom is coordinated pyramidally by three oxygen atoms of three formate groups (CSD refcode: KTFORM, 1200243) [56]. Thereafter in 1974, Harrison and Thornton published an X-ray diffraction study devoted to the structure determination of Sn(O2CH)2 arranged in infinite two-dimensional sheets in which tin(II) atomes are bridged by formate groups (CSD refcode: SNFORM, 1260779) [57].
Solid-state structures of tin(IV) formate complexes were established from the 1980s. In 1987, Molloy et al. reported the crystal structure of triphenyltin formate prepared from a mixture of triphenyltin hydroxide and formic acid in refluxing toluene [58]. [Ph3Sn(O2CH)]n is organized in a polymeric chain propagating along c axis and compared by the authors to a flattened helix. Triphenyltin moieties are bridged by bidentate formate ligands (CSD refcode: FAJZIZ, 1151980; Figure 6). In 1990, Mistry et al. resolved the structure of bis(formato)dimethyltin(IV), [Me2Sn(O2CH)2]n, which was described as a sheet polymer with linear Me2Sn moieties nearly symmetrically bridged by formate anions. Sn atoms exhibit an octahedral geometry (CCDC ref.: SIFLEY, 1258788; Figure 7) [59]. More recently, Power et al. published the synthesis and characterization of several organotin(IV) formates prepared from hydrides and hydroxides of organotins in the presence of formic acid [60]. X-ray crystallographic structures investigations were conducted on five compounds. [n-Bu2Sn(O2CH)2]n (CCDC ref.: EKONEY, 783726; Figure 8), [(PhCH2)3Sn(O2CH)]n (CCDC ref.: EKONAU, 783725), and [Cy3Sn(O2CH)]n (Cy = cyclohexyl) (crude structure owing to the low quality of crystals) exhibit polymeric structures, while Mes3SnOC(O)H (Mes = 2,4,6-trimethylphenyl) (CCDC ref.: EKONOI, 783728) and Dmp3SnOC(O)H (dmp = 2,6-dimethylphenyl) (CCDC ref.: EKONIC, 783727; Figure 9) are monomeric tin formates. In 2010, McMurtie and Arnold reported the crystal structure of meso-tetraphenylporphyrinatotin(IV) diformate, prepared from Sn(TPP)(OH)2 (TPP = dianion of 5,10,15,20-tetraphenylporphyrin) and formic acid, and in which the central tin atom exhibits an octahedral geometry with two axially O-coordinated formato ligands (Figure 10) [61].

3.2. X-Ray Crystal Structures of Formato Tin Complexes Resulting from Reactivity with CO2

While the formation of the above compounds required the use of formic acid in their preparation, there are also a few examples of formato tin complexes directly isolated under carbon dioxide atmosphere, resulting from the insertion of CO2 into the Sn–H bond. In our opinion, six complexes, published mainly during the last decade and characterized by X-ray diffraction, correspond to this category of compounds (Table 3). In 2009, Roesky et al. reported the structural and spectroscopic characterization of several complexes resulting from hydrostannylation reactions of carbon dioxide, ketones, aldehydes, alkynes and carbodiimides with the tin(II) hydride precursor, [HC{(Me)CN(2,6-i-Pr2C6H3)}2]SnH [62]. In the case of the reaction involving CO2, bubbling at room temperature very quickly causes a change in color (from yellow to colorless) of the reaction mixture (toluene solution). The treatment with n-hexane then leads to the formation of a microcrystalline powder (at −32 °C) characterized as being the stannylene formate complex, [HC{(Me)CN(2,6-i-Pr2C6H3)}2]SnOC(O)H (8) (Scheme 16). The presence of the formate ligand evidenced by infrared was confirmed by single-crystal X-ray diffraction. The molecular structure shows moreover a distorted pseudo-tetrahedral environment for the tin atom, completed by one electron lone pair (Figure 11).
During the period 2007–2010, the Albertin’s group isolated several formato tin complexes by studying the reactivity of transition-metal stannyl complexes with carbon dioxide. Three of these complexes were characterized by single-crystal X-ray diffraction studies. Firstly, the reaction of Re(SnH3)(CO)2[P(OEt)3]3, in solution in benzene and at room temperature, with CO2 (under atmospheric pressure) and during 24 h led to the binuclear OH-bridging bis(formate) complex [Re{Sn[OC(O)H]2(μ-OH)}(CO)2{P(OEt)3}3]2 (9) (Scheme 17) [63]. The authors described the core of 9 as a dimeric complex in which two octahedrally coordinated rhenium units are joined by a bridging tetraformatebis(μ-hydroxo)ditin moiety. Tin atoms exhibit a highly distorted square pyramidal coordination geometry, bearing two η1-formate ligands per metal atom. The central Sn2O2 distannoxane ring is planar (Figure 12). Infrared and 13C{1H} NMR spectra attest also the presence of formates by showing two bands at 1674–1594 cm−1OCO, asym) and one resonance at δ 166 ppm [OC(O)H], respectively (Table 4).
In 2008, the same group reported a similar structure by standing the ruthenium-cyclopentadienyl-trihydridostannyl complex, Ru(SnH3)(Cp)[P(OEt)3](PPh3), under a CO2 atmosphere for 5 h (Scheme 18) [64]. The resulting formate species was characterized as [Ru[Sn{OC(O)H}2(μ-OH)](Cp){P(OEt)3}(PPh3)]2 (10), exhibiting a structure based on two μ-OH groups bridging the tin atoms (Figure 13).
In 2010, Albertin et al. studied the reactivity of hydride-trihydridestannyl complexes MH(SnH3)P4 [M = Fe, Os; P = P(OEt)3, PPh(OEt)2] toward carbon dioxide leading to solids characterized as hydroxobis(formate)stannyl derivatives, MH[Sn(OH){OC(O)H}2]P4 [65]. Reactions took place at room temperature, in solution in toluene and under atmospheric pressure of CO2 (Scheme 19). Starting from OsH(SnH3)(PPh(OEt)2)4, the compound OsH[Sn(OH){OC(O)H}2][PPh(OEt)2]4 (11) was obtained and its molecular structure was established by single-crystal X-ray diffraction showing the insertion of two CO2 molecules into two Sn–H bonds (Scheme 19). The Os–H bond remains intact. The hydride ligand is unequivocally demonstrated by 1H NMR spectroscopy showing a multiplet at δ −11 ppm. The osmium atom is hexacoordinated according to a distorted octahedral geometry bearing four phosphonite ligands, one hybride ligand, and one hydroxobis(formate) stannyl ligand. The hydride and stannyl ligand are in cis positions. Regarding the coordination of the tin atom, Sn is tetracoordinate displaying a tetrahedral geometry (Figure 14).
In 2016, Strohmann et al. published the crystal structure of diaquadi-μ-hydroxido-tris[trimethyltin(IV)] diformatotrimethylstannate(IV), [(Me3Sn)3(μ-OH)2(H2O)2][Me3Sn{OC(O)H}2] (12), obtained as byproduct from trapping reactions with trimethyltin chloride and formed from atmospheric CO2 [66]. The structure of 12 consists of a short trimethyltin hydroxide chain (tristannoxane), positively charged, and a bisformatostannate, negatively charged (Figure 15). From a supramolecular point of view, formate ligands bridge four cationic tristannoxanes via hydrogen-bonding interactions leading to a two-dimensional network.
Lately, in 2017, Jones et al. reported the efficient and highly selective use of the two-coordinate amido-tin(II) hydride complex, LSnH (L = -N(C6H2-i-Pr{C(H)Ph2}2-4,2,6)(Si-i-Pr3)), acting as catalytic species for the reductive hydroboration of CO2 into methanol equivalents, MeOBR2 (2-methoxy-4,4,5,5-tetramethyl-1,3,2-dioxaborolane and catecholatoboryl methoxide), using boranes, HBR2, as hydrogen sources [67]. The tin formate complex, LSn(κ2-O,O’-O2CH) (13), was thus isolated during catalytic studies and also by exclusively reacting LSnH with CO2 (Scheme 20) or alternatively from LSnBr with potassium formate, KOC(O)H. Compound 13 was crystallographically characterized and consists of a monomeric complex in which the tin atom is unusually chelated by a formate ligand (Figure 16).
Table 3. Comparison of selected structural parameters found in formates of organotin complexes.
Table 3. Comparison of selected structural parameters found in formates of organotin complexes.
CompoundsSn−O(C)
(Å)
C−O(Sn)
(Å)
C−O
(Å)
Sn−O−C
(deg)
O−C−O
(deg)
CSD Entry
Deposition Number
Ref.
[HC{(Me)CN(2,6-i-Pr2C6H3)}2]SnOC(O)H (8)2.1353(15)1.299(2)1.209(3)116.40(13)126.78(18)COPLOJ
709476
[62]
[Re{Sn[OC(O)H]2(μ-OH)}(CO)2{P(OEt)3}3]2 (9)2.040(5)
2.071(7)
1.29(1)
1.26(1)
1.19(2)
1.21(1)
123.1(6)
132.2(6)
126(1)
128.8(9)
CIHYUO
655465
[63]
[Ru[Sn{OC(O)H}2(μ-OH)](Cp){P(OEt)3}(PPh3)]2 (10)2.210(2)
2.069(2)
1.262(4)
1.274(5)
1.216(5)
1.205(6)
128.7(2)
133.5(2)
127.9(3)
125.7(4)
KOMBUK
708897
[64]
OsH[Sn(OH){OC(O)H}2][PPh(OEt)2]4 (11)2.061(5)
2.089(5)
1.16(1)
1.25(1)
1.11(1)
1.21(1)
132.8(7)
124.9(7)
131(1)
124(1)
ALAGUQ
813529
[65]
[(Me3Sn)3(μ-OH)2(H2O)2][Me3Sn{OC(O)H}2] (12)2.2991(17)
2.2990(17)
1.269(3)1.224(3)125.94(17)128.1(3)EWIGIC
1505528
[66]
N(C6H2-i-Pr{C(H)Ph2)2-4,2,6)(Si-i-Pr3)Sn(κ2-O,O’-O2CH) (13)2.353(2)
2.333(2)
1.109(4)
1.310(4)
na91.4(2)
87.5(2)
125.6(4)ACEQUX
1518144
[67]
Table 4. Selection of spectroscopic data (NMR and IR) assigned to formates of tin complexes.
Table 4. Selection of spectroscopic data (NMR and IR) assigned to formates of tin complexes.
Compounds119Sn{1H} NMR
(δ, ppm)
13C{1H} NMR
−OC(H)O (δ, ppm)
1H
NMR
−OC(H)O (δ, ppm)
IR
(cm−1)
Ref.
[HC{(Me)CN(2,6-i-Pr2C6H3)}2]SnOC(O)H (8)−360 a166.92 a8.97
3J(119Sn,H) = 52 Hz
1641
(νC=O)
2700
(νC−H)
[62]
[Re{Sn[OC(O)H]2(μ-OH)}(CO)2{P(OEt)3}3]2 (9)−487.7 b166.4 b9.21 bc1665, 1594
(νCHO)
[63]
[Ru[Sn{OC(O)H}2(μ-OH)](Cp){P(OEt)3}(PPh3)]2 (10)−263.5 cd166.1 d8.36 d1660, 1627
(νCHO)
[64]
OsH[Sn(OH){OC(O)H}2][PPh(OEt)2]4 (11)−451.0 b165.5 e8.93 b1673, 1634
(νCHO)
[65]
N(C6H2-i-Pr{C(H)Ph2)2-4,2,6)(Si-i-Pr3)Sn(κ2-O,O’-O2CH) (13)−134.0 a173.80 a8.71 a1549
(νCHO)
[67]
a Measured in benzene-d6; b measured in toluene-d8; c measured at −70 °C; d measured in acetone-d6; e determined for OsH[Sn(OH){OC(O)H}2][P(OEt)3]4.

4. Phosphinoformato Tin Complexes

4.1. Foreword

This type of compound, rarely described so far, results from the insertion of CO2 into the Sn–P bond then leads to the formation of the Sn–OC(O)P moiety which can be qualified as a phosphinoformato derivative, by analogy with phosphino formic acid (H2PCOOH) according to the recent work published by Kaiser et al. [68]. To the best of our knowledge, only two examples of this type of complex have been described so far (Table 5), and their characterization is quite recent (last decade).
Table 5. Selection of structural parameters found in complexes 14 and 15.
Table 5. Selection of structural parameters found in complexes 14 and 15.
CompoundsSn–O(C)
(Å)
P–C(O2)
(Å)
C–O(Sn)
(Å)
C–Oexo
(Å)
Sn–O–C
(deg)
O–C–O
(deg)
O–C–P
(deg)
CSD Entry
Deposition Number
Ref.
[(i-Pr2P)2N·CO2]2Sn (14)2.344(4)
2.344(5)
1.875(7)
1.867(6)
1.272(8)
1.273(8)
1.210(8)
1.231(7)
128.2(5)
132.2(4)
126.8(6)
129.5(6)
114.5(5)
116.0(5)
117.8(5)
115.4(5)
RANBUF
831553
[69]
(F5C2)3SnCH2P(t-Bu)2·CO2 (15)2.233(2)1.891(6)1.268(4)1.216(4)122.9(2)128.1(3)114.9(2)
117.0(3)
FONREI
1892689
[70]

4.2. X-Ray Crystal Structures

The first example of the phosphinoformato tin complex, characterized by X-ray diffraction as [(i-Pr2P)2N·CO2]2Sn (14), was resolved in 2011 by Kemp et al. [69]. Complex 14 was isolated as a white precipitate by bubbling CO2 at room temperature (for 10 min) through a pentane solution of [(i-Pr2P)2N]2Sn, which caused a color change of the solution from orange to yellow (Scheme 21). Structural analysis of single-crystals confirmed the capture of two molecules of CO2 by [(i-Pr2P)2N]2Sn forming a six-membered ring complex in which CO2 is inserted into one Sn–P bond of each ligand (Figure 17). Moreover, the authors studied the stability of 14 in particular by thermogravimetric analysis and showed the easy release of CO2 from 90 °C. They showed also the possibility of recovering the starting complex by simply heating a solid sample of 14 under argon. Compound 14 is stable for several months stored under CO2 at room temperature or under argon at −25 °C. On the basis of these observations, the incorporation of CO2 as an adduct is preferentially considered rather than as a real insertion.
More recently, in 2019, Mitzel et al. reported the reactivity of the geminal frustrated Lewis pair (F5C2)3SnCH2P(t-Bu)2 with a variety of small molecules and, in particular, with CO2 [70]. The formation of the CO2 adduct of (F5C2)3SnCH2P(t-Bu)2, characterized as (F5C2)3SnCH2P(t-Bu)2·CO2 (15), was first demonstrated by NMR at −70 °C in THF-d8 (Table 6). The authors mention the instability of the complex at room temperature causing the release of CO2 (Scheme 22). The CO2 molecule is bonded to tin and phosphorus atoms, thus leading to the formation of a five-membered ring with an exocyclic C–O bond. The PCO2 unit is planar (Figure 18). Other adducts resulting from the reactivity of (F5C2)3SnCH2P(t-Bu)2, in particular with SO2 and CS2, were also isolated in the solid state and describe comparable solid-state structures.
Table 6. Selection of spectroscopic data (NMR and IR) assigned to phosphinoformato of tin complexes.
Table 6. Selection of spectroscopic data (NMR and IR) assigned to phosphinoformato of tin complexes.
Compounds119Sn{1H} NMR
(δ, ppm)
13C{1H} NMR
–OC(O)P
(δ, ppm)
31P{1H} NMR
–OC(O)P
(δ, ppm)
IR
(cm−1)
Ref.
[(i-Pr2P)2N·CO2]2Sn (14)−184 a
(t, 1JSn-P = 2626 Hz)
168.7 a
(d, 1JC-P = 95 Hz, PCO2)
26.4 a
(s, broad)
1629
(νC=O)
1206
[69]
(F5C2)3SnCH2P(tBu)2·CO2 (15)−308.3 bc
(hept, very broad, 2JSn,F = 368 Hz).
162.3 bd
(d, 1JC-P = 80 Hz, PCO2)
31.8 bd
(s, 2JSn,P = 56 Hz)
na[70]
a Measured in benzene-d6; b measured in tetrahydrofuran-d8; c determined at 208 K; d determined at 203 K.

5. Metallocarboxylato Tin Complexes

5.1. Foreword

The last part of this collection is devoted to bimetallic transition-metal/tin complexes with bridging carbon dioxide, of general formula M(CO2)Sn, also named metallocarboxylato tin complexes. Compared to the categories previously described, their synthesis does not involve the direct use of CO2 but is based on the reactivity between metallocarboxylic acids and halogenated precursors of organotins. However, we have found it interesting and complementary to also include this class of compounds in this review. Metallocarboxylato complexes and in particular tin derivatives began to be studied in the mid-1980s, often considered as suitable models of intermediates for the catalytic conversion of carbon dioxide [4]. Most of the structures shown below were resolved in the 1990s. The D. Gibson’s group in Louisville (USA) was one of the most active in this field characterizing several of these compounds. Metallocarboxylato-tin complexes have also been the subject of specific infrared considerations [71]. Indeed, νOCO adsorption bands constitute well-suited probes for orienting and determining the coordination modes of the CO2 ligand.

5.2. X-Ray Crystal Structures

To the best of our knowledge, in 1987, Gladysz et al. reported the first X-ray characterization of a transition metal/tin bridging CO2 complex by isolating (η-C5H5)Re(NO)(PPh3)(CO2SnPh3) (16) [72]. Compound 16 was obtained in high yield (86%) by reacting (η-C5H5)Re(NO)(PPh3)(CO2K) with 1.1 equivalent of Ph3SnCl at −78 °C in THF (Scheme 23). Yellow crystal prisms of 16 grew from a mixture of CH2Cl2/hexane. Compound 16 is air- and water-stable, and could also be alternatively prepared from (η-C5H5)Re(NO)(PPh3)(CO2H) and (Ph3Sn)2O at 25 °C in THF. Moreover, the decarboxylation of 16 occurs by heating at 180 °C for 10 min (from solid) or at 140 °C, for 20 h in solution in xylene. Using Me3SnCl, the authors claim the formation of the analogue (η-C5H5)Re(NO)(PPh3)(CO2SnMe3) with a yield of 95%. The study was also extended to the synthesis of rhenium/germanium and rhenium/lead bridging CO2 complexes using Ph3GeBr and Ph3PbCl, as precursors, respectively. The crystallographic structure of 16 confirms the presence of two moieties based on rhenium and tin, respectively, linked by a CO2 bridging ligand. The tin atom of 16 adopts a distorted trigonal bipyramid geometry in which the equatorial positions are occupied by the two oxygen atoms of the CO2 ligand and supplemented by a phenyl group (Figure 19). The CO2 ligand can be regarded as a carboxylato fragment, which is corroborated by infrared analysis, bidentally bound to the tin atom and defined as μ(n1-C: n2-O,O′) ligand.
In 1991, Gibson et al. published the solid state structure of (η-C5H5)Fe(CO)(PPh3)(CO2SnPh3) (17) prepared by mixing, under nitrogen and at low temperature, Ph3SnCl to a THF solution of CpFe(CO)(PPh3)COOK·3H2O [73]. The structure was found to be isomorphous to 16, with the tin atom occupying a trigonal bipyramid arrangement (Figure 20). Compared to the Gladysz complex, the authors underline for 17 a greater difference between the lengths of the Sn–O bonds, corresponding to a value of 0.219 Å (Table 7). However, in the literature devoted to triaryltin organocarboxylate monomers this difference is generally much more marked in the range of 0.48 to 0.81 Å [74].
Two years later in 1993, the same group reported the solid state isolation of a new Fe/Sn CO2-bridged complex, which was characterized by single crystal X-ray diffraction analysis as the indenyl derivative (η5-C9H7)Fe(CO)(PPh3)(CO2SnPh3) (18). Compound 18 was prepared under nitrogen, in THF solution and at 273 K, from (η5-C9H7)Fe(CO)2(PPh3)+·I and Ph3SnCl [75]. As shown by the structure displayed in Figure 21, the Sn atom geometry consists of a distorted trigonal bipyramid. However, the steric hindrance of the indenyl ligand and the resulting interactions with the phosphine ligand are suspected to cause notable structural differences compared to complexes 16 and 17: (i) Sn–O bond lengths differ from 0.467(3) Å; (ii) the difference in the carboxyl C–O bond lengths also increases [0.100(4) Å]; (iii) the Fe–C–O bond angles exhibit higher distortions [10.2(3)°] (Table 7).
The Gibson’s group continued to be very prolific in this field until the end of the 1990s, structurally characterizing several new iron and rhenium/tin CO2 bridged complexes. In 1994, the reaction of Cp*Re(CO)(NO)COOH with Ph3SnCl in the presence of Na2CO3 afforded the formation of Cp*Re(CO)(NO)(CO2SnPh3) (19) showing a μ2-η3 coordination mode for the CO2 ligand (Figure 22) [76]. As reflected by the lengths of the Sn–O and C–O bonds (Table 7), an asymmetric bonding mode of carboxylic oxygens to carbon and tin atoms was observed.
In 1995, Gibson et al. resolved by X-ray diffraction, the structure of four new CO2-bridged transition metals/tin complexes [77]. CpFe(CO)(PPh3)(CO2SnMe3) (20) and CpFe(CO)(PPh3)(CO2Sn(n-Bu)3) (21) were synthesized by reaction in THF between CpFe(CO)(PPh3)CO2K+ and ClSnMe3 and ClSnPh3, respectively. In both cases, the bridging carboxyl ligand adopts a μ2-η3 coordination mode in which the carboxyl atom is bound to the iron atom and both oxygens are bound to the tin atom. Cp*Fe(CO)2(CO2SnPh3) (22) was prepared from Cp*Fe(CO)3+BF4 and Ph3SnCl in CH2Cl2 at 0 °C in the presence of KOH. Compound 22 displays a similar structure to that of complexes 20 and 21 in which tin atoms are five-coordinated and adopt a distorted trigonal-bipyramidal geometry. The CO2-bridged rhenium-tin complex Cp*Re(CO)(NO)(CO2SnMe3) (23) was isolated as crystalline solid from Cp*Re(CO)(NO)COOH and ClSnMe3 in the presence of KOH. The X-ray structure of 23 highlighted a μ2-η2 coordination mode of the bridging CO2 ligand in which the carbonyl carbon and one oxygen atom are bonded to the rhenium atom and the tin atom, respectively. As a result, the tin atom adopts a distorted tetrahedral geometry (Figure 23).
In 1997, two new specimen of CO2-bridged rhenium/tin complexes were characterized at the solid-state by the Gibson’s group [78]. The transmetalation reaction involving Cp*Re(CO)(NO)(CO2SnMe3) (23) and Me2SnCl2 led to the formation in good yield of Cp*Re(CO)(NO)(CO2Sn(Cl)Me2) (24) jointly with Me3SnCl (Scheme 24). The X-ray structure of complex 24 shows the presence of a μ2-η3 CO2 ligand bridging the two moieties of rhenium and tin. The tin atom exhibits a five-coordinate geometry adopting an edge-capped tetrahedral arrangement rather than a distorted trigonal bipyramid.
An equimolar mixture of Cp*Re(CO)(NO)COOH and compound 24 leads, in the presence of water, to the formation of [Cp*Re(CO)(NO)(CO2)]2SnMe2 (25) via the formation of an hydroxyl species, suggested by the authors as being Cp*Re(CO)(NO)(CO2Sn(OH)Me2) (Scheme 25). Complex 25 was also isolated by reacting Cp*Re(CO)(NO)COOH with Me2SnCl2 in the presence of Na2CO3. The structure of 25 was described as two Cp*Re(CO)(NO) units linked to a SnMe2 moiety through two μ2-η3 CO2 ligands. The central tin atom is hexacoordinated and displayed a skewed trapezoidal-bipyramidal geometry (Figure 24).
In 1997, Komiya et al. reported the synthesis of the CO2 complex of iron(0), Fe(CO2)(depe)2 [depe = 1,2-bis(diethylphosphino)ethane)], resulting from the replacement of N2 in Fe(N2)(depe)2 with CO2. Fe(CO2)(depe)2 reacts then with Ph3SnCl in Et2O at −78°C to lead to the formation of FeCl(CO2SnPh3)(depe)2 (26) (Scheme 26) [79]. The same reaction, using Me3SnCl as tin precursor, gives rise to the analogous complex FeCl(CO2SnMe3)(depe)2. The X-ray structure of 26 shows that iron and tin atoms are linked by a CO2 molecule, acting as μ(n1-C: n2-O,O′) ligand (Figure 25).
During the last decade, two new metallocarboxylato of tin complexes have been characterized at the solid-state. First of all, in 2011, Adams et al. investigated the reaction of Os3(CO)12 with Ph3SnOH in the presence of [Bu4N][OH] in methanol, isolating two new trinuclear osmium clusters characterized as Os3(CO)10(μ-η2-O=COSnPh3)(μ-OMe) (27) and Os3(CO)10(μ-OMe)(μ-OH) (with yields of 18% and 7%, respectively) [80]. The tiphenylgermyl homologue of 27, Os3(CO)10(μ-η2-O=COGePh3)(μ-OMe), was also obtained from the reaction of Ph3GeOH with Os3(CO)12 and using a comparable synthetic protocol (in the presence of [Bu4N][OH] and in methanol). The authors described the structure of 27 as being a trinuclear cluster of three osmium atoms bearing one bridging methoxy ligand and one bridging triphenlystannylcarboxylate ligand. Remarkably, the SnO distances show a large difference in length (2.075 (5) Å and 2.858 (5) Å, respectively), which results from an unusual mode of coordination showing the organotin fragment in a pendant arm arrangement. The carbon atom and one oxygen atom of O=COSnPh3 are bonded to two distinct osmium atoms, while the tin atom is connected via the second oxygen atom of the ligand (Figure 26).
Finally, more recently and during investigations devoted to the synthesis and isolation of platinum carbonyl clusters stabilized by Sn(II)-based fragments, Zacchini et al. solved in 2016 the X-ray structure of [NEt4]4[Pt9(CO)8(SnCl2)3(SnCl3)2(Cl2SnOCOSnCl2)]·2.5CH3COCH3 (28), which contains an unusual [(Cl2SnOCOSnCl2)]2− ligand described as bis-stannyl-carboxylate or carbon dioxide μ3: κ3-C,O,O′-CO2 [81]. Cluster 28 was obtained as a side-product from the reaction at room temperature in acetone, involving [NEt4]4[Pt15(CO)30], [NEt4]Cl and a large excess of SnCl2·2H2O, and leading to the major product [NEt4]4[Pt6(CO)6(SnCl2)2(SnCl3)4]. The authors described the unprecedented structure of 28 as consisting of a formally neutral Pt9 core bonded to eight carbonyls, three μ4-SnCl2, two μ3-[SnCl3]- and the six electron donor bis-stannyl-carboxylate [(Cl2SnOCOSnCl2)]2− ligand (Figure 27). The possible formation of the metallocarboxylate group is explained by the nucleophilic addition of H2O to a terminal CO ligand, followed by deprotonation. Alternatively, the authors also suggest that [(Cl2SnOCOSnCl2)]2− could be considered as a carbonite ion [CO2]2− [82], stabilized by coordination at one Pt atom and two Sn atoms. A selection of spectroscopic data (NMR and IR) relating to metallocarboxylato-tin complexes is reported in Table 8.
Table 7. Selection of structural parameters found in metallocarboxylato tin complexes.
Table 7. Selection of structural parameters found in metallocarboxylato tin complexes.
CompoundsSn−O(C)
(Å)
M−C(O2)
(Å)
C−O
(Å)
O−C−O (deg)O−Sn−O (deg)CSD Entry
Deposition Number
Ref.
(η5-C5H5)Re(NO)(PPh3)(CO2SnPh3) (16)2.257(7)
2.175(7)
2.058(9)1.269(11)
1.313(11)
112.2(8) 57.8(2)FODLAL
1158240
[72]
(η5-C5H5)Fe(CO)(PPh3)(CO2SnPh3) (17)2.123(4)
2.342(4)
1.931(5)1.270(6)
1.305(6)
113.4(4)57.4(1)JIDLUD
1185927
[73]
(η5-C9H7)Fe(CO)(PPh3)(CO2SnPh3) (18)2.069(2)
2.536(2)
1.933(3)1.236(3)
1.336(3)
115.9(3)55.42(7)HADNUV
1171252
[75]
Cp*Re(CO)(NO)(CO2SnPh3) (19)2.092(3)
2.399(1)
2.100(9)1.24(1)
1.322(9)
114.6(7)56.79(5)YEGCUI
1300837
[76]
CpFe(CO)(PPh3)(CO2SnMe3) (20)2.476(3)
2.089(3)
1.936(4)1.260(6)
1.321(5)
114.3(3)55.94(9)YOTBEO
1305570
[77]
CpFe(CO)(PPh3)(CO2Sn(n-Bu)3) (21)2.105(4)
2.432(4)
1.934(6)1.266(7)
1.316(7)
114.2(5)56.5(1)YOTBIS
1305571
[77]
Cp*Fe(CO)2(CO2SnPh3) (22)2.102(2)
2.394(2)
1.956(3)1.252(3)
1.312(3)
114.5(2)56.90(6)YOTBOY
1305572
[77]
Cp*Re(CO)(NO)(CO2SnMe3 (23)2.054(4)
2.806(4)
2.103(5)1.237(6)
1.311(6)
117.0(5)50.3(1)YOTBUE
1305573
[77]
[Cp*Re(CO)(NO)(CO2)]2SnMe2 (24)2.079(6)
2.524(8)
2.11(1)1.26(1)
1.31(1)
116.8(9)55.7(2)NOXRUN
1223123
[78]
Cp*Re(CO)(NO)(CO2Sn(Cl)Me3) (25)2.109(8)
2.287(8)
2.11(1)1.24(1)
1.32(1)
115.4(9)58.8(3)NOXROH[78]
FeCl(CO2SnPh3)(depe)2 (26)2.097(6)
2.312(7)
1.87(1)1.28(1)
1.32(1)
108.4(9)57.0NOMQIP
1222229
[79]
Os3(CO)10(μ-η2-O=COSnPh3)(μ-OMe) (27)2.076(5)2.080(6)1.277(8)
1.295(8)
115.6(8)116.0(4)IXODAB
807088
[80]
[NEt4]4[Pt9(CO)8(SnCl2)3(SnCl3)2-(Cl2SnOCOSnCl2)]·2.5CH3COCH3 (28)2.088(9)
2.122(9)
2.03(1)1.28(1)
1.31(2)
116(1)115.5(7)
117.8(8)
KAKLOA
1439435
[81]
Table 8. Selection of spectroscopic data (NMR and IR) assigned to metallocarboxylato tin complexes.
Table 8. Selection of spectroscopic data (NMR and IR) assigned to metallocarboxylato tin complexes.
Compounds119Sn{1H} NMR
(δ, ppm)
13C{1H} NMR
M–CO2
(δ, ppm)
IR
ν(OCO)
(cm−1)
Ref.
(η-C5H5)Re(NO)(PPh3)(CO2SnPh3) (16)−167.1 a
J119Sn,31P = 4.6 Hz
207.6 a1395, 1188[72]
Cp*Re(CO)(NO)(CO2SnPh3) (19)na206.94 b1429, 1174[76]
CpFe(CO)(PPh3)(CO2SnMe3) (20)na219.63 a
d, JP,C = 30.2 Hz
220.42 c
d, JP,C = 32.5 Hz
1433, 1132[77]
CpFe(CO)(PPh3)(CO2Sn(n-Bu)3) (21)na219.56 a
d, JP,C = 30.9 Hz
220.29 c
d, JP,C = 31.4 Hz
1480, 1132[77]
Cp*Fe(CO)2(CO2SnPh3) (22)na215.10 a1470, 1146[77]
Cp*Re(CO)(NO)(CO2SnMe3) (23)na197.58 a
196.10 c
1512, 1162[77]
Cp*Re(CO)(NO)(CO2Sn(Cl)Me2) (24)na205.66 d1385, 1246[78]
[Cp*Re(CO)(NO)(CO2)]2SnMe2 (25)na202.14 d1469, 1186[78]
FeCl(CO2SnPh3)(depe)2 (26)na243.8 d
qui, JP,C = 22 Hz
1306[80]
a Measured in chloroform-d; b methylenchloride-d2; c measured in tetrahydrofuran-d8; d measured in benzene-d6.

6. Conclusions

In the conclusion of the first part of this inventory, “CO2 Derivatives of Molecular Tin Compounds. Part 1: Hemicarbonato and Carbonato Complexes”, we previously claimed that the CO2 derivatives of molecular tin compounds could truly be considered as a class of compounds in their own right. The additional solid-state structures described in this second part and focusing more specifically on carbamato, formato, phosphinoformato and metallocarboxylato complexes, contribute assuredly to reinforce this point of view. Once again, a rich diversity of architectures, including discrete, dimeric, polynuclear and polymeric structures, and exhibiting various modes of coordination, sometimes totally unusual, were highlighted. Thus, this collection of compounds supports again the great reactivity of molecular tin complexes toward carbon dioxide and underlines the facile insertion of CO2 into Sn–X bonds (X = OR, N, H, P). This reactivity, which has been known for a long time, has already been efficiently used for catalytic applications involving CO2 and organotins, in particular as has been shown, with the aim of accessing carbamate compounds and their derivatives. However, the latest advances recently published are very interesting and promising, suggesting new perspectives in terms of reactivity of organotins with respect to small molecules and in particular for the activation of carbon dioxide. This is especially the case of studies relating to the reductive hydroboration of CO2 or those involving the formation of CO2-Sn/phosphorus frustrated Lewis pairs (FLP) adducts, which start new insights. In the field of inorganic chemistry, the isolation of a new cluster incorporating a bis-stannyl-carboxylate group, also viewed as the coordination of a carbonite fragment, constitutes an innovative result which can be considered as unprecedented. Thus, the second half of the last century was incontestably a period of strong developments and fundamental advances in organotin chemistry, but the coming decades promise to be just as fruitful and exciting in terms of progress and inventiveness, and in particular, in the quest to activate carbon dioxide.

Funding

The author thanks the Centre National de la Recherche Scientifique (C.N.R.S-France) and the University of Bourgogne Franche-Comté for financial support.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

A special thanks to three anonymous reviewers for their judicious corrections and suggestions which have contributed to improving the manuscript.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Plasseraud, L. CO2 derivatives of molecular tin compounds. Part 1: Hemicarbonato and carbonato complexes. Inorganics 2020, 8, 31. [Google Scholar] [CrossRef]
  2. Aresta, M.; Nobile, C.F.; Albano, V.G.; Forni, E.; Manassero, M. New nickel-carbon dioxide complex: Synthesis, properties, and crystallographic characterization of (carbon dioxide)bis(tricyclohexylphosphine)nickel. J. Chem. Soc. Chem. Commun. 1975, 636–637. [Google Scholar] [CrossRef]
  3. Álvarez, M.; Galindo, A.; Pérez, P.J.; Carmona, E. Molybdenum and tungsten complexes with carbon dioxide and ethylene ligands. Chem. Sci. 2019, 10, 8541–8546. [Google Scholar] [CrossRef] [Green Version]
  4. Gibson, D.H. The organometallic chemistry of carbon dioxide. Chem. Rev. 1996, 96, 2063–2096. [Google Scholar] [CrossRef] [PubMed]
  5. Mascetti, J. Carbon dioxide coordination chemistry and reactivity of coordinated CO2. In Carbon Dioxide as Chemical Feedstock; Aresta, M., Ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2010; pp. 55–88. [Google Scholar]
  6. Mascetti, J. Metal coordination of CO2. In Encyclopedia of Inorganic and Bioinorganic Chemistry; Wiley: Hoboken, NJ, USA, 2014; pp. 1–17. [Google Scholar]
  7. Aresta, M.; Dibenedetto, A.; Quaranta, E. CO2 coordination to metal centres: Modes of bonding and reactivity. In Reaction Mechanisms in Carbon Dioxide Conversion; Springer: Berlin/Heidelberg, Germany, 2016; pp. 35–69. [Google Scholar]
  8. Paparo, A.; Okuda, J. Carbon dioxide complexes: Bonding modes and synthetic methods. Coord. Chem. Rev. 2017, 334, 136–149. [Google Scholar] [CrossRef]
  9. Holscher, M.; Gurtler, C.; Keim, W.; Muller, T.E.; Peters, M.; Leitner, W. Carbon dioxide as a carbon resource—Recent trends and perspectives. Z. Naturforschung B 2012, 67b, 961–975. [Google Scholar] [CrossRef]
  10. Dibenedetto, A.; Angelini, A.; Stufano, P. Use of carbon dioxide as feedstock for chemicals and fuels: Homogeneous and heterogeneous catalysis. J. Chem. Tech. Biotech. 2014, 89, 334–353. [Google Scholar] [CrossRef]
  11. Wang, H.; Gao, P.; Zhao, T.; Wei, W.; Sun, Y. Recent advances in the catalytic conversion of CO2 to value added compounds. Sci. China Chem. 2015, 58, 79–92. [Google Scholar] [CrossRef]
  12. Liu, Q.; Wu, L.; Jackstell, R.; Beller, M. Using carbon dioxide as a building block in organic synthesis. Nat. Commun. 2015, 6, 5933. [Google Scholar] [CrossRef] [PubMed]
  13. Kleij, A.W.; North, M.; Urakawa, A. CO2 catalysis. ChemSusChem 2017, 10, 1036–1038. [Google Scholar] [CrossRef] [Green Version]
  14. Dabral, S.; Schaub, T. The use of carbon dioxide (CO2) as a building block in organic synthesis from an industrial perspective. Adv. Synth. Catal. 2019, 361, 223–246. [Google Scholar] [CrossRef] [Green Version]
  15. Jana, A.; Tavčar, G.; Roesky, H.W.; John, M. Germanium(II) hydride mediated reduction of carbon dioxide to formic acid and methanol with ammonia borane as the hydrogen source. Dalton Trans. 2010, 39, 9487–9489. [Google Scholar] [CrossRef]
  16. Villegas-Escobar, N.; Schaefer, H.F., III; Toro-Labbé, A. Formation of formic acid derivatives through activation and hydroboration of CO2 by low-valent group 14 (Si, Ge, Sn, Pb) catalysts. J. Phys. Chem. A 2020, 124, 1121–1133. [Google Scholar] [CrossRef]
  17. Groom, C.R.; Bruno, I.J.; Lightfoot, S.C.; Ward, S.C. The Cambridge Structural Database. Acta Cryst. 2016, B72, 171–179. [Google Scholar] [CrossRef] [PubMed]
  18. Wang, H.; Xin, Z.; Li, Y. Synthesis of ureas from CO2. Top. Curr. Chem. 2017, 375, 1–26. [Google Scholar]
  19. Yang, Z.-Z.; He, L.-N.; Gao, J.; Liu, A.-H.; Yu, B. Carbon dioxide utilization with C–N bond formation: Carbon dioxide capture and subsequent conversion. Energy Environ. Sci. 2012, 5, 6602–6639. [Google Scholar] [CrossRef]
  20. Li, J.-Y.; Song, Q.-W.; Zhang, K.; Liu, P. Catalytic conversion of carbon dioxide through C–N bond formation. Molecules 2019, 24, 182. [Google Scholar] [CrossRef] [Green Version]
  21. Dalpozzo, R.; Della Ca’, N.; Gabriele, B.; Mancuso, R. Recent advances in the chemical fixation of carbon dioxide: A green route to carbonylated heterocycle synthesis. Catalysts 2019, 9, 511. [Google Scholar] [CrossRef] [Green Version]
  22. Tominaga, K.-I.; Sasaki, Y. Synthesis of 2-oxazolidinones from CO2 and 1,2-aminoalcohols catalyzed by n-Bu2SnO. Synlett 2002, 2, 307–309. [Google Scholar] [CrossRef]
  23. Pulla, S.; Felton, C.M.; Gartia, Y.; Ramidi, P.; Ghosh, A. Synthesis of 2-oxazolidinones by direct condensation of 2-aminoalcohols with carbon dioxide using chlorostannoxanes. ACS Sustain. Chem. Eng. 2013, 1, 309–312. [Google Scholar] [CrossRef]
  24. Chaturvedi, D. Perspectives on the synthesis of organic carbamates. Tetrahedron 2012, 68, 15–45. [Google Scholar] [CrossRef]
  25. Dell’Amico, D.B.; Calderazzo, F.; Labella, L.; Marchetti, F.; Pampaloni, G. Converting carbon dioxide into Carbamato Derivatives. Chem. Rev. 2003, 103, 3857–3898. [Google Scholar] [CrossRef]
  26. Bresciani, G.; Biancalana, L.; Pampaloni, G.; Marchetti, F. Recent advances in the chemistry of metal carbamates. Molecules 2020, 25, 3603. [Google Scholar] [CrossRef]
  27. Zhu, G.; Li, H.; Cao, Y.; Liu, H.; Li, X.; Chen, J.; Tang, Q. Kinetic study on the novel efficient clean decomposition of methyl N-phenyl carbamate to phenyl isocyanate. Ind. Eng. Chem. Res. 2013, 52, 4450–4454. [Google Scholar] [CrossRef]
  28. Heyn, R.H.; Jacobs, I.; Carr, R.H. Chapter three—Synthesis of aromatic carbamates from CO2: Implications for the polyurethane industry. Adv. Inorg. Chem. 2014, 66, 83–115. [Google Scholar]
  29. Abla, M.; Choi, J.-C.; Sakakura, T. Halogen-free process for the conversion of carbon dioxide to urethanes by homogeneous catalysis. Chem. Commun. 2001, 2238–2239. [Google Scholar] [CrossRef] [PubMed]
  30. Yuan, H.-Y.; Choi, J.-C.; Onozawa, S.-y.; Fukaya, N.; Choi, S.J.; Yasuda, H.; Sakakura, T. The direct synthesis of N-phenylcarbamates from CO2, aniline, and dibutyltin dialkoxide. J. CO2 Util. 2016, 16, 282–286. [Google Scholar] [CrossRef]
  31. Ballivet-Tkatchenko, D.; Chermette, H.; Plasseraud, L.; Walter, O. Insertion reaction of carbon dioxide into Sn–OR bond. Synthesis, structure and DFT calculations of di- and tetranuclear isopropylcarbonato tin(IV) complexes. Dalton Trans. 2006, 5167–5175. [Google Scholar] [CrossRef] [PubMed]
  32. Germain, N.; Müller, I.; Hanauer, M.; Paciello, R.A.; Baumann, R.; Trapp, O.; Schaub, T. Synthesis of industrially relevant carbamates towards isocyanates using carbon dioxide and organotin(IV) alkoxides. ChemSusChem 2016, 9, 1586–1590. [Google Scholar] [CrossRef] [PubMed]
  33. Germain, N.; Hermsen, M.; Schaub, T.; Trapp, O. Synthesis of carbamates from carbon dioxide promoted by organostannanes and alkoxysilanes. Appl. Organomet. Chem. 2017, 31, e3733. [Google Scholar] [CrossRef]
  34. Ballivet-Tkatchenko, D.; Jerphagnon, T.; Ligabue, R.; Plasseraud, L.; Poinsot, D. The role of distannoxanes in the synthesis of dimethyl carbonate from carbon dioxide. Appl. Catal. Gen. 2003, 255, 93–99. [Google Scholar] [CrossRef]
  35. Bloodworth, A.J.; Davies, A.G. Organometallic reactions. Part I. The addition of tin alkoxides to isocyanates. J. Chem. Soc. 1965, 5238–5244. [Google Scholar] [CrossRef]
  36. Bloodworth, A.J.; Davies, A.G. Organometallic reactions. Part II. The addition of bistrialkyltin oxides to isocyanates. J. Chem. Soc. 1965, 6245–6249. [Google Scholar] [CrossRef]
  37. Harrison, P.G.; Zuckerman, J.J. Structural studies in main group chemistry III. Tin-119m Mössbauer investigation of some tin-nitrogen bonded compounds. J. Organomet. Chem. 1973, 55, 261–266. [Google Scholar] [CrossRef]
  38. Devendra, R.; Edmonds, N.R.; Söhnel, T. Interaction between trialkyltin alkoxide and phenyl isocyanate in the formation of tin carbamate: A computational and experimental study. J. Mol. Catal. A 2014, 395, 72–86. [Google Scholar] [CrossRef]
  39. Bloodworth, A.J.; Davies, A.G.; Vasishtha, S.C. Organometallic reactions. Part XIV. The decarboxylation of trialkyltin carbamates by isocyanates and isothiocyanates: A new route to carbodi-imides. J. Chem. Soc. C 1968, 2640–2646. [Google Scholar] [CrossRef]
  40. Shibata, I.; Yamasaki, H.; Baba, A.; Matsuda, H. A novel Darzens-type reaction promoted by tributylstannylcarbamate. J. Org. Chem. 1992, 57, 6909–6914. [Google Scholar] [CrossRef]
  41. Shibata, I.; Mori, Y.; Yamasaki, H.; Baba, A.; Matsuda, H. Novel Michael addition promoted by tributylstannylcarbamate. Synthesis of diacylcyclopropanes. Tetrahedron Lett. 1993, 34, 6567–6570. [Google Scholar] [CrossRef]
  42. Zakharov, N.; Struchkov, Y.T.; Ganina, V.I.; Kuz’min, E.A. X-ray crystallographic investigation of nonbinding interactions and coordination in organometallic compounds. XV. The crystal structure of the trimeththylstannyl ester of trimethylstannylcarbamic acid Me3SnNHC(=O)–OSnMe3. Zhurnal Strukturnoi Khimii 1980, 21, 162–167. [Google Scholar]
  43. Plasseraud, L. Organotin(IV) complexes containing Sn−O−Se moieties: A structural inventory. Synthesis 2018, 50, 3653–3661. [Google Scholar] [CrossRef]
  44. Kümmerlen, J.; Sebald, A.; Reuter, H. The structure of (iBu3Sn)2CO3 and (Me3Sn)2CO3 in solution and in the solid state studied by 13C/119Sn NMR spectroscopy and X-ray diffraction. J. Organomet. Chem. 1992, 427, 309–323. [Google Scholar] [CrossRef]
  45. Abis, L.; Belli Dell’Amico, D.; Calderazzo, F.; Caminiti, R.; Garbassi, F.; Ianelli, S.; Pelizzi, G.; Robino, P.; Tomei, A. N,N-Dialkylcarbamato complexes as precursors for the chemical implantation of metal cations on a silica support. Part I. Tin. J. Mol. Catal. Chem. 1996, 108, L113–L117. [Google Scholar] [CrossRef]
  46. Horley, G.A.; Mahon, M.F.; Molloy, K.C. Synthesis and characterization of novel homoleptic N,N-dialkylcarbamato complexes of antimony: Precursors for the deposition of antimony oxides. Inorg. Chem. 2002, 41, 5052–5058. [Google Scholar] [CrossRef] [PubMed]
  47. Komarov, N.V.; Ryzhkova, N.A.; Andreev, A.A. Synthesis of alkoxystannanes by reactions of O-(organylstannyl) carbamates with alcohols. Russ. Chem. Bull. 2004, 53, 936–938. [Google Scholar] [CrossRef]
  48. Levashov, A.S.; Andreev, A.A.; Buryi, D.S.; Konshin, V.V. A reaction of tin tetra(N,N-diethylcarbamate) with phenylacetylene as a new route to tetra(phenylethynyl)tin. Russ. Chem. Bull. 2015, 63, 775–776. [Google Scholar] [CrossRef]
  49. Buryi, D.S.; Levashov, A.S. The reaction of tin tetracarbamates with organyl chlorosilanes: A novel synthetic route towards O-silylurethanes. Russ. J. Gen. Chem. 2019, 89, 924–928. [Google Scholar] [CrossRef]
  50. Stewart, C.A.; Dickie, D.A.; Parkes, M.V.; Saria, J.A.; Kemp, R.A. Reactivity of bis(2,2,5,5-tetramethyl-2,5-disila-1-azacyclopent-1-yl)tin with CO2, OCS, and CS2 and comparison to that of bis[bis(trimethylsilyl)amido]tin. Inorg. Chem. 2010, 49, 11133–11141. [Google Scholar] [CrossRef]
  51. Sita, L.R.; Babcock, J.R.; Xi, R. Facile metathetical exchange between carbon dioxide and the divalent group 14 bisamides M[N(SiMe3)2]2 (M = Ge and Sn). J. Am. Chem. Soc. 1996, 118, 10912–10913. [Google Scholar] [CrossRef]
  52. Stewart, C.A.; Dickie, D.A.; Tang, Y.; Kemp, R.A. Insertion reactions of CO2, OCS, and CS2 into the Sn–N bonds of (Me2N)2Sn: NMR and X-ray structural characterization of the products. Inorg. Chim. Acta 2011, 376, 73–79. [Google Scholar] [CrossRef]
  53. Harris, L.A.-M.; Coles, M.P.; Fulton, J.R. Synthesis and reactivity of tin amide complexes. Inorg. Chim. Acta 2011, 369, 97–102. [Google Scholar] [CrossRef]
  54. Elöd, E.; Kolbach, F. Tin salts of organic acids. Z. Anorganische Allgemeine Chem. 1927, 164, 297–312. [Google Scholar] [CrossRef]
  55. Donaldson, J.D.; Knifton, J.F. Complex tin(II) formates. J. Chem. Soc. 1964, 6107–6113. [Google Scholar] [CrossRef]
  56. Jelen, A.; Lindqvist, O. The crystal structure of potassium triformatostannate(II), KSn(HCOO)3. Acta Chem. Scand. 1969, 23, 3071–3080. [Google Scholar] [CrossRef]
  57. Harrison, P.G.; Thornton, E.W. Derivatives of bivalent germanium, tin, and lead. Part 21. Tin(II) formate: A reinvestigation. J. C. S. Dalton 1978, 1274–1278. [Google Scholar] [CrossRef]
  58. Molloy, K.C.; Quill, K.; Nowell, I.W. Organotin biocides. Part 8. The crystal structure of triphenyltin formate and a comparative variable-temperature tin-119 Mössbauer spectroscopic study of organotin formates and acetates. J. Chem. Soc. Dalton Trans. 1987, 101–106. [Google Scholar] [CrossRef]
  59. Mistry, F.; Rettig, S.J.; Trotter, J.; Aubke, F. Structure of bis(formato)dimethyltin(IV). Acta Cryst. 1990, C46, 2091–2093. [Google Scholar] [CrossRef]
  60. Ellis, B.D.; Atkins, T.M.; Peng, Y.; Sutton, A.D.; Gordon, J.C.; Power, P.P. Synthesis and thermolytic behavior of tin(IV) formates: In search of recyclable metal–hydride systems. Dalton Trans. 2010, 39, 10659–10663. [Google Scholar] [CrossRef] [PubMed]
  61. McMurtie, J.C.; Arnorld, D.P. Tin(IV) porphyrin complexes. Crystal structures of meso-tetraphenyl-porphyrinatotin(IV) diacetate, bis(dichloro-acetate), bis(trifluoroacetate) and diformate, and structural correlations for tin(IV) porphyrin complexes with O-bound anionic ligands. J. Struct. Chem. 2010, 51, 107–113. [Google Scholar] [CrossRef]
  62. Jana, A.; Roesky, H.W.; Schulzke, C.; Döring, A. Reactions of tin(II) hydride species with unsaturated molecules. Angew. Chem. Int. Ed. 2009, 48, 1106–1109. [Google Scholar] [CrossRef] [PubMed]
  63. Albertin, G.; Antoniutti, S.; Castro, J.; García-Fontán, S.; Zanardo, G. Preparation and reactivity of stannyl complexes of manganese and rhenium. Organometallics 2007, 26, 2918–2930. [Google Scholar] [CrossRef]
  64. Albertin, G.; Antoniutti, S.; Bacchi, A.; Pelizzi, G.; Zanardo, G. Synthesis and reactivity of trihydridostannyl complexes of ruthenium and osmium. Organometallics 2008, 27, 4407–4418. [Google Scholar] [CrossRef]
  65. Albertin, G.; Antoniutti, S.; Castro, J. Reactivity of dihydrides MH2P4 (M = Fe, Ru, Os) with SnCl2: Preparation of bis(trihydridestannyl) derivatives. Organometallics 2010, 29, 3808–3816. [Google Scholar] [CrossRef]
  66. Otte, F.; Koller, S.G.; Golz, C.; Strohmann, C. Crystal structures of diaquadi-μ-hydroxido-tris[trimethyltin(IV)] diformatotrimethylstannate(IV) and di-μ-hydroxido-tris[trimethyltin(IV)] chloride monohydrate. Acta Cryst. 2016, E72, 1499–1502. [Google Scholar] [CrossRef]
  67. Hadlington, T.J.; Kefalidis, C.E.; Maron, L.; Jones, C. Efficient reduction of carbon dioxide to methanol equivalents catalyzed by two-coordinate amido-germanium(II) and -tin(II) hydride complexes. ACS Catal. 2017, 7, 1853–1859. [Google Scholar] [CrossRef]
  68. Zhu, C.; Frigge, R.; Turner, A.M.; Kaiser, R.I.; Sun, B.-J.; Chen, S.-Y.; Chang, A.H.H. First identification of unstable phosphino formic acid (H2PCOOH). Chem. Commun. 2018, 54, 5716–5719. [Google Scholar] [CrossRef]
  69. Dickie, D.A.; Coker, E.N.; Kemp, R.A. Formation of a reversible, intramolecular main-group metal–CO2 adduct. Inorg. Chem. 2011, 50, 11288–11290. [Google Scholar] [CrossRef]
  70. Holtkamp, P.; Friedrich, F.; Stratmann, E.; Mix, A.; Neumann, B.; Stammler, H.-G.; Mitzel, N.W. A neutral germinal tin/phosphorus frustrated Lewis pair. Angew. Chem. Int. Ed. 2019, 58, 5114–5118. [Google Scholar] [CrossRef] [PubMed]
  71. Pinkes, J.R.; Cutler, A.R. Iron-tin carbon dioxide complexes (η5-C5Me5)(CO)2Fe–CO2SnR3 (R = Me, Ph): Observations pertaining to unsymmetrical metallocarboxylates and carboxylate–carbonyl 13C-label exchange. Inorg. Chem. 1994, 33, 759–764. [Google Scholar] [CrossRef]
  72. Senn, D.R.; Gladysz, J.A.; Emerson, K.; Larsen, R.D. Synthesis, structure, and reactivity of transition-metal/main-group-metal bridging carboxylate complexes of the formula (η5-C5H5)Re(NO)(PPh3) (CO2MLn) (M = Li, K, Ge, Sn, Pb). Inorg. Chem. 1987, 26, 2737–2739. [Google Scholar] [CrossRef]
  73. Gibson, D.H.; Richardson, J.F.; Ong, T.-S. 1-Carbonyl-1-η5-cyclopentadienyl-2,2,2-triphenyl-1-triphenylphosphine-μ-carboxylato-1κC:2κO:2κO’-irontin. Acta Cryst. 1991, C47, 259–261. [Google Scholar] [CrossRef]
  74. Holmes, R.R.; Day, R.O.; Chandrasekhar, V.; Vollano, J.F.; Holmes, J.M. Pentacoordinated molecules. 65. Discrete, dimeric, and polymeric structures of triphenyltin esters of chlorobenzoic acids. Inorg. Chem. 1986, 25, 2490–2494. [Google Scholar] [CrossRef]
  75. Gibson, D.H.; Richardson, J.F.; Mbadike, O.P. 1-Carbonyl-μ-carboxylato-1κC:2κO:2κO’-1-η5-indenyl-2,2,2-triphenyl-l-(triphenylphosphine)irontin. Acta Cryst. 1993, B49, 784–786. [Google Scholar] [CrossRef] [Green Version]
  76. Gibson, D.H.; Mehta, J.M.; Ye, M.; Richardson, J.F.; Mashuta, M.S. Synthesis and characterization of rhenium metallocarboxylates. Organometallics 1994, 13, 1070–1072. [Google Scholar] [CrossRef]
  77. Gibson, D.H.; Ye, M.; Sleadd, B.A.; Mehta, J.M.; Mbadike, O.P.; Richardson, J.F.; Mashuta, M.S. Characterization and thermolysis reactions of CO2-bridged iron-tin and rhenium-tin complexes. Structure-reactivity correlations. Organometallics 1995, 14, 1242–1255. [Google Scholar] [CrossRef]
  78. Gibson, D.H.; Mehta, J.M.; Mashuta, M.S.; Richardson, J.F. Synthesis and characterization of new rhenium—Tin complexes with μ23-bridging CO2 ligands. Organometallics 1997, 16, 4828–4832. [Google Scholar] [CrossRef]
  79. Hirano, M.; Akita, M.; Tani, K.; Kumagai, K.; Kasuga, N.C.; Fukuoka, A.; Komiya, S. Activation of coordinated carbon dioxide in Fe(CO2)(depe)2 by group 14 Electrophiles. Organometallics 1997, 16, 4206–4213. [Google Scholar] [CrossRef]
  80. Adams, R.D.; Chen, M.; Trufan, E. Complexes containing bridging tin- and germanium-substituted metallocarboxylate ligands from the reactions of Ph3SnOH and Ph3GeOH with Os3(CO)12 in the presence of base. J. Organomet. Chem. 2011, 696, 2894–2898. [Google Scholar] [CrossRef]
  81. Bortoluzzi, M.; Ceriotti, A.; Ciabatti, I.; Della Pergola, R.; Femoni, C.; Iapalucci, M.C.; Storione, A.; Zacchini, S. Platinum carbonyl clusters stabilized by Sn(II)-based fragments: Syntheses and structures of [Pt6(CO)6(SnCl2)2(SnCl3)4]4−, [Pt9(CO)8(SnCl2)3(SnCl3)2(Cl2SnOCOSnCl2)]4− and [Pt10(CO)14{Cl2Sn(OH)SnCl2}2]2−. Dalton Trans. 2016, 45, 5001–5013. [Google Scholar] [CrossRef] [PubMed]
  82. Paparo, A.; Okuda, J. Carbonite, the dianion of carbon dioxide and its metal complexes. J. Organomet. Chem. 2018, 869, 270–274. [Google Scholar] [CrossRef]
Scheme 1. Molecular representations of nitrogen-containing molecules accessible from CO2 used as a C1 building block.
Scheme 1. Molecular representations of nitrogen-containing molecules accessible from CO2 used as a C1 building block.
Inorganics 09 00018 sch001
Scheme 2. Synthesis of 2-oxazolidinones from CO2 and 1,2-aminoalcohols catalyzed by n-Bu2SnO, data from [22].
Scheme 2. Synthesis of 2-oxazolidinones from CO2 and 1,2-aminoalcohols catalyzed by n-Bu2SnO, data from [22].
Inorganics 09 00018 sch002
Scheme 3. Key step requiring the insertion of CO2 into the Sn–N bond of the stannaoxazolidine species formed in situ, data from [22].
Scheme 3. Key step requiring the insertion of CO2 into the Sn–N bond of the stannaoxazolidine species formed in situ, data from [22].
Inorganics 09 00018 sch003
Scheme 4. Molecular representation of chlorostannoxane compounds used as tin precursors for the synthesis of 2-oxazolidinones by direct condensation of 2-aminoalcohols with carbon dioxide (R = n-Bu, Ph; R′ = n-Bu, Ph), data from [23].
Scheme 4. Molecular representation of chlorostannoxane compounds used as tin precursors for the synthesis of 2-oxazolidinones by direct condensation of 2-aminoalcohols with carbon dioxide (R = n-Bu, Ph; R′ = n-Bu, Ph), data from [23].
Inorganics 09 00018 sch004
Scheme 5. Halogen-free process for the conversion of carbon dioxide to carbamates, data from [29].
Scheme 5. Halogen-free process for the conversion of carbon dioxide to carbamates, data from [29].
Inorganics 09 00018 sch005
Scheme 6. Tin-promoted synthesis of N-phenyl carbamate from aniline and carbon dioxide (R = Me, Et, n-Pr, and n-Bu), data from [30].
Scheme 6. Tin-promoted synthesis of N-phenyl carbamate from aniline and carbon dioxide (R = Me, Et, n-Pr, and n-Bu), data from [30].
Inorganics 09 00018 sch006
Scheme 7. Formation of organo ester of N-organo-N-trialkylstannylcarbamic acid from trialkyltin alkoxides and isocyanates, data from [35].
Scheme 7. Formation of organo ester of N-organo-N-trialkylstannylcarbamic acid from trialkyltin alkoxides and isocyanates, data from [35].
Inorganics 09 00018 sch007
Scheme 8. Formation of trialkyltin ester of N-organo-N-trialkylstannylcarbamic acid from bis(trialkyltin) oxides and isocyanates, data from [36].
Scheme 8. Formation of trialkyltin ester of N-organo-N-trialkylstannylcarbamic acid from bis(trialkyltin) oxides and isocyanates, data from [36].
Inorganics 09 00018 sch008
Scheme 9. Darzens-type reaction (top) and Michael addition (bottom) promoted by methyl ester of N-ethyl-N-tributylstannylcarbamic acid, data from [40,41].
Scheme 9. Darzens-type reaction (top) and Michael addition (bottom) promoted by methyl ester of N-ethyl-N-tributylstannylcarbamic acid, data from [40,41].
Inorganics 09 00018 sch009
Figure 1. Molecular structure of 1, adapted from [42] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, N dark blue, C grey).
Figure 1. Molecular structure of 1, adapted from [42] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, N dark blue, C grey).
Inorganics 09 00018 g001
Scheme 10. Reaction scheme leading to Compounds 2 and 3, data from [45].
Scheme 10. Reaction scheme leading to Compounds 2 and 3, data from [45].
Inorganics 09 00018 sch010
Scheme 11. Alternative route leading to compound 3, data from [46].
Scheme 11. Alternative route leading to compound 3, data from [46].
Inorganics 09 00018 sch011
Figure 2. Molecular structure of 3, adapted from [46] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, N dark blue, C grey).
Figure 2. Molecular structure of 3, adapted from [46] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, N dark blue, C grey).
Inorganics 09 00018 g002
Scheme 12. Reaction of tetrakis(N,N-diethylcarbamato)tin(IV) with phenylacetylene leading to tetra(phenylethynyl)tin, data from [47].
Scheme 12. Reaction of tetrakis(N,N-diethylcarbamato)tin(IV) with phenylacetylene leading to tetra(phenylethynyl)tin, data from [47].
Inorganics 09 00018 sch012
Scheme 13. Reaction scheme leading to Compound 4, data from [50].
Scheme 13. Reaction scheme leading to Compound 4, data from [50].
Inorganics 09 00018 sch013
Figure 3. Molecular structure of 4, adapted from [50] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, N dark blue, C grey, Si light yellow).
Figure 3. Molecular structure of 4, adapted from [50] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, N dark blue, C grey, Si light yellow).
Inorganics 09 00018 g003
Scheme 14. Reaction scheme leading to Compound 5, data from [52].
Scheme 14. Reaction scheme leading to Compound 5, data from [52].
Inorganics 09 00018 sch014
Figure 4. Molecular structure of 5, adapted from [52] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, N dark blue, C grey). Intermolecular hydrogen bonds are shown by cyan dash lines.
Figure 4. Molecular structure of 5, adapted from [52] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, N dark blue, C grey). Intermolecular hydrogen bonds are shown by cyan dash lines.
Inorganics 09 00018 g004
Scheme 15. Reaction scheme leading to Compounds 6 and 7, data from [53].
Scheme 15. Reaction scheme leading to Compounds 6 and 7, data from [53].
Inorganics 09 00018 sch015
Figure 5. Molecular structure of 6, adapted from [53] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, N dark blue, C grey).
Figure 5. Molecular structure of 6, adapted from [53] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, N dark blue, C grey).
Inorganics 09 00018 g005
Figure 6. Molecular structure of [Ph3Sn(O2CH)]n (CSD refcode: FAJZIZ, 1151980), adapted from [58] (MERCURY view). Hydrogen atoms are omitted for clarity except that of the formate group (Sn light blue, O red, C grey, H white).
Figure 6. Molecular structure of [Ph3Sn(O2CH)]n (CSD refcode: FAJZIZ, 1151980), adapted from [58] (MERCURY view). Hydrogen atoms are omitted for clarity except that of the formate group (Sn light blue, O red, C grey, H white).
Inorganics 09 00018 g006
Figure 7. Infinite sheets based-structure of [Me2Sn(O2CH)2]n (CCDC ref.: SIFLEY, 1258788), adapted from [59] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, C grey).
Figure 7. Infinite sheets based-structure of [Me2Sn(O2CH)2]n (CCDC ref.: SIFLEY, 1258788), adapted from [59] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, C grey).
Inorganics 09 00018 g007
Figure 8. Polymeric structure of [n-Bu2Sn(O2CH)2]n (CCDC ref.: EKONEY, 783726), adapted from [60] (MERCURY view). Hydrogen atoms are omitted for clarity and for each n-butyl chain; only the α-carbon atoms bonded to tin are shown (Sn light blue, O red, C grey).
Figure 8. Polymeric structure of [n-Bu2Sn(O2CH)2]n (CCDC ref.: EKONEY, 783726), adapted from [60] (MERCURY view). Hydrogen atoms are omitted for clarity and for each n-butyl chain; only the α-carbon atoms bonded to tin are shown (Sn light blue, O red, C grey).
Inorganics 09 00018 g008
Figure 9. Molecular structure of Dmp3SnOC(O)H (dmp = 2,6-dimethylphenyl) (CCDC ref.: EKONIC, 783727), adapted from [60] (MERCURY view). Hydrogen atoms are omitted for clarity except that of the formate group (Sn light blue, O red, C grey, H white).
Figure 9. Molecular structure of Dmp3SnOC(O)H (dmp = 2,6-dimethylphenyl) (CCDC ref.: EKONIC, 783727), adapted from [60] (MERCURY view). Hydrogen atoms are omitted for clarity except that of the formate group (Sn light blue, O red, C grey, H white).
Inorganics 09 00018 g009
Figure 10. Molecular structure of Sn(TPP)[(OC(O)H]2 (TPP = dianion of 5,10,15,20-tetraphenylporphyrin) (CCDC ref.: EPAZAX, 711428), adapted from [61] (MERCURY view). Hydrogen atoms are omitted for clarity except that of the formate group (Sn light blue, O red, N dark blue, C grey, H white).
Figure 10. Molecular structure of Sn(TPP)[(OC(O)H]2 (TPP = dianion of 5,10,15,20-tetraphenylporphyrin) (CCDC ref.: EPAZAX, 711428), adapted from [61] (MERCURY view). Hydrogen atoms are omitted for clarity except that of the formate group (Sn light blue, O red, N dark blue, C grey, H white).
Inorganics 09 00018 g010
Scheme 16. Reaction scheme leading to Compound 8, data from [62].
Scheme 16. Reaction scheme leading to Compound 8, data from [62].
Inorganics 09 00018 sch016
Figure 11. Molecular structure of 8, adapted from [62] (MERCURY view). Hydrogen atoms are omitted for clarity except that of the formate group (Sn light blue, O red, N dark blue, C grey, H white).
Figure 11. Molecular structure of 8, adapted from [62] (MERCURY view). Hydrogen atoms are omitted for clarity except that of the formate group (Sn light blue, O red, N dark blue, C grey, H white).
Inorganics 09 00018 g011
Scheme 17. Reaction scheme leading to compound 9, data from [63].
Scheme 17. Reaction scheme leading to compound 9, data from [63].
Inorganics 09 00018 sch017
Figure 12. Molecular structure of 9, adapted from [63] (MERCURY view). Hydrogen atoms are omitted for clarity except those of hydroxyl and formate groups (Re blue, Sn light blue, O red, P orange, C grey, H white).
Figure 12. Molecular structure of 9, adapted from [63] (MERCURY view). Hydrogen atoms are omitted for clarity except those of hydroxyl and formate groups (Re blue, Sn light blue, O red, P orange, C grey, H white).
Inorganics 09 00018 g012
Scheme 18. Reaction scheme leading to Compound 10, data from [64].
Scheme 18. Reaction scheme leading to Compound 10, data from [64].
Inorganics 09 00018 sch018
Figure 13. Molecular structure of 10, adapted from [64] (MERCURY view). Hydrogen atoms are omitted for clarity except those of hydroxyl and formate groups (Ru blue green, Sn light blue, O red, P orange, C grey, H white).
Figure 13. Molecular structure of 10, adapted from [64] (MERCURY view). Hydrogen atoms are omitted for clarity except those of hydroxyl and formate groups (Ru blue green, Sn light blue, O red, P orange, C grey, H white).
Inorganics 09 00018 g013
Scheme 19. Reaction scheme leading to Compound 11, data from [65].
Scheme 19. Reaction scheme leading to Compound 11, data from [65].
Inorganics 09 00018 sch019
Figure 14. Molecular structure of 11, adapted from [65] (MERCURY view). Ph and OEt substituents are omitted for clarity (Os blue, Sn light blue, O red, P orange, C grey, H white).
Figure 14. Molecular structure of 11, adapted from [65] (MERCURY view). Ph and OEt substituents are omitted for clarity (Os blue, Sn light blue, O red, P orange, C grey, H white).
Inorganics 09 00018 g014
Figure 15. Molecular structure of 12, adapted from [66] (MERCURY view). Hydrogen atoms of methyl groups are omitted for clarity (Sn light blue, O red, C grey, H white).
Figure 15. Molecular structure of 12, adapted from [66] (MERCURY view). Hydrogen atoms of methyl groups are omitted for clarity (Sn light blue, O red, C grey, H white).
Inorganics 09 00018 g015
Scheme 20. Reaction scheme leading to Compound 13, data from [67].
Scheme 20. Reaction scheme leading to Compound 13, data from [67].
Inorganics 09 00018 sch020
Figure 16. Molecular structure of 13, adapted from [67] (MERCURY view). Hydrogen atoms are omitted for clarity except that of the formate group (Sn light blue, O red, N dark blue, Si orange, C grey, H white).
Figure 16. Molecular structure of 13, adapted from [67] (MERCURY view). Hydrogen atoms are omitted for clarity except that of the formate group (Sn light blue, O red, N dark blue, Si orange, C grey, H white).
Inorganics 09 00018 g016
Scheme 21. Reaction scheme leading to Compound 14, data from [69].
Scheme 21. Reaction scheme leading to Compound 14, data from [69].
Inorganics 09 00018 sch021
Figure 17. Molecular structure of 14, adapted from [69] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, P orange, N dark blue, C grey).
Figure 17. Molecular structure of 14, adapted from [69] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, P orange, N dark blue, C grey).
Inorganics 09 00018 g017
Scheme 22. Reaction scheme leading to Compound 15, data from [70].
Scheme 22. Reaction scheme leading to Compound 15, data from [70].
Inorganics 09 00018 sch022
Figure 18. Molecular structure of 15, adapted from [70] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, P orange, C grey, F green).
Figure 18. Molecular structure of 15, adapted from [70] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, P orange, C grey, F green).
Inorganics 09 00018 g018
Scheme 23. Reaction scheme leading to Compound 16, data from [72].
Scheme 23. Reaction scheme leading to Compound 16, data from [72].
Inorganics 09 00018 sch023
Figure 19. Molecular structure of 16, adapted from [72] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, P orange, N dark blue, Re black, C grey).
Figure 19. Molecular structure of 16, adapted from [72] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, P orange, N dark blue, Re black, C grey).
Inorganics 09 00018 g019
Figure 20. Molecular structure of 17, adapted from [73] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, P orange, Fe white, C grey).
Figure 20. Molecular structure of 17, adapted from [73] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, P orange, Fe white, C grey).
Inorganics 09 00018 g020
Figure 21. Molecular structure of 18, adapted from [75] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, P orange, Fe white, C grey).
Figure 21. Molecular structure of 18, adapted from [75] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, P orange, Fe white, C grey).
Inorganics 09 00018 g021
Figure 22. Molecular structure of 19, adapted from [76] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, P orange, N blue, Re black, C grey).
Figure 22. Molecular structure of 19, adapted from [76] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, P orange, N blue, Re black, C grey).
Inorganics 09 00018 g022
Figure 23. Molecular structure of 23, adapted from [77] (MERCURY view). Hydrogen atoms and one molecule of CHCl3 are omitted for clarity (Sn light blue, O red, N blue, Re black, C grey).
Figure 23. Molecular structure of 23, adapted from [77] (MERCURY view). Hydrogen atoms and one molecule of CHCl3 are omitted for clarity (Sn light blue, O red, N blue, Re black, C grey).
Inorganics 09 00018 g023
Scheme 24. Reaction scheme leading to compound 24, data from [78].
Scheme 24. Reaction scheme leading to compound 24, data from [78].
Inorganics 09 00018 sch024
Scheme 25. Reaction scheme leading to compound 25, data from [78].
Scheme 25. Reaction scheme leading to compound 25, data from [78].
Inorganics 09 00018 sch025
Figure 24. Molecular structure of 25, adapted from [78] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, N dark blue, Re black, C grey).
Figure 24. Molecular structure of 25, adapted from [78] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, N dark blue, Re black, C grey).
Inorganics 09 00018 g024
Scheme 26. Reaction scheme leading to Compound 26, data from [79].
Scheme 26. Reaction scheme leading to Compound 26, data from [79].
Inorganics 09 00018 sch026
Figure 25. Molecular structure of 26, adapted from [79] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, P orange, Cl light green, Fe white, C grey).
Figure 25. Molecular structure of 26, adapted from [79] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, P orange, Cl light green, Fe white, C grey).
Inorganics 09 00018 g025
Figure 26. Molecular structure of 27, adapted from [80] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, Os dark blue, C grey,).
Figure 26. Molecular structure of 27, adapted from [80] (MERCURY view). Hydrogen atoms are omitted for clarity (Sn light blue, O red, Os dark blue, C grey,).
Inorganics 09 00018 g026
Figure 27. Molecular structure of 28, adapted from [81] (MERCURY view). Hydrogen atoms, two and a half molecules of acetone and four tetra-n-butylammonium cations are omitted for clarity (Sn light blue, O red, Cl light green, Pt light grey, C grey).
Figure 27. Molecular structure of 28, adapted from [81] (MERCURY view). Hydrogen atoms, two and a half molecules of acetone and four tetra-n-butylammonium cations are omitted for clarity (Sn light blue, O red, Cl light green, Pt light grey, C grey).
Inorganics 09 00018 g027
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Plasseraud, L. CO2 Derivatives of Molecular Tin Compounds. Part 2: Carbamato, Formato, Phosphinoformato and Metallocarboxylato Complexes. Inorganics 2021, 9, 18. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics9030018

AMA Style

Plasseraud L. CO2 Derivatives of Molecular Tin Compounds. Part 2: Carbamato, Formato, Phosphinoformato and Metallocarboxylato Complexes. Inorganics. 2021; 9(3):18. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics9030018

Chicago/Turabian Style

Plasseraud, Laurent. 2021. "CO2 Derivatives of Molecular Tin Compounds. Part 2: Carbamato, Formato, Phosphinoformato and Metallocarboxylato Complexes" Inorganics 9, no. 3: 18. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics9030018

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop