Next Article in Journal
Characterization of Volatilized Compounds in Conventional and Organic Vegetable-Source Alternative Meat-Curing Ingredients
Next Article in Special Issue
Dipolar Copper(I) Complexes: A Novel Appealing Class of Highly Active Second-Order NLO-Phores
Previous Article in Journal
Antiviral and Immune Enhancement Effect of Platycodon grandiflorus in Viral Diseases: A Potential Broad-Spectrum Antiviral Drug
Previous Article in Special Issue
The Influence of pH on Long-Range Electron Transfer and Proton-Coupled Electron Transfer in Ruthenium-Modified Azurin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Anions of α-Amino Acids as (O,N)-Donor Ligands in Si-, Ge- and Sn-Coordination Chemistry

Institut für Anorganische Chemie, TU Bergakademie Freiberg, D-09596 Freiberg, Germany
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(4), 834; https://doi.org/10.3390/molecules30040834
Submission received: 30 December 2024 / Revised: 2 February 2025 / Accepted: 3 February 2025 / Published: 11 February 2025
(This article belongs to the Special Issue Exclusive Feature Papers in Inorganic Chemistry, 2nd Edition)

Abstract

:
This review gives an overview of Si-, Ge- and Sn-compounds with (O,N)-bi- and -oligodentate ligands, which have the α-amino carboxylic acid motif N–C(R,R′)–C(=O)O in common (R,R′ = H or hydrocarbyl). While the amino acids themselves are encountered as mono- and di-anionic ligands, modifications at the N-terminus (e.g., extension of the ligand backbone by, e.g., additional alkane carboxylic acid groups) give rise to a wealth of ligands, which bear the α-amino carboxylic acid motif. With particular interest in the coordination features of these ligands, crystallographically characterized complexes are the focus of this review.

Graphical Abstract

1. Introduction

Coordination chemistry of the heavier tetrels Si, Ge and Sn has attracted researchers’ interest for decades. Simple complexes of tetravalent tetrels with a higher coordination number were reported more than one century ago, e.g., hexafluorosilicate ([SiF6]2–) [1], a related ammonia adduct ([SiF4(NH3)2]) [2], hexachlorostannate ([SnCl6]2–) [3] and the Ge-analog ([GeCl6]2–) [4]. (The somewhat later dates of reports of various Ge compounds can be attributed to the discovery of the element germanium late, in 1886 [5].) In further studies, chelating ligands have given rise to a wealth of compounds in which these tetrels exhibit coordination numbers five, six or even higher. Thus, coordination chemistry of these tetrels has been thoroughly reviewed over the past decades; for examples, see [6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25]. As these tetrels (when tetravalent) are oxidophilic, which is manifest in the natural abundance of minerals such as cassiterite (SnO2) and quartz (SiO2), chelators with anionic O-anchor groups have proven successful in binding additional lone-pair donor sites (e.g., amine or imine N atoms) to the tetrel atoms, thus enhancing their coordination numbers. Figure 1 shows a selection of tetrel complexes IXI [26,27,28,29,30,31,32,33,34,35,36] with two simple bidentate (O,N)-donor chelating ligands (mono-anions of 8-oxyquinoline and of N,N-dimethylaminoethanol). They reveal that chelators with the N-C-C-O sequence can bind to the tetrels with the formation of five-membered ring motifs and that they can enhance the tetrel coordination number to, e.g., five, six or seven. Moreover, as the pair of complexes IX and X shows, formation of ionic tetrel complexes must be taken into consideration as well, and pair X, XI demonstrates that isomeric coordination modes of the bidentate ligands may be encountered with otherwise related complexes.
In principle, a great variety of (O,N)-chelators is available from the natural chiral pool (i.e., α-amino acids), and further chiral and non-chiral derivatives thereof (of natural or artificial origin, such as aminoacetic acid (glycine) and α-amino isobutyric acid, respectively) are available as well. Interestingly, tetrel coordination chemistry with this class of ligands has been explored only scarcely so far (vide infra), but more complex (oligodentate) derivatives of these ligands also contribute to the portfolio of tetrel complexes investigated in the past decades. In combination with our group’s ongoing investigation of (O,N)-chelate complexes of silicon [37,38,39,40,41,42,43,44], this served as our motivation to review the current state of insights into the Si-, Ge- and Sn-coordination chemistry with α-amino acid-derived chelating ligands. Figure 2a–e give a selected impression as to which types of ligands may feature an α-amino acid motif. (For the sake of simplicity, these molecules are drawn in their carboxylic acid form. We are aware that α-amino acids form zwitterionic species.) The two examples shown in Figure 2f indicate that, starting from α-amino acids, modifications at the N and/or C-terminus may give access to various other (O,N)-chelating ligands. In this regard, this review mentions compounds which feature a genuine α-amino carboxylic acid motif in the ligand backbone and compounds which lack the amine or amide motif at the N-terminus (e.g., Schiff bases) and/or the carboxylate at the C-terminus (e.g., alcoholates, phosphonates). Thus, for the sake of clarity, the former are labeled with Arabic numerals and the latter with Roman numerals. Moreover, labels of generic drawings, which resemble groups of compounds, are italicized to emphasize their variable character. The crystal structures of compounds, which are referred to in this review, can be accessed from the Cambridge Structural Database (CSD). For the readers’ convenience, the access codes (CSD reference codes) are summarized in Appendix A (Table A1, Table A2 and Table A3, for the compounds labeled with Roman numerals, with Arabic numerals and for additional compounds apart from these labeling schemes, respectively).

2. Si-, Ge- and Sn-Complexes of α-Amino Carboxylic Acid-Derived (O,N)-Chelating Ligands

2.1. Si-, Ge- and Sn-Complexes Derived from α-Amino Monocarboxylic Acids

2.1.1. Si-Complexes Derived from α-Amino Monocarboxylic Acids

Literature reports from the past decades suggest that α-amino carboxylic acids can be utilized as mono-anionic (O,N)-chelators at Si, e.g., in complexes generated by reaction of Me3SiCl or Ph3SiCl and the mono-sodium salts of amino acids such as glycine, alanine, methionine, leucine, valine or phenylalanine [45]. This series of proposed pentacoordinate silicon chelates is complemented by a series of proposed hexacoordinate silicon bis-chelates obtained in a similar manner from Me2SiCl2 and two equivalents of the amino acids’ mono-sodium salts [46]. The protic solvent used for the syntheses (i.e., methanol), however, is likely to react with the chlorosilane starting materials (i.e., formation of methoxysilanes [47]), and none of the proposed chelates has been supported by crystallographic evidence. So far, only six examples of Si-complexes with (O,N)-chelating mono-anions of α-amino acids have been characterized crystallographically (Figure 3) [48]. The authors obtained these complexes from reactions of bis(trimethylsilylated) α-amino acids and isocyanatosilanes (Si(NCO)4, MeSi(NCO)3, PhSi(NCO)3). The hexacoordinate Si-complex 1 derived from l-alanine features cis-arranged monodentate NCO groups trans to the chelates’ carboxylate O atoms. The two formally dative N–Si bonds are trans to one another. The pentacoordinate silicon complexes 2a2e feature both a mono- and a di-anionic (O,N)-chelator derived from the same amino acid. In complexes 2a2d (l-alanine and l-phenylalanine used as ligands), the NH2 group of the mono-anionic chelator occupies an axial position in the almost trigonal–bipyramidal Si-coordination sphere. In the l-tert-leucine derivative 2e, both axial positions are occupied by the carboxylate groups. Selected geometric features of the Si-coordination spheres of these compounds are listed in Table 1. Poor solubility and decomposition in [D6]DMSO inhibited NMR spectroscopic investigations of the stereochemistry of the pentacoordinate Si-compounds 2a2e in solution [48].
In previous studies, Tacke et al. reported on zwitterionic Si-bis-chelates, in which two di-anions of α-amino acids serve as the (O,N)-chelators (Figure 4) [49,50]. The syntheses of this class of compounds involved a dehydrogenative coupling of the amino acids with the silane H3Si–CH2–(NC5H6-2,2,6,6-Me4). In these compounds 3a3f, the di-anions of glycine, l-alanine, l-phenylalanine, l-valine, l-tert-leucine and l-proline, respectively, are employed as chelators. A common feature of these complexes is the trigonal–bipyramidal Si-coordination sphere with axial orientation of both Si–O bonds. In such a trigonal–bipyramidal arrangement, chiral amino acids (as in 3b3f) may give rise to the formation of two different diastereomers (3 and 3′). The solid-state structures reveal that the one or the other may crystallize (e.g., 3b corresponds to type 3, 3c corresponds to type 3′). In solution, as monitored NMR spectroscopically, pairs of these diastereomers may coexist in epimerization equilibria (as in the case of 3b), whereas others (e.g., 3c) do not undergo epimerization in solution at room temperature.
So far, there is no crystallographic evidence for the successful isolation of silacycles of α-amino acid di-anions with a tetracoordinate Si atom (neither cyclic 4 nor spirocyclic types 4 in Figure 5). Nonetheless, the formation of compounds of type 4 as reaction intermediates has been reported in the literature [51]. In principle, silylation of the anionic C- and N-termini of the amino acid is possible, and two-fold silylated derivatives can be isolated, e.g., the bis(triisopropylsilyl) derivative of tryptophan 5 [52]. The five-membered heterocycles, however, exhibit so-called ring-strain release Lewis acidity [53], and therefore they may require a fifth donor atom in the Si-coordination sphere for further stabilization. The successful isolation and structural characterization of ammonia adduct 6a [39] and of N-methylimidazole adducts 6b and 6c [37] are in support of the Si atom’s affinity to a fifth donor atom when incorporated in a five-membered cycle with an amino acid’s di-anion. These compounds were prepared from the respective amino acid and a silicon precursor with Si–N bonds, which either releases the required Lewis base (i.e., hexamethyldisilazane with the release of ammonia in the case of 6a) or triggers the addition of an alternative Lewis base (i.e., starting from Me2Si(NHtBu)2 and N-methylimidazole in the cases of 6b and 6c). In adducts 6a6c, the Si-coordination spheres are close to trigonal–bipyramidal, in all cases with the carboxylate O and the additional Lewis base in the axial position while the anionic N-terminus of the amino acid occupies an equatorial site. In chloroform solution, these adducts dissociate (with the formation of a variety of products in the case of 6a [39] and in coexistence with the silacycle+NMI system in a dynamic equilibrium for 6b, 6c and the related l-valine derivative [37]).
Table 1 contains selected parameters of the Si-coordination spheres of the compounds mentioned in Section 2.1.1. As compound 1 is the only hexacoordinate Si-complex in this series, we cannot deduce any trends in dependence on substitution patterns. Comparison of the other 14 compounds, all of which feature pentacoordinate Si, reveals some mutual features. The distorted trigonal–bipyramidal Si-coordination spheres are founded on axial positioning of O atoms or a combination of one O atom and one non-ionic N-donor moiety. None of the complexes have any amide N atoms in the axial position. The complexes of the series 2a2e reveal the most pronounced deviation of this axis from linearity. However, there is no clear trend toward forming a square-base pyramidal coordination sphere by pronounced widening of one of the equatorial angles. Each of the compounds 2a2e features two widened equatorial angles in the range of 120.1–125.9°. The (all-equatorial) Si–N(amide) bonds of the 14 complexes were found in a rather narrow range of 1.69–1.74 Å. The (all-axial) Si–O bonds of the di-anionic chelators span a wider range (1.79–1.88 Å), which exceeds the span of the bond length range of the axial Si–N bonds to the non-ionic donor sites (1.97–2.04 Å). As to the latter, the positional change of this formally dative bond into an equatorial position (encountered with compound 2e, 1.88 Å) has greater influence on this bond length than the variation in substitution patterns in the other 13 compounds. This shortening of the formally dative Si–N bond is accompanied by lengthening of the Si–O bond to the same chelating ligand. The resultant Si–X (X = O,N) bond lengths of the mono-anionic chelator in 2e are similar to those encountered with the mono-anionic chelators in the hexacoordinate Si-complex 1.

2.1.2. Ge-Complexes Derived from α-Amino Monocarboxylic Acids

While in the field of Si-coordination chemistry, a series of different classes of complexes with mono- and/or di-anionic α-amino acid-derived (O,N)-chelate complexes have been reported in the past two decades, there is no crystallographic evidence for the isolation of related Ge complexes. Moreover, no monodentate binding modes of an HN–C–C(=O)O motif to Ge are yet represented in the Cambridge Structural Database (CSD). Nonetheless, some amino acid-derived germacycles were reported by Lavayssiere et al., e.g., the Et2Ge derivatives of the di-anions of valine and of α-amino isoburyic acid [54,55]. The authors report the use of Et2GeCl2 and a supporting base (Et3N) as well as the route via aminogermane Et2Ge(NEt2)2 for syntheses of that kind of germacycle. In a very recent report, Ignatyev et al. mentioned the formation of the compound (glycinate)Ge(OH)3, prepared from GeO2 and glycine, in which the Ge atom is chelated by a mono-anion of glycine [56].

2.1.3. Sn-Complexes Derived from α-Amino Monocarboxylic Acids

As found for Ge, the CSD currently lacks entries for (O,N)-chelate complexes of Sn with mono- and/or di-anionic α-amino acid-derived ligands. However, Ho et al. reported the structure of trimethyltin glycinate 7a [57] (Figure 6), obtained from Me3SnOH and glycine upon azeotropic removal of the water formed in their reaction. In this polymeric compound, the Sn atom is pentacoordinate because of the two-fold monodentate (and thus bridging) coordination mode of the amino acid’s mono-anion. In a related manner, the (O,S)-di-anion of cysteine operates as a bridging ligand in compound 7b [58]. 7b and related derivatives of l-cysteine (with SnEt2, SnBu2, SnPh2 and SnBn2 moiety) were prepared from the respective diorganotin dichloride, l-cysteine and NaOEt. In contrast to 7a, compound 7b features a five-membered chelate about the Sn atom, but this (S,N)-bidentate binding mode is special to the anion of this particular amino acid. Some related chelates of Sn(IV) (8a [59], 8b [60]) and Sn(II) (8c [61]) were reported for mono-anions of cysteine esters. Even though they are somewhat out of scope (lacking the (O,N)-chelating motif), they allow for a comparison of the Sn-coordination of the amino acid’s NH2 group. Comparison of 7a and 7b shows that the chelate and the presence of S (instead of the third C atom) in the trigonal–bipyramidal Sn-coordination sphere strengthen the H2N–Sn-coordination to some extent, whereas the trans-disposed Sn–O bonds are similar in both compounds. Compound 8a, albeit featuring an axial Sn–Cl bond, resembles a cutout of the structure of 7b with similar bond lengths associated with the (N,S)-chelate at Sn. Comparison of 8a and 8b indicates that enhanced Lewis acidity of the Sn center (Sn-bound Me groups replaced by Cl and another (N,S)-chelate) strengthens the coordinative H2N–Sn bonds. Longer Sn–N and Sn–S bonds are observed with the related Sn(II) compound 8c.
Even though there is no crystallographic analysis in support of the (O,N)-chelating properties of α-amino acid mono-anions at tin, some stannacycles of that kind were reported in the literature. Djurdjevic and Djokic reported complexation of Sn(II) by glycinate [62], and Nath and Yadaf reported the (O,N)-chelate of the mono-anion of methionine at SnMe3 [63]. In the context of the known polymeric structure of (glycinate)SnMe3 (7a) [57], a detailed elucidation of the coordination features of these two compounds is desirable for comparison. Moreover, related chelates were reported for other stannane and amino acid combinations by Cashion et al. [64] and Nath et al. [65].

2.1.4. Concluding Remarks Regarding Si-, Ge- and Sn-Complexes Derived from α-Amino Monocarboxylic Acids

In the course of the past two decades, structural studies of Si-complexes of amino acids revealed that this portfolio of compounds is capable of binding to Si in a chelating manner, both as a mono- and as a di-anionic (O,N)-chelator. In contrast, crystallographic characterization of tin compounds of simple amino acids dates back more than four decades, and yet the structural data available to date are still very scarce. Besides the mono-anion of glycine, which was reported as a bridging ligand between SnMe3 groups, only cysteine and derivatives thereof were found to chelate the Sn atom (both Sn(IV) and Sn(II)) in a (S,N)-bidentate fashion. In all cases, the amino group of the ligand operates as a neutral donor moiety. Structural evidence of simple α-amino acid anions in (O,N)-chelates at Sn is yet to be delivered. This and the current lack of corresponding structural data for related Ge compounds clearly mark out some fields for further exploration. In principle, bidentate (O,N)-chelating motifs with anions of simple α-amino acids should be feasible at the heavier tetrels as well. Even though Sn is larger and more thiophilic than Si, some studies of the even heavier and more thiophilic Pb(II) (i.e., compounds Pb(leucinate)(NO3) [66], Pb(phenylalaninate)2 [67], Pb(prolinate)2 [68] and Pb(aspartate) [69], cf. entries A3a, A3b, A3c and A3d, respectively, in Table A3) have clearly shown that mono-anions of various α-amino acids can be bound to Pb as (O,N)-chelators, thus being capable of binding at the smallest, i.e., Si(IV) (e.g., in compounds 1 and 2a2e), and the largest, i.e., Pb(II), of the heavier tetrels. The particularly high stability of Pb(II) and the therefore possible syntheses in aqueous solution with no risk of oxidation of the tetrel and the availability of Pb(II) salts with weakly coordinating anions such as nitrate or perchlorate may have contributed to the success of the exploration of this field of coordination chemistry from this end of the tetrel series, because amino acids (as zwitterionic compounds with good solubility in water) may represent good reactants in an aqueous medium, in particular. The (O,N)-chelation by di-anionic amino acid-derived ligands was reported for Si and Ge (even though representatives of the latter still require detailed structural elucidation). In Sn-coordination chemistry, the (O,N)-di-anions of amino acids still need to be explored as ligands.
Apart from structural elucidation and gaining additional information about the tetrel coordination behavior of mono- and di-anions of simple α-amino acids, Cota et al. pointed out that complexes such as 2a and 2c (which comprise a SiMe motif and amino acid ligands as hydrolysable groups) may be interesting precursors for methylsilanetriol (MeSi(OH)3) as a food supplement [48]. Moreover, the combination of amino acid-derived chelators and amines at the same Si atom (as in 6a) requires further exploration as starting materials for syntheses of the respective α-amino carboxylic amides [39].

2.2. Si-, Ge- and Sn-Complexes Derived from Dipeptides

2.2.1. Si-Complexes Derived from Dipeptides

In 2021, the first structure of an organosilicon complex with a di-anionic homodipeptide ligand was published [39]. This compound 9a (Figure 7) formed in a triethylamine supported reaction of Me2SiCl2 and l-valine and represents a Si-templated synthesis of a dipeptide from an unprotected amino acid. In 2022, a crystallographic analysis of a related compound (9b) was published by Hattori and Yamamoto. The heterodipeptide ligand of compound 9b was deliberately synthesized in the Si-coordination sphere by using Ph2Si(imidazolyl)2, α-amino isobutyric acid and the tBu-ester of glycine in a Ta(OEt)5-catalyzed manner [51]. The authors reported the suitability of this route for the deliberate formation of a great variety of Si-dipeptide complexes. Cota et al. already mentioned the formation of peptides as a side-reaction when utilizing silylated amino acids as starting materials (in the context of their preparation of compounds 1 and 2a2e) [48].
In compounds 9a and 9b, the Si atoms are situated in distorted trigonal–bipyramidal coordination spheres with the Si–O bond and the formally dative Si–NH2 bond in axial positions. In this regard, they are related to compounds 2a2d and 6a6c. The axial angles (165.2 and 165.4° in 9a, 166.4–167.6° in 9b) exhibit slightly greater deviation from linearity than in compounds 2a2d, which can be attributed to the tridentate (O,N,N)-chelate with two five-membered rings about Si. The lengths of the axial Si–N bonds correspond very well to the related axial bond lengths of compounds 2a2d and 6a6c. The di-anionic motif in compounds 9a and 9b, however, exhibits shorter axial Si–O bonds than in 6a6c and longer equatorial Si–N bonds (with respect to 2a2d and 6a6c).

2.2.2. Ge-Complexes Derived from Dipeptides

Crystallographic studies of related germanium complexes with di-anionic dipeptide ligands were reported for compounds 10a [70] and 10b [71], and further compounds (such as (gly-ala)GeMe2 [71]) were prepared in a similar manner, i.e., from Me2GeBr2 and the free dipeptide in toluene with Et3N as a sacrificial base. Selected bond lengths of their Ge coordination spheres are shown in Figure 7. The coordination spheres of Ge in 10a and 10b are essentially related to those of Si in 9a and 9b. The larger covalent radius of Ge causes a more pronounced deviation of the axial angle from linearity (O-Ge-NH2 161.8(1)° in 10a, 164.4(2)° in 10b). Interestingly, the different bonds of the dipeptide ligand respond differently to the change to the heavier tetrel. Whereas both the axial and the equatorial Si–N bonds are ca. 0.1 Å longer in the Ge compounds, the Ge–O bonds are ca. 0.2 Å longer.
Another class of Ge-complexes with dipeptide ligands (GeCl moiety with a mono-anionic and a di-anionic glycylglycine-derived ligand) has been reported by Giuffrida et al. [72]. The authors studied the molecular conformation of this compound with the aid of computational methods.

2.2.3. Sn-Complexes Derived from Dipeptides

In contrast to the few examples of structurally characterized dipeptide complexes of Ge, a great variety of related diorganotin complexes has been synthesized (and investigated crystallographically). The syntheses started from the free dipeptide, and the synthesis routes can be divided into three general approaches: (A) reaction of a dipeptide and a diorganotin oxide with formation of water, (B) reaction of a dipeptide and a diorganotin dialkoxide with liberation of the respective alcohol, (C) reaction of a dipeptide with sodium alkoxide (for in situ preparation of the dipeptide di-anion with liberation of the respective alcohol) and a diorganotin dichloride with formation of NaCl. As a representative example, compound 11a [73] is listed in Figure 7. In this particular compound, the Sn atom is also coordinated in a distorted trigonal–bipyramidal manner, and with a more pronounced position of the heavier tetrel out of the tridentate (O,N,N)-chelate clamp (the axial angle is 153.0(2)°). With respect to the related Ge compounds, the formal move out of the chelate is accompanied by rather proportional lengthening of the three bonds to the chelating ligand. The variety of dipeptide-Sn-complexes, which were structurally characterized in the course of these studies, revealed a portfolio of tin coordination patterns in their solid-state structures, which depend on the absence vs. presence of additional intramolecular coordination interactions (Figure 8). They range from tin pentacoordination in a distorted trigonal–bipyramidal fashion (Figure 8a, as in 11a), via [5 + 2]- or [5 + 1]-coordination of two O atoms or one O atom of an adjacent carboxylate group (Figure 8b,c, respectively) to [5 + 1]-coordination by a nitrogen atom of an adjacent histidine moiety (Figure 8d).
Table 2 contains selected bond lengths and angles of compounds 11a11p [73,74,75,76,77,78,79,80,81,82,83,84,85]. The features of the Sn-coordination spheres of the pentacoordinate Sn-compounds of this series are related to those of 11a. Additional remote [5 + 1]- or [5 + 2]-coordination, however, caused pronounced widening of the C-Sn-C angle in most cases (less pronounced in 11k). Interestingly, with respect to bonds to the tridentate dipeptide ligand, the increase in Sn-coordination number mainly causes lengthening of the formally covalent bonds (of Sn–O, in particular), whereas the formally dative bond to the NH2 group is less affected. We attribute this lengthening of the Sn–N bond to the fact that the increase in coordination number occurs in this equatorial plane where Sn–N is located. Pronounced lengthening of Sn–O rather Sn–NH2 on the O-Sn-NH2 axis indicates a pronounced affinity of Sn atom of these compounds to the coordination of the amino group. This is also supported by the fact that the Sn–NH2 bonds in compounds 11a11p are markedly shorter (by ca. 0.2 Å) than the corresponding bonds in the pentacoordinate Sn(IV) compounds 7a, 7b and 8a. The remote [5 + 1]- or [5 + 2]-coordination in this class of compounds can be interpreted as an effect in the solid state. Solution-state 119Sn NMR data of compounds such as 11j (δ 119Sn–198.1 ppm) and 11k (δ 119Sn –200.7 ppm) indicate pentacoordination of the Sn atoms (in this case, in [D6]DMSO solution) [81]. In deuterated methanol, Sn pentacoordination is retained as well, as shown for 11b (δ 119Sn–160.6 ppm) [74] as well as 11m (δ 119Sn–175.9 ppm) and 11n (δ 119Sn–183.4 ppm) [83].
This set of structurally characterized diorganotin compounds with (O,N,N)-chelating dipeptide ligands is complemented by further organotin compounds of amino acids and dipeptides, e.g., [86,87,88]. The particular interest in this class of compounds arose from the physiological activity of many organotin compounds, which made these combinations with bio-related molecules (i.e., dipeptides) attractive candidates for the exploration of their, for example, anti-tumor [76,83] anti-bacterial [81] and anti-inflammatory activity [80].

2.2.4. Concluding Remarks Regarding Si-, Ge- and Sn-Complexes Derived from Dipeptides

Regarding the structural motifs encountered with di-anionic ligands of peptides at the heavier tetrels, only compounds with a single (O,N,N)-ligand chelating the tetrel have been reported so far. The ligands’ di-anionic nature in combination with the stability of E = Si(IV), Ge(IV), Sn(IV) and these tetrels’ capability of forming stable hexacoordinate complexes, however, should allow for access to dipeptide complexes of the type E(O,N,N)2. Even though structural evidence is scarce, some representatives of E(O,N,N)2 with other tridentate di-anionic chelators capable of chelating via the formation of five-membered rings only (e.g., the compounds XII [89] and XIII [90] shown in Figure 9) serve as motivation for the exploration of such complexes with dipeptide-derived ligands as well.
Moreover, to our knowledge, structural characterization of Si-, Ge- or Sn-complexes of anions of longer oligopeptides has not been reported so far. However, Hattori and Yamamoto [51] reported the utilization of silicon dipeptide complexes as starting materials for selective peptide chain elongation. This provides a perspective on a way to access and characterize silicon complexes of longer oligopeptides. Also, the utilization of silicon compounds as templates for the deliberate syntheses of peptides from unprotected [39,91] or protected amino acids [51] is a very young and promising field of amino acid coordination chemistry at heavier main group elements. In addition to organosilicon compounds, organoaluminium compounds appear to be useful tools in peptide syntheses as well [92].
From the viewpoint of coordination chemistry, the special amino acid motif of proline, which should exhibit different coordination features when incorporated into the dipeptides’ N-terminus positions because of its cyclic amine moiety, still warrants exploration.

2.3. Si-, Ge- and Sn-Complexes Derived from Amineoligoacetic Acids and Related Hydroxyalkylamineacetic Acids

2.3.1. Si-Complexes Derived from Amineoligoacetic Acids and Related Hydroxyalkylamineacetic Acids

Silatranes (general motif 12 in Figure 10) represent an old and well-reviewed class of silicon complexes derived from triethanolamine. Substitution of one, two or three alcoholate groups with carboxylate groups leads to the classes of silatranones, silatranediones and silatranetriones, respectively. The molecular structures of silatranones 12a12f [93,94,95,96,97] and silatranedione 12g [98] were analyzed by X-ray crystallography, and the molecules were found to resemble the molecular architecture of related silatranes. For example, in methylsilatrane [99] and phenylsilatrane [100], the Si–N bond lengths are 2.16 Å, the Si–O bond lengths are in the range of 1.65–1.68 Å and the sum of O-Si-O angles amounts to 356°. The ranges of Si–N bond lengths (spanning 2.09 Å in 12b and 2.15 Å in 12a) are very similar, and the sums of O-Si-O angles (spanning 355° in 12c and 357° in 12b) are very similar, too. Even though the Si–O bonds are located in similarly planarized equatorial environments around the Si atoms, their bond lengths are noticeably different for the alkoxy- vs. carboxy-O atoms. The former span a range of 1.64–1.66 Å, whereas the latter are longer (in the range of 1.71–1.72 Å in compounds 12a12f and 1.70 Å in silatranedione 12g). The equatorially situated Si–O(carboxylate) bonds, however, are markedly shorter than the axial Si–O(carboxylate) bonds in the compounds listed in Table 1 and compounds 9a and 9b. Silatranetriones have also been reported [101], but their molecular structures have been analyzed by spectroscopic means only. Their 29Si NMR shifts (e.g., δ –135.8 ppm for methylsilatranetrione N(CH2COO)3SiMe in [D6]DMSO, whereas the corresponding methylsilatrane N(CH2CH2O)3SiMe, methylsilatranone (12f) and methylsilatranedione (12g) have 29Si NMR shifts of δ –69.7, –73.2 and –77.2 ppm, respectively, in the same solvent [101]) indicate a higher coordination number of their Si atoms, and the authors attributed this to the additional coordination of a solvent molecule. The presence of a single set of 1H and 13C NMR signals for the three CH2COO moieties indicates rapid site exchange of the coordinated solvent [101]. A detailed elucidation of their molecular structures is yet to be delivered to allow for discussion of the effects of solvent coordination on the bond lengths and angles of the atrane cage.
Structural analyses of quasisilatranones and quasisilatranediones (i.e., silatranone-related Si-compounds with tridentate (O,N,O)-chelators with a central NH moiety) have not been reported so far. Moreover, we are not aware of reports of silicon complexes of chelating anions derived from ethylenediamine tetraacetic acid, diethylenetriamine pentaacetic acid or related extended versions of nitrilotriacetic acid.

2.3.2. Ge-Complexes Derived from Amineoligoacetic Acids and Related Hydroxyalkylamineacetic Acids

In sharp contrast to the structural information available in the field of silatranone-type silicon complexes, the current state of crystallographic characterization of related Ge compounds lies beyond the Si congeners, with no overlap that would allow for a direct comparison of molecular structures. Even though germatranones, germatranediones and germatranetriones have been reported in the literature [101], only one crystal structure has been reported out of this group of compounds, i.e., the hydrate of hydroxygermatranetrione 13 [102] (Figure 11). In this complex, the Ge atom is hexacoordinated, and a water molecule serves as a sixth ligand moiety, located in the otherwise equatorial GeO3 plane of the atrane motif. In this compound, the length of the Ge–OOC bond trans to the Ge-bound water molecule (1.88 Å) is only marginally shorter than the mutually trans Ge–OOC bonds (1.89 and 1.90 Å). They are closer to the length of the dative Ge–OH2 bond (1.92 Å) than to the length of the formally covalent Ge–OH bond (1.76 Å). The structure of the corresponding anionic complex 13′ (K+ salt, two OH groups bound to Ge) has been reported as well [103]. The conformation of this anion resembles the molecular shape of 13. The stronger coordination of the additional OH group trans to the carboxylate (Ge–OH bond length 1.82 Å vs. Ge–OH2 bond length 1.92 Å in 13) causes particular lengthening of the trans-disposed Ge–OOC bond (1.97 Å), while the mutually trans Ge–OOC bonds (1.91 Å) remain shorter. The related compound 14 features a phosphonic acid moiety in one of its equatorial atrane positions (with respect to the N-Ge-OH axis) [104]. In this case, the anionic phosphonate serves as a ligand, which enhances the Ge coordination number to six by dimerization of this anionic complex via Ge-O-P-O-Ge bridges. The two Ge–OP bonds (1.87 and 1.88 Å) are shorter than the two Ge–OOC bonds (1.90 and 1.93 Å). The Ge–OH bond (1.77 Å) is similar to the corresponding bond in 13. The Ge–N bonds are very similar in compounds 13, 13′ and 14 (2.08, 2.08 and 2.09 Å, respectively).
Moreover, detailed structural analyses are available for Ge-complexes derived from oligoamine–oligoacetic acids. This class of compounds is accessible in acidic aqueous medium from GeCl4 [105]. Most of them feature the ethylenediamine tetraacetic acid motif (compounds 15 [106,107,108,109,110], Figure 12, Table 3), and one of them has a propane-1,3-diamine backbone (compound 16 [111]). This difference, i.e., extension of the ligand backbone by one additional carbon atom, allows the diamine tetraacetate to bind to the Ge atom in a hexadentate manner, as shown in Figure 12 for compound 16, whereas in compounds of type 15, the diamine oligoacetate ligand always binds in a pentadentate manner, leaving a dangling ligand arm R, while an OH group occupies the sixth ligand site at Ge. As shown in Table 3, the seven representatives of compound class 15 (15a15f [106,107,108,109,110], where the crystal structure of 15e features two independent molecules in the asymmetric unit) exhibit very similar bond length patterns on their Ge coordination spheres. There is a trend that the Ge–N bond a of the atrane-like N atom (which features three Ge-binding arms) is similar to those in compounds 13, 13′ and 14, whereas the Ge–N bond b of the N atom with a dangling substituent is slightly longer. As bond a is trans to a more tightly bound anionic O-donor than bond b (i.e., bond f is shorter than bond d), we attribute the shortness of bond a (and the similarities of this bond length to those in compounds 13, 13′ and 14) to the bond-enforcing action of the tripodal ligand motif in atranes and related compounds. These interpretations are consistent with the bond lengths observed in compound 16, which features shorter Ge–N bonds. In 16, both N atoms are part of tripodal ligand motifs, and the Ge-coordination sphere is devoid of such a tightly bound O-donor site (i.e., OH) as in compounds of type 15, which would compete with the trans-disposed Ge–N bond. If in compounds of type 15 the dangling arm R is a chelator such as CH2CH2N(CH2COO)2 (i.e., 15 being a derivative of diethylenetriamine pentaacetic acid), additional cations can be added to this remote ligand site with the formation of related heteronuclear complexes, such as [“Ge”-CH2CH2N(CH2COO)(CH2COOH)]2Cu, with “Ge” being Ge-complex motif 15 with the R′ = CH2CH2 backbone [112] (cf. entry A3e in Table A3). A different kind of heteronuclear germatranone originates from 2-propanol-1,3-diamine tetraacetic acid (i.e., from the 2-OH functionalized derivative of the ligand used in 16). This ligand gives rise to the formation of germatranedione cages with the alcoholate ligand arm bridging the Ge site and the adjacent heterometallic site in complexes of types 17 and 17′. In motif 17 (encountered with lanthanide complexes 17a17e [103,113,114,115,116]), the alcoholate donor site at Ge is trans to a Ge–OH bond, and the two Ge–OOC bonds are mutually trans. In type 17′, the two Ge–OOC bonds are mutually cis, and another equatorial O-donor ligand (“O”) is bound in one of the cis positions next to the alcoholate. In compound 17f, “O” is a water molecule. In MLn = Cu(bipy) [117] and Zn(H2O)2 complexes [118] (cf. Table A3 entries A3f and A3g, respectively) of a more complicated multinuclear architecture, a carbonyl O atom of an adjacent complex moiety occupies this position. The Ge–N, Ge–OOC and Ge–OH bond lengths in the complexes 17a17f (Table 4) are longer than corresponding bonds in germatranetrione 13.

2.3.3. Sn-Complexes Derived from Amineoligoacetic Acids and Related Hydroxyalkylamineacetic Acids

In contrast to the Si- and Ge-chemistry shown in Section 2.3.1 and Section 2.3.2, respectively, structural characterization of tin complexes of amineoligoacetic acid derivatives has been successful with a greater variety of ligands. Particularly interesting is the set of compounds derived from diacetic acid derivatives (compounds 18, Figure 13). The N-bound substituent R may be H or hydrocarbyl (a non-coordinating residue), and for a set of diorganotin compounds derived therefrom (18a18g [119,120,121,122,123,124,125]), two principle coordination motifs are encountered. As shown for 18a, the di-anionic (O,N,O)-ligand was found to bind to Sn in an axial–equatorial–axial manner, and the equatorial Sn–C bonds span an angle closer to the expected 120° for equatorial positions in a trigonal bipyramid. With the aid of additional ligands L (such as water) and dimerization via additional remote coordination of a carboxylic O atom of an adjacent complex molecule, the Sn-coordination sphere transitions toward pentagonal–bipyramidal with axial Sn–C bonds and equatorial positioning of N and O atoms (as shown in Figure 13 for 18c). As shown in Table 5, this adoption of pentagonal–bipyramidal coordination is similar for compounds 18c18g, all resembling a [6 + 1]-coordination sphere ([4 + 1] in the equatorial plane) with the carboxylate O of an adjacent molecule being the remote lone-pair donor. The C-Sn-C angle widening (toward axial positions) is more pronounced than in the dipeptide complexes 11e, 11g and 11i, which also feature additional remote coordination in the solid state (cf. Table 2). Moreover, the related complex 19 (Figure 13, [126]) adopts an intermediate role in this regard. Its Sn atom is hexacoordinated, and its C-Sn-C angle is 141.7°. The different electronic situation in 19, however, causes further differences, and therefore rather withdraws this complex from comparison within a series. That is, the different bridging atom (alcoholate instead of carboxylate) forms short bonds to both Sn atoms (2.07 and 2.34 Å of the intra- and intermolecular bond length, respectively), whereas the Sn–O(carboxylate) bond is rather long (2.49 Å). Other changes in the ligand (such as a coordinating substituent R, a 2-hydroxyethyl group) may give rise to different coordination patterns, i.e., the stannatranedione motif with hexacoordinate Sn in 20 [127]. In spite of the additional ligand arm, the Sn–N bond length in 20 (2.32 Å) is similar to those in compounds 18a18g. Whereas in compounds 18 and in 20 the (O,N,O)-ligand binds to Sn in a mer-like fashion (positioning of both O,N,O and Sn close to an idealized plane), the anionic complex 21 [128] represents a structurally characterized example of this ligand binding to octahedrally coordinated Sn in a fac configuration. In addition to the different configuration of the ligand and the set of electron-withdrawing substituents at Sn, the Sn–N bond in 21 (2.22 Å) is markedly shorter than those in compounds 18 and 20. Extension of the ligand backbone, toward ethylendiamine diacetate (OOCCH2NHCH2CH2NHCH2COO)2–, also allows for the formation of octahedral Sn-complexes, as shown for the SnMe2 derivative, which has a trans-N-Sn-C and trans-O-Sn-O arrangement [122] (cf. Table A3 entry A3h).
Using nitrilotriacetate, Sn-complexes with a penta- (22) [129], hexa- (23) [130], hepta- (24) [131] and octacoordinated Sn atom (25) [132,133] were prepared, the molecular structures of which were characterized (Figure 14). Compound 22 was prepared from the corresponding triiodostannyl complex and nitrilotriacetic acid in dichloromethane in a base-supported substitution reaction with retention of the Sn–Os bond. Compound 23 formed in a Sn–C(Ph) bond cleavage reaction of Me2N(CH2)3SnPh3 and nitrilotriacetic acid in DMF. Hence, in addition to the liberation of benzene as the expected leaving group, the NMe2 moiety was converted into the corresponding amine N-oxide, which furnishes the Sn-coordination sphere of 23. Salts of the anions 24 and 25 were accessible in aqueous solution.
The Sn-coordination sphere in 22 is trigonal–bipyramidal, with the Sn–Os bond in the axial position and a trans-disposed N–Sn bond. In 23, the tripodal ligand adopts a configuration as shown in Figure 14, which is related to the structures of compounds 20 and 21, combining their Sn-OOC positions in one molecule. This arrangement in an octahedral Sn-coordination sphere may play an important role in the Sn–N bond length of this molecule, because this bond length is in the lower range for this set of three octahedral Sn-complexes (2.23 Å in 23, 2.22 Å in 21, 2.32 Å in 20), whereas the penta-, hepta- and octacoordinated Sn-compounds in Figure 14 feature markedly longer Sn–N bonds (2.36 Å in 22, 2.32 and 2.35 Å for the tetra- and the tridentate ligand site, respectively, in 24, 2.37 Å in the K-salt of 25, 2.38 and 2.40 Å in the methylammonium salt of 25).
The molecular conformation of anion 24 can be related to the atrane-type cage (as in 22) for the tetradentate ligand, and the three bonds to the donor atoms of the tridentate ligand are staggered about the O-Sn-O angles of the former. In contrast, in di-anionic octacoordinated Sn-complex 25, the two tetradentate ligand moieties adopt conformations which are comparable to those in 23, and a Sn–O bond of one ligand opposes the vacant face of the other ligand (and vice versa), as indicated by the bold-style O and N atoms in Figure 14 (vacant face at bold N of one ligand exposed to the bold O of the other ligand and vice versa).
This trend toward coordination numbers 7 or even 8 of the Sn atom, which contrasts the Sn-coordination chemistry from related Ge- and Si-chemistry, is also evident in Sn-complexes of oligoamine oligoacetic acid-derived ligands. Various derivatives of Sn(EDTA) (EDTA being the tetra-anion ethylenediamine tetraacetate) of the general molecular building pattern 26 [134,135,136,137,138,139] (Figure 15) were characterized crystallographically, and their mutual feature is a heptacoordinate Sn in a coordination sphere, which can be described as a distorted pentagonal bipyramid. As shown in Figure 15, the axial positions are occupied by carboxylate O atoms (forming the shortest Sn–O bonds in this coordination sphere), and the sets of atoms in the idealized pentagonal plane (two N atoms, two further carboxylate O atoms, one additional lone-pair donor ligand L) are distorted against one another in terms of a twist of the N2Sn plane against the SnO2L plane. This building pattern (although more distorted off the bipyramidal shape of compounds of type 26) is in general retained in Sn(IV) complexes derived from diethylenetriaminepentaacetic acid (motif 27, both the protio form [140,141] and the ammonium salt [142] were characterized crystallographically). Their building pattern is related to that of compounds 26 by replacing an equatorial O atom with the third amine N, accompanied by the replacement of ligand L with O atoms of the additional two carboxylate moieties. Thus, in compounds of type 27, the Sn atom is octacoordinated and completely encapsulated in the octadentate ligand. Also related to the molecular structures of 26 and 27 is tin(II) compound (EDTA)Sn2 (28) [143]. The solid-state structure of 28 features two different Sn(II) coordination spheres; one Sn atom is located in a SnO4-coordination sphere, surrounded by carboxylate O atoms (it serves as the bridging Sn site, indicated by “(Sn)” in Figure 15), whereas the other Sn atom is located in the EDTA chelate in a SnN2O4-coordination sphere (in a [4 + 2]-coordination with two markedly longer Sn-O distances trans to the Sn⋯N bonds), which relates to those of 26 by replacing the seventh ligand L with the Sn-located lone pair, and it relates to 27 in terms of the pronounced distortion off the axial geometry. In sharp contrast to 28, tin(II) compound (EDTA)H2Sn (29) [144] adopts a tetrel coordination sphere related to those of Ge compounds of type 15 (the site of the Ge-bound OH group in 15 is replaced with the Sn-located lone pair in 29).

2.3.4. Concluding Remarks Regarding Si-, Ge- and Sn-Complexes Derived from Amineoligoacetic Acids and Related Hydroxyalkylamineacetic Acids

The exploration of Ge- and Sn-coordination chemistry benefits from the availability of compounds such as EDTA complexes from an aqueous medium. This allows for access to a variety of complexes which feature, for example, an atranetrione or -dione ligand backbone about Ge or Sn. Because of the size difference between Ge and Sn, different coordination modes of EDTA and related anions are encountered (vide supra). Interestingly, EDTA complexes of hs-Fe3+, which has an ionic radius between those of the formal ionic radii of Ge4+ and Sn4+ [145], include the motifs encountered with Ge (in [Fe(HEDTA)(H2O)] [146]) and with Sn (in K[Fe(EDTA)(H2O)] [147]) (cf. Table A3, entries A3i and A3j, respectively). The class of complexes presented in Section 2.3.3. comprises some structurally characterized examples of Sn(II) compounds. In this regard, it serves as a motivator for further exploration of related Sn(II) complexes with ligands mentioned in Section 2.1 and Section 2.2. Moreover, the wealth of coordination chemistry encountered with the Ge- and Sn-complexes in Section 2.3. raises questions as to the facets of Si-coordination chemistry of anions of amineoligoacetic acids beyond silatranediones, with the Si-hexacoordinate silatranetriones as an entrance into this world of coordination chemistry of interesting compounds yet to be characterized in detail.

3. Perspectives Outside the Box

3.1. α-Amino Acid-Derived Schiff Bases as Ligands

In principle, the primary amine motif of an α-amino acid motivates researchers to create related ligands by Schiff base condensation of this group with suitable carbonyl compounds (such as o-hydroxyarylcarbaldehydes, -ketones or acetylacetone). In 1992, Smith et al. reported the first crystal structure of an (O,N,O)-chelated tin compound with a Schiff base derived from an α-C substituted α-amino acid, a compound of type XIV (Figure 16, compound XIVa [148]). It features a di-anionic (O,N,O)-tridentate Schiff base ligand derived from salicylaldehyde and valine, and further compounds of that kind were reported in the course of this investigation. The Schiff base ligands were accessible by condensation reaction of the aldehyde and the free amino acid, and the tin complexes were prepared from the Schiff base ligands and diorganotin oxides [148]. Further compounds of that kind were reported thereafter, e.g., compounds XIVb [149] and XIVc [150] with ligands derived from leucine and alanine, respectively, and it was reported that racemization of the chiral Schiff base compounds may occur [150]. Moreover, modifications at the aryl part of the ligand backbone enhanced the portfolio of these compounds (e.g., XIVd and XIVe [151]); they were pursued as this kind of tin compound was of interest for the exploration of their cytotoxic activity. In the course of the past decade, acetylacetone-derived Schiff bases (e.g., XVa [152], XVb [153], XVc [154]) have enhanced the portfolio of this class of tin compounds even further (also with the aim of exploring their biological activity, e.g., for use against moscito larvae [154]). As found with other diorganotin(IV) compounds with di-anionic tridentate (O,N,O)-ligands (cf. Section 2.2.3 and Section 2.3.3), Schiff base complexes such as compounds of type XIV may establish additional intermolecular interactions, which enhance the Sn-coordination number to 6 (e.g., compound XIVf [155]), [6 + 1] (compound XIVg [156]) or 7 (in XIVh [157]). Within the portfolio of amino acid-derived Schiff base complexes of tin, compounds XVI [158] and XVII [159] are particularly noteworthy as they enhance our understanding of the coordination properties of this class of tridentate ligands. The former features a mono-anionic (O,N,O)-ligand, which also binds to tin in a tridentate manner. The latter represents (to date) the only crystallographically characterized tin complex which features two of these amino acid-derived Schiff base ligands. In both cases, the Sn atom is hexacoordinate, and the (O,N,O)-ligands occupy meridional positions in the octahedral Sn-coordination spheres.
In the past two decades, these studies were complemented by the investigation of some related Ge- and Si-compounds. In 2002, Nath and Goyal reported the syntheses of some organosilicon compounds with amino acid-derived Schiff base ligands and the investigation of their antimicrobial activity (even though the authors also mentioned that these compounds are sensitive to hydrolysis) [160]. Detailed insights into the molecular structures of such complexes were delivered for aldehyde-derived compounds such as XVIII [159], XIXa [161] and XIXb [159], acetylacetone-derived compounds such as XXa [162] and XXb [163] and even for compounds of sterically more demanding imine backbones (XXIa [164] and XXIb [165]) (see Figure 17). It was also pointed out that racemization can present an issue in case of the preparation of silicon compounds of amino acid-derived Schiff base ligands [161].

3.2. α-Amino Acid-Derived Alcohols as Ligands

In addition to the preparation of complexes with mono- or di-anionic α-amino acid ligand motifs, amino acids (especially the chiral representatives thereof) offer great potential for transferring their chiral information to other classes of ligands. The chiral amino alcohol H2N-C*HiPr-CPh2-OH, which can be obtained from valine, can be utilized for the preparation of chiral Schiff base ligands and Si-complexes thereof as well, compound XXII (Figure 18) being a representative thereof [166]. Compound XXIII represents a Si-complex with a related aminoalcohol motif in a tridentate (O,N,O)-ligand that features an amine moiety as the formal dative bond donor [167]. To name a second class of oligodentate ligands, chiral triethanolamine derivatives (and complexes thereof) may also be obtained from chiral amino acids, e.g., compounds of type XXIV, which are based on alanine and valine as the chiral starting materials [168]. While other routes may also be employed (e.g., the -CHPhCH2O- and -CH2CHMeO- moieties in the ligands of compounds XXV [169] and XXVI [170], respectively, originate from ring-opening of styreneoxide and propyleneoxide, respectively), the amino acids from the natural chiral pool offer a variety of building blocks, which can be converted into different kinds of other chiral ligands.

3.3. Ligands Derived from Non-Carboxylic α-Amino Acids

Even though silicon analogs of the type R2N-CH2-Si(=O)OH are not known, the α-amino carboxylic acid motif is related to other heavier main group element (E)-based acids of the type R2N-CH2-E(=O)R′OH (R, R′ being various kinds of residues, e.g., alkyl, aryl, H and, for R′, extending to OH, alkyoxy and others) such as aminoalkylphosphonic or -sulfonic acid derivatives. For some simple representatives thereof, coordination chemistry has been explored to some extent, e.g., syntheses and characterization of the copper complexes of H2N-CH2-PR(=O)O (R = Me, Ph) [171] and H2N-CH2-S(=O)2O [172] (cf. Table A3, entries A3k, A3l and A3m, respectively), in which these mono-anions operate as (O,N)-chelators. A search for crystallographic evidence of these special kinds of α-amino acid motifs acting as chelators in Si/Ge/Sn-coordination chemistry merely yielded some hits for phosphonic and phosphinic acid derivatives (XXVII [173], 14 [104], XXVIII [174], Figure 19). Interestingly, only the Ge representative, which has N clamped to Ge by additional carboxylate buttresses, exhibits a rather short bond between the amine N atom and the heavier tetrel. The coordinative N–E bonds for the E = Sn, Si representatives are markedly longer. For the series of Si-compounds XXVIIIa-e shown in Figure 19, this was rather unexpected; the anion of the trifunctional aminomethylphosphinic acid (nitrilotris(phenylphosphinic acid) [175]) allowed for rather long N–Si distances in spite of the additional buttresses. Of note, the Sn representative XXVII features N–Sn-coordination at the Sn(II) site of this Sn(II)2Sn(IV) trinuclear complex. Sn-complexes of [4-HOOC-C6H4-CH2N(CH2PO3)2]4– (Sn–N 2.92 Å) [176], [–CH(CH2CH2)2NCH2PO32–]2 (Sn–N 2.94, 2.96, 3.23 Å; 2.81, 3.13, 3.19 Å) [177] and CyN(CH2PO32–)2 (Sn–N 2.96 Å) [178] (cf. Table A3, entries A3n, A3o and A3p, respectively) contribute to the portfolio for the heavier tetrel. All of them feature amine-Sn-coordination at Sn(II) with rather long Sn–N distances (listed in parentheses). As for Sn(IV) representatives, structures of other Sn-compounds of α-amino phosphonic and phosphinic acid derivatives were reported in which the respective molecule or anion binds to Sn via O only, leaving behind a vacant N lone-pair donor site. Figure 19 shows some selected examples (XXIX [179], XXX [180], XXXI [181]), and the SnCl2Me2-complex of [(EtO)2(O=)PCH(Ph)-NH-CH2]2 [182] (cf. Table A3 entry A3q) represents a further example. Of note, coordination of amine N to phosphonate-functionalized Sn(IV) at a rather short distance (2.45 Å) is in principle feasible, as shown by Dakternieks et al. for compound MeN(CH2CH2CH2)2SnMe(O(OH)O=PtBu) [183] (cf. Table A3 entry A3r).
As for sulfur analogs, compound XXXII shown in Figure 20 [184] represents the only crystallographically characterized compound with a five-membered (N-C-S-O)Sn-chelate motif, and representatives of Ge and Si are yet to be explored. In principle, compounds such as the SiMe3-derivative XXXIII shown in Figure 20 [185] may turn out to be suitable starting materials for syntheses of (N-C-S-O)E (E = Si, Ge, Sn)-chelates. The compound itself would be a neutral (N-C-S-O)-chelator, but the SiMe3 groups as leaving groups offer its transformation into an anionic chelator. In its Li salt (XXXIV), the anion forms a five-membered (N-C-S-O)Li-chelate [185].

4. Conclusions

This overview of heavier tetrel derivatives of α-amino acids, where anions of the latter operate as (O,N)-chelating ligands, demonstrates the variety of ligands with the α-amino acid motif in Si-, Ge- and Sn-coordination chemistry. Some classes of complexes (e.g., complexes of di-anions of dipeptides) were reported with structural characterization for the series of the three tetrels and thus allow for comparison of the coordination behaviors of the different tetrels. For some classes (e.g., complexes derived from amine diacetic acid), series of crystallographically characterized representatives were reported for one tetrel only (E = Sn), which at least allow for comparison of structural features within the set of complexes. In the context of the availability of a great variety of amino acids from the natural chiral pool, however, the portfolio of structurally characterized Si-, Ge- and Sn-complexes of these ligands is rather limited. Moreover, rather common amine oligoacetic acids (such as nitrilotriacetic acid, ethylenediamine tetraacetic acid) are also highly underrepresented in some fields of tetrel coordination chemistry as far as unequivocal characterization of the molecular structures of the resultant complexes is concerned. Some emerging fields of tetrel coordination chemistry with α-amino acid-derived ligands indicate that the overall topic of tetrel coordination chemistry with these kinds of (O,N)-chelators is a live and growing field. Just to name sections of the timeline and give some examples: The first crystallographic characterization of a tin complex of a chiral amino acid-derived Schiff base ligand dates back to 1992 [148], and examples with Si and Ge followed in 2012 [161] and 2018 [159], respectively. Whereas complexes of di-anions of dipeptides were structurally characterized as early as the 1970s for Sn [85] and 1980s for Ge [70,71], the first crystallographic report of a dipeptide-derived Si-complex was published in 2021 [39]. However, gaps in some series of classes of complexes predominate. Detailed structural characterization is yet to be delivered for, e.g., silatranetriones, which are interesting because of their tendency toward Si-hexacoordination (as indicated by 29Si NMR spectroscopic data) [101]; for EDTA complexes of silicon, which would complement the examples of heptacoordinate Sn- and hexacoordinate Ge-complexes of this well-known chelator; for Ge-complexes of mono- and di-anions of simple α-amino acids, where there are examples of chelating mono- and di-anions of that kind at Si [37,39,48,49,50]; and structural evidence for the bridging coordination of the glycinate mono-anion in polymeric (gly)SnMe3 [57], which raise questions as to the preferred chelating or bridging coordination of these ligands in Ge compounds. In addition to the apparent gaps in series of compounds, this overview sheds light on further fields of coordination chemistry yet to be explored. Tetrel coordination chemistry of other α-amino acids (such as α-amino phosphonic or phosphinic acids) is an almost blank space on the map of structurally characterized compounds. Moreover, in spite of the enhanced stability of di-valent tin (with respect to Ge(II) or Si(II)), the field of Sn(II)-compounds of α-amino acids is rather underrepresented. In this regard, chelates of the type (O,N)2Sn(II) with amino acid-derived mono-anionic (O,N)-chelating ligands may represent interesting ligands themselves in transition metal coordination chemistry because of the lone-pair donor capability of Sn(II). For other chelates, some transition metal complexes have been characterized, e.g., (8-oxyquinolinate)2GeCr(CO)5 [186] and (8-oxyquinolinate)2SnCr(CO)5 [187] (cf. Table A3 entries A3s and A3t, respectively). For α-amino acid-derived complexes of Sn (or even Ge or Si), however, this kind of transition metal coordination chemistry with group 14 ylene ligands is an unexplored field. So far, stannatranetrione compound 22 [129], which features an Os–Sn bond, is the only crystallographically characterized TME compound (TM = any transition metal, E = Si,Ge,Sn), which features an E-bound α-amino acid-derived ligand backbone.

Author Contributions

A.S. and J.W. contributed to the writing and proofreading of this review. All authors have read and agreed to the published version of the manuscript.

Funding

Parts of this work were carried out with financial support from the European Union (European Social Fund, ESF) and the federal state of Saxony (Sächsische Aufbaubank, SAB, Dresden, Germany) via a Ph.D. scholarship (“Landesinnovationsstipendium”) to A. Seidel under Project 100670490-AP2.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

Table A1. List of literature references (Ref.) and CSD reference codes (CSD Refcode) of crystallographically characterized compounds mentioned in this paper with Roman numerals. The crystallographic data (in CIF format) can be accessed from the Cambridge Structural Database (CSD) using the respective CSD Refcode via https://www.ccdc.cam.ac.uk/structures/ (accessed on 1 February 2025 for availability check).
Table A1. List of literature references (Ref.) and CSD reference codes (CSD Refcode) of crystallographically characterized compounds mentioned in this paper with Roman numerals. The crystallographic data (in CIF format) can be accessed from the Cambridge Structural Database (CSD) using the respective CSD Refcode via https://www.ccdc.cam.ac.uk/structures/ (accessed on 1 February 2025 for availability check).
CompoundRef.CSD
Refcode
CompoundRef.CSD
Refcode
CompoundRef.CSD
Refcode
I[26]IHAFAYXIVe[151]KAXVIPXXIII[167]FEYMON
II[27]EYASOOXIVf[155]OGOCIZXXIVa[168]RABLUB
III[28]MOCCITXIVg[156]NEYDEDXXIVb[168]RABMAI
IV[29]DMSNQNXIVh[157]JATHUIXXV[169]IJAXEW
V[30]FEDYOCXVa[152]ZEHQIPXXVI[170]OGABAD
VI[31]FICNUAXVb[153]KALYEEXXVII[173]QIFBIT
VII[32]CEWJAPXVc[154]ESOMACXXVIIIa[174]UGEKOK
VIII[33]WONQOGXVI[158]JILXAHXXVIIIb[174]UGELEB
IX[34]CEKWARXVII[159]ZIGWIYXXVIIIc[174]UGELAX
X[35]WEQRIUXVIII[159]ZIGTUHXXVIIId[174]UGELIF
XI[36]VUKMASXIXa[161]ACIJEDXXVIIIe[174]UGEKUQ
XII[89]VUNKARXIXb[159]ZIGWAG 1XXIX[179]KISROT
XIII[90]IQATOKXXa[162]MUFQAHXXX[180]QIHWEJ
XIVa[148]JUNHABXXb[163]OFOYOZXXXI[181]JIPYEP 2
XIVb[149]TOZNEBXXIa[164]JICTAUXXXII[184]SUPHOC
XIVc[150]EHIZOKXXIb[165]YILVOIXXXIII[185]GIJJIT
XIVd[151]KAXVELXXII[166]WEKFAUXXXIV[185]GIJJAL
1 Another entry of a crystal structure of this compound can be found under CSD Refcode ZIGWEU. 2 Another entry of a crystal structure of this compound can be found under CSD Refcode JIPYOZ.
Table A2. List of literature references (Ref.) and CSD reference codes (CSD Refcode) of crystallographically characterized compounds mentioned in this paper with Arabic numerals. The crystallographic data (in CIF format) can be accessed from the Cambridge Structural Database (CSD) using the respective CSD Refcode via https://www.ccdc.cam.ac.uk/structures/ (accessed on 1 February 2025 for availability check).
Table A2. List of literature references (Ref.) and CSD reference codes (CSD Refcode) of crystallographically characterized compounds mentioned in this paper with Arabic numerals. The crystallographic data (in CIF format) can be accessed from the Cambridge Structural Database (CSD) using the respective CSD Refcode via https://www.ccdc.cam.ac.uk/structures/ (accessed on 1 February 2025 for availability check).
CompoundRef.CSD
Refcode
CompoundRef.CSD
Refcode
CompoundRef.CSD
Refcode
1[48]WUYTIU11g[78]KIHYAB17e[114]GELJUE
2a[48]WUYTOA11h[79]VUPRON17f[116]RAGHIS
2b[48]WUYTOG11i[80]IGUQIL18a[119]OWEYAS
2c[48]WUYVAO11j[81]MOCRAY18b[120]IGEYOI
2d[48]WUYVES11k[81]MOCREC18c[121]KELLAO
2e[48]WUYVIW11l[82]VUPDAL18d[122]TOSRIC
3a[49]GAGJEE11m[83]PATHEX18e[123]GAMYUP
3b[49]GAGJII11n[83]PATHIB18f[124]SARREL
3c[49]GAGJOO11o[84]SASZUH18g[125]LUKDEB01
3d[50]JAZHAU11p[85]GLDPSN19[126]VAZHAG
3e[50]JAZHEY12a[93]CASYOK20[127]VIPDAZ
3f[50]JAZHOI12b[94]CECXEN21[128]QOVVEE
5[52]UYAYEZ12c[95]FMESIA22[129]QINWOZ
6a[39]JAPZOT12d[95]FMESIB23[130]HAWBOW
6b[37]HIVJOP12e[96]FOFXED24[131]LOFJEV
6c[37]HIVJUV12f[97]GEKXID25 1[132]FIBWAO
7a[57]GLYMSN1012g[98]DIRSAY25 2[132]FIBWOC
7b[58]BIGSUG13[102]VUHCIK25 3[133]IWEROR
8a[59]ECYSSN13′[103]LAFHIL25 4[133]IWERIL
8b[60]BONYAE14[104]TEBJOA26 5[134]HIVZUJ
8c[61]VUKLIW15a[106]VEFGOC26 6[135]SNEDTA
9a[39]JAPZUZ15b[107]VOCPIN26 7[136]NUQZEE
9b[51]QANSOR15c[108]YUFYIH26 8[137]QELZUC
10a[70]VASWER15d[109]XOHHEH26 9[137]QEMBAL
10b[71]VAWHEG15e[110]VOCPOT26 10[138]TORTUP
11a[73]DOHBIL15f[108]YUFYON26 11[139]WEVNAP
11b[74]NOCQOM16[111]XUZZEZ27 12[140]KIJWUV
11c[75]QOBXOU17a[103]GEMPEV27 12[141]KIJWUV01
11d[76]VUKHAK17b[113]ELOYIO27 13[142]LOHLEZ
11e[76]VUKHEO17c[114]GELJOY28[143]EDTASN
11f[77]ZUQBOC17d[115]XAVRUL29[144]HEDTAT
1 Potassium salt. 2 Rubidium salt. 3 Methylammonium salt. 4 Guanidinium salt. 5 L = HO(SnCl3OH2). 6 L = H2O. 7 L = OH (counter-ions: Ba2+, Cl). 8 L = OH (counter-ion: Ba2+). 9 L = F (counter-ion: NH4+). 10 L = OH (counter-ion: Na+). 11 L = Cl (counter-ion: H3O+). 12 counter-ion: H3O+. 13 counter-ion: NH4+.
Table A3. List of literature references (Ref.) and CSD reference codes (CSD Refcode) of selected crystallographically characterized compounds mentioned in this paper. The crystallographic data (in CIF format) can be accessed from the Cambridge Structural Database (CSD) using the respective CSD Refcode via https://www.ccdc.cam.ac.uk/structures/ (accessed on 1 February 2025 for availability check).
Table A3. List of literature references (Ref.) and CSD reference codes (CSD Refcode) of selected crystallographically characterized compounds mentioned in this paper. The crystallographic data (in CIF format) can be accessed from the Cambridge Structural Database (CSD) using the respective CSD Refcode via https://www.ccdc.cam.ac.uk/structures/ (accessed on 1 February 2025 for availability check).
EntryRef.CSD
Refcode
EntryRef.CSD
Refcode
EntryRef.CSD
Refcode
A3a[66]BOHNESA3h[122]TOSRUOA3o[177]PEGRUQ
A3b[67]ODEWEAA3i[146]EDTAFE01A3p[178]ZAYRIC
A3c[68]RIRRUGA3j[147]CERFEKA3q[182]TASSAH
A3d[69]NAGZAZA3k[171]TIRYAUA3r[183]TUPMEW
A3e[112]DANRUGA3l[171]TIRXUNA3s[186]IGICEG
A3f[117]XAVROFA3m[172]AMSOCUA3t[187]RODLOL
A3g[118]TEKSEJA3n[176]JOVXAV

References

  1. Stolba, F. Über die Bedeutung der Kieselflusssäure für die chemische Analyse. Z. Anal. Chem. 1864, 3, 298–313. [Google Scholar] [CrossRef]
  2. Davy, J. XVIII. An Account of some Experiments on different Combinations of Fluoric Acid. Philos. Trans. R. Soc. Lond. 1812, 102, 352–369. [Google Scholar] [CrossRef]
  3. Cohen, E. Physikalisch-chemische Studien am Zinn. Z. Phys. Chem. Stoechiom. Verwandtschafts. 1900, 35, 588–597. [Google Scholar] [CrossRef]
  4. Laubengayer, A.W.; Billings, O.B.; Newkirk, A.E. Chlorogermanic Acid and the Chlorogermanates. Properties and Crystal Structure of Cesium Hexachlorogermanate. J. Am. Chem. Soc. 1940, 62, 546–548. [Google Scholar] [CrossRef]
  5. Winkler, C. Germanium, Ge, ein neues, nichtmetallisches Element. Chem. Ber. 1886, 19, 210–211. [Google Scholar] [CrossRef]
  6. Sivaramakrishna, A.; Pete, S.; Mhaskar, C.M.; Ramann, H.; Ramanaiah, D.V.; Arbaaz, M.; Niyaz, M.; Janardan, S.; Suman, P. Role of hypercoordinated silicon(IV) complexes in activation of carbon–silicon bonds: An overview on utility in synthetic chemistry. Coord. Chem. Rev. 2023, 485, 215140. [Google Scholar] [CrossRef]
  7. Lemière, G.; Millanvois, A.; Ollivier, C.; Fensterbank, L. A Parisian Vision of the Chemistry of Hypercoordinated Silicon Derivatives. Chem. Rec. 2021, 21, 1119–1129. [Google Scholar] [CrossRef]
  8. Singh, G.; Kaur, G.; Singh, J. Progressions in hyper–coordinate silicon complexes. Inorg. Chem. Commun. 2018, 88, 11–20. [Google Scholar] [CrossRef]
  9. Peloquin, D.M.; Schmedake, T.A. Recent advances in hexacoordinate silicon with pyridine-containing ligands: Chemistry and emerging applications. Coord. Chem. Rev. 2016, 323, 107–119. [Google Scholar] [CrossRef]
  10. Wagler, J.; Böhme, U.; Kroke, E. Higher-Coordinated Molecular Silicon Compounds. In Functional Molecular Silicon Compounds I—Regular Oxidation States; Scheschkewitz, D., Ed.; Springer: Berlin/Heidelberg, Germany, 2013; Volume 115, pp. 29–105. [Google Scholar] [CrossRef]
  11. Puri, J.K.; Singh, R.; Kaur Chahal, V. Silatranes: A review on their synthesis, structure, reactivity and applications. Chem. Soc. Rev. 2011, 40, 1791–1840. [Google Scholar] [CrossRef]
  12. Kost, D.; Kalikhman, I. Hydrazide-based hypercoordinate silicon compounds. Adv. Organomet. Chem. 2004, 50, 1–106. [Google Scholar] [CrossRef]
  13. Chuit, C.; Corriu, R.J.P.; Reye, C.; Young, J.C. Reactivity of Penta- and Hexacoordinate Silicon Compounds and Their Role as Reaction Intermediates. Chem. Rev. 1993, 93, 1371–1448. [Google Scholar] [CrossRef]
  14. Sidorkin, V.F.; Pestunovich, V.A.; Voronkov, M.G. The Physical Chemistry of Silatranes. Russ. Chem. Rev. 1980, 49, 414–427. [Google Scholar] [CrossRef]
  15. Kadomtseva, A.V.; Mochalov, G.M.; Kuzina, O.V. Biologically Active Coordination Compounds of Germanium. Synthesis and Physicochemical Properties. Russ. J. Org. Chem. 2021, 57, 879–888. [Google Scholar] [CrossRef]
  16. Levason, W.; Reid, G.; Zhang, W. Coordination complexes of silicon and germanium halides with neutral ligands. Coord. Chem. Rev. 2011, 255, 1319–1341. [Google Scholar] [CrossRef]
  17. Singh, R.V.; Gupta, P.; Chaudhary, P.; Deshmukh, C.N. Coordination Compounds of Germanium(IV) Formed with Soft and Hard Donor Atoms: A Look into the Past and Present Work. Main Group Met. Chem. 2005, 28, 93–118. [Google Scholar] [CrossRef]
  18. Holloway, C.E.; Melnik, M. Germanium Organometallic Compounds: Classification and Analysis of Crystallographic and Structural Data. Main Group Met. Chem. 2002, 25, 185–266. [Google Scholar] [CrossRef]
  19. Lavigne, A.A.; Tancrede, J.M.; Pike, R.M. Coordination Compounds of Germanium. Coord. Chem. Rev. 1968, 3, 497–508. [Google Scholar] [CrossRef]
  20. Devi, J.; Kumar, B.; Taxak, B. Recent advancements of organotin(IV) complexes derived from hydrazone and thiosemicarbazone ligands as potential anticancer agents. Inorg. Chem. Commun. 2022, 139, 109208. [Google Scholar] [CrossRef]
  21. Srivastav, N.; Kumar, K.; Singh, R.; Kaur, V. Tricyclic tin(iv) cages: Synthetic aspects and intriguing features of stannatranes and pseudostannatranes. New J. Chem. 2020, 44, 3168–3184. [Google Scholar] [CrossRef]
  22. Manju; Mishra, N.; Kumar, D. Coordination Chemistry of Schiff Base Tin Complexes. Russ. J. Coord. Chem. 2014, 40, 343–357. [Google Scholar] [CrossRef]
  23. Pellerito, C.; Nagy, L.; Pellerito, L.; Szorcsik, A. Biological activity studies on organotin(IV)n+ complexes and parent compounds. J. Oranomet. Chem. 2006, 691, 1733–1747. [Google Scholar] [CrossRef]
  24. Pettinari, C. Organotin(IV) Derivatives of Imidazoles, Pyrazoles and Related Pyrazolyl and Imidazolyl Ligands. Main Group Met. Chem. 1999, 22, 661–692. [Google Scholar] [CrossRef]
  25. Holloway, C.E.; Melnik, M. Tin Coordination Compounds: Classification and Analysis of Crystallographic and Structural Data. Main Group Met. Chem. 1998, 21, 371–488. [Google Scholar] [CrossRef]
  26. Wagler, J.; Schley, M.; Gerlach, D.; Böhme, U.; Brendler, E.; Roewer, G. Surprising Insights in the Various Molecular Structures of Hypercoordinate Bis(oxinato)silicon Complexes. Z. Naturforsch. B 2005, 60, 1054–1064. [Google Scholar] [CrossRef]
  27. Wächtler, E.; Kämpfe, A.; Krupinski, K.; Gerlach, D.; Kroke, E.; Brendler, E.; Wagler, J. New Insights into Hexacoordinated Silicon Complexes with 8-Oxyquinolinato Ligands: 1,3-Shift of Si-Bound Hydrocarbyl Substituents and the Influence of Si-Bound Halides on the 8-Oxyquinolinate Coordination Features. Z. Naturforsch. B 2014, 69, 1402–1418. [Google Scholar] [CrossRef]
  28. Cai, Y.P. Germanium-8-Hydroxyquinoline hydroxides: Synthesis, structure and luminescence properties. Inorg. Chim. Acta 2019, 489, 211–216. [Google Scholar] [CrossRef]
  29. Schlemper, E.O. Crystal and molecular structure of dimethyltin bis(8-hydroxyquinolinate). Inorg. Chem. 1967, 6, 2012–2017. [Google Scholar] [CrossRef]
  30. Archer, S.J.; Koch, K.R.; Schmidt, S. Nitrogen donor chelates of bivalent tin chloride. Part II. The crystal structures of SnCl2·(2,2′-bipyridyl) and SnCl2·(1,10-phenanthroline). Inorg. Chim. Acta 1987, 126, 209–218. [Google Scholar] [CrossRef]
  31. Shi, D.; Hu, S. Structural studies of chloro(quinoline-8-olato)diethyltin(IV),[SnEt2Cl(C8H6NO)]. Chin. J. Struct. Chem. 1987, 6, 193–197. [Google Scholar]
  32. Schürmann, M.; Schmiedgen, R.; Huber, F.; Silvestri, A.; Ruisi, G.; Barbieri Paulsen, A.; Barbieri, R. Mono-aryltin(IV) and mono-benzyltin(IV) complexes with pyridine-2-carboxylic acid and 8-hydroxyquinoline. X-ray structure of p-chloro-phenyl-tris(8-quinolinato)tin(IV)·2CHCl3. J. Organomet. Chem. 1999, 584, 103–117. [Google Scholar] [CrossRef]
  33. Hagemann, M.; Mix, A.; Berger, R.J.F.; Pape, T.; Mitzel, N.W. Strong Intramolecular Si−N Interactions in the Chlorosilanes Cl3−nHnSiOCH2CH2NMe2 (n = 1−3). Inorg. Chem. 2008, 47, 10554–10564. [Google Scholar] [CrossRef]
  34. Khrustalev, V.N.; Portnyagin, I.A.; Borisova, I.V.; Zemlyansky, N.N.; Ustynyuk, Y.A.; Antipin, M.Y.; Nechaev, M.S. Donor-Stabilized Germyl Cations. Stable Pentacoordinate Germanium Chloride [PhGe(OCH2CH2NMe2)2][Cl]. Organometallics 2006, 25, 2501–2504. [Google Scholar] [CrossRef]
  35. Portnyagin, I.A.; Nechaev, M.S.; Khrustalev, V.N.; Zemlyansky, N.N.; Borisova, I.V.; Antipin, M.Y.; Ustynyuk, Y.A.; Lunin, V.V. An Unusual Reaction of (β-Dimethylaminoethoxy)triethyltin with Phenyltin Trichloride. The First X-ray Structural Evidence of the Existence of Complexes R2SnXY·R2SnXY (R = Alkyl, Aryl; X, Y = Hal, OR, X ≠ Y) Both as Unsymmetrical Adducts [R2SnX2·R2SnY2] and Symmetrical Dimers [R2SnXY]2. Eur. J. Inorg. Chem. 2006, 2006, 4271–4277. [Google Scholar] [CrossRef]
  36. Verchère, A.; Mishra, S.; Jeanneau, E.; Guillon, H.; Decams, J.-M.; Daniele, S. Heteroleptic Tin(IV) Aminoalkoxides and Aminofluoroalkoxides as MOCVD Precursors for Undoped and F-Doped SnO2 Thin Films. Inorg. Chem. 2020, 59, 7167–7180. [Google Scholar] [CrossRef]
  37. Seidel, A.; Gericke, R.; Kutzner, B.; Wagler, J. Lewis Acid-Base Adducts of α-Amino Acid-Derived Silaheterocycles and N-Methylimidazole. Molecules 2023, 28, 7816. [Google Scholar] [CrossRef]
  38. Kuß, S.; Brendler, E.; Wagler, J. Molecular Structures of the Pyridine-2-olates PhE(pyO)3 (E = Si, Ge, Sn)—[4+3]-Coordination at Si, Ge vs. Heptacoordination at Sn. Crystals 2022, 12, 1802. [Google Scholar] [CrossRef]
  39. Kowalke, J.; Brendler, E.; Wagler, J. Valinate and SiMe2—An interesting couple in pentacoordinate Si-complexes: Templated generation of the dipeptide val-val and formation of an organosilicon-ammonia-adduct. J. Organomet. Chem. 2021, 956, 122126. [Google Scholar] [CrossRef]
  40. Kämpfe, A.; Brendler, E.; Kroke, E.; Wagler, J. Tp*Cu(I)–CN–SiL2–NC–Cu(I)Tp*—A hexacoordinate Si-complex as connector for redox active metals via π-conjugated ligands. Dalton Trans. 2015, 44, 4744–4750. [Google Scholar] [CrossRef] [PubMed]
  41. Kämpfe, A.; Brendler, E.; Kroke, E.; Wagler, J. 2-Acylpyrroles as Mono-anionic O,N-Chelating Ligands in Silicon Coordination Chemistry. Chem. Eur. J. 2014, 20, 9409–9418. [Google Scholar] [CrossRef] [PubMed]
  42. Schöne, D.; Gerlach, D.; Wiltzsch, C.; Brendler, E.; Heine, T.; Kroke, E.; Wagler, J. A Distorted Trigonal Antiprismatic Cationic Silicon Complex with Ureato Ligands: Syntheses, Crystal Structures and Solid State 29Si NMR Properties. Eur. J. Inorg. Chem. 2010, 2010, 461–467. [Google Scholar] [CrossRef]
  43. Brendler, E.; Wächtler, E.; Wagler, J. Hypercoordinate Silacycloalkanes: Step-by-Step Tuning of N→Si Interactions. Organometallics 2009, 28, 5459–5465. [Google Scholar] [CrossRef]
  44. Wagler, J.; Böhme, U.; Brendler, E.; Roewer, G. Equilibrium between Tetra-, Penta-, and Hexacoordinate Imine and Enamine Chelates of Silicon: Crystal Structure and Variable-Temperature NMR. Organometallics 2005, 24, 1348–1350. [Google Scholar] [CrossRef]
  45. Nath, M.; Goyal, S. Triorganosilicon(IV) Derivatives of Amino Acids. Synth. React. Inorg. Met.-Org. Nano-Met. Chem. 2002, 32, 1205–1221. [Google Scholar] [CrossRef]
  46. Nath, M.; Goyal, S. Dimethylsilicon(IV) Derivatives of Amino Acids. Phosphorus Sulfur Silicon Relat. Elem. 2002, 177, 841–851. [Google Scholar] [CrossRef]
  47. Kelusky, E.C.; Fyfe, C.A. Molecular motions of alkoxysilanes immobilized on silica surfaces: A deuterium NMR study. J. Am. Chem. Soc. 1986, 108, 1746–1749. [Google Scholar] [CrossRef]
  48. Cota, S.; Beyer, M.; Bertermann, R.; Burschka, C.; Götz, K.; Kaupp, M.; Tacke, R. Neutral Penta- and Hexacoordinate Silicon(IV) Complexes Containing Two Bidentate Ligands Derived from the α-Amino Acids (S)-Alanine, (S)-Phenylalanine, and (S)-tert-Leucine. Chem. Eur. J. 2010, 16, 6582–6589. [Google Scholar] [CrossRef] [PubMed]
  49. Tacke, R.; Bertermann, R.; Burschka, C.; Dragota, S.; Penka, M.; Richter, I. Spirocyclic Zwitterionic λ5Si-Silicates with Two Bidentate Ligands Derived from α-Amino Acids or α-Hydroxycarboxylic Acids: Synthesis, Structure, and Stereodynamics. J. Am. Chem. Soc. 2004, 126, 14493–14505. [Google Scholar] [CrossRef] [PubMed]
  50. Dragota, S.; Bertermann, R.; Burschka, C.; Penka, M.; Tacke, R. Diastereo- and Enantiomerically Pure Zwitterionic Spirocyclic λ5Si-[(Ammonio)methyl]silicates with an SiO2N2C Skeleton Containing Two Bidentate Chelate Ligands Derived from α-Amino Acids. Organometallics 2005, 24, 5560–5568. [Google Scholar] [CrossRef]
  51. Hattori, T.; Yamamoto, H. Synthesis of Silacyclic Dipeptides: Peptide Elongation at Both N- and C-Termini of Dipeptide. J. Am. Chem. Soc. 2022, 144, 1758–1765. [Google Scholar] [CrossRef] [PubMed]
  52. Soto-Cairoli, B.; Kock, I.; de Pomar, J.J.; Yang, G.; Guzmán, J.M.; González, J.R.; Antomattei, A.; Soderquist, J.A. Selective Triisopropylsilylation of α-Amino Acids: Protection Without Racemization. Heterocycles 2010, 80, 409–426. [Google Scholar] [CrossRef]
  53. Denmark, S.E. The Interplay of Invention, Discovery, Development, and Application in Organic Synthetic Methodology: A Case Study. J. Org. Chem. 2009, 74, 2915–2927. [Google Scholar] [CrossRef] [PubMed]
  54. Lavayssiere, H.; Dousse, G.; Satge, J. Heterocycles du germanium, du phosphore (III), de l’arsenic (III) et du soufre a ligand aminoacide ou hydroxyacide. J. Organomet. Chem. 1977, 137, C37–C42. [Google Scholar] [CrossRef]
  55. Lavayssiere, H.; Dousse, G.; Satge, J. Oxazolidones et dioxolonnes du germanium (IV) et (II) du phosphore (III), de l’arsenic (III) et du soufre. Phosphorus Sulfur Silicon Relat. Elem. 1990, 53, 411–422. [Google Scholar] [CrossRef]
  56. Ignatyev, I.S.; Kondratenko, Y.A.; Lezov, D.V.; Ugolkov, V.L.; Kochina, T.A. Hexacoordinate germanium compounds with BIS-TRIS and amino acid ligands. Mendeleev Commun. 2023, 33, 601–604. [Google Scholar] [CrossRef]
  57. Ho, B.Y.K.; Molloy, K.C.; Zuckerman, J.J.; Reidinger, F.; Zubieta, J.A. The crystal structure and variable temperature 119mSn Mössbauer study of trimethyltin glycinate, a one-dimensional, amino-bridged polymer. J. Organomet. Chem. 1980, 187, 213226. [Google Scholar] [CrossRef]
  58. Ma, C.; Zhang, J.; Zhang, R. Syntheses and Characterizations of Diorganotin(IV) Complexes with L-Cysteine: Crystal Structure of [(CH3)2Sn(l-C3H5NO2S)·H2O]. Heteroat. Chem. 2003, 14, 636–641. [Google Scholar] [CrossRef]
  59. Domazetis, G.; MacKay, M.F.; Magee, R.J.; James, B.D. Interaction of Organotin Compounds with Biological Ligands and the Molecular Structure of Ethyl L-Cysteinato S,N.(chlorodimethyl)stahyl)stannate. Inorg. Chim. Acta 1979, 34, L247–L248. [Google Scholar] [CrossRef]
  60. Calogero, S.; Valle, G.; Cusack, P.A.; Smith, P.J.; Donaldson, J.D. The X-ray Crystal Structure of Dichloro-S,S’-Bis(O-methylcysteinato)tin(IV): The First Determination of an Inorganic Tin-Amino Acid Derivative. Inorg. Chim. Acta 1982, 67, 95–98. [Google Scholar] [CrossRef]
  61. Anderson, J.E.; Sawtelle, S.M.; Thompson, J.S.; Kretchmar Nguyen, S.A.; Calabrese, J. Ligand Addition Reactions and the Electron-Transfer Properties of SnCl2·2H2O and SnCl4·5H2O. Molecular Structure of Bis(ethylcysteinato)tin(II). Inorg. Chem. 1992, 31, 2778–2785. [Google Scholar] [CrossRef]
  62. Djurdjevic, P.; Djokic, D. Protein Interactions with Bivalent Tin. 1. Hydrolysis and Complexation of Tin(II) Ion with Glycine. J. Inorg. Biochem. 1996, 62, 17–29. [Google Scholar] [CrossRef] [PubMed]
  63. Nath, M.; Yadaf, R. Studies of Organotin(IV) derivatives of DL-Methionine and L-Asparagine. Bull. Chem. Soc. Jpn. 1998, 71, 1355–1362. [Google Scholar] [CrossRef]
  64. Cashion, J.D.; Domazetis, G.; James, B.D. Mössbauer spectra of organotin amino-acid and glutathione derivatives. J. Organomet. Chem. 1980, 185, 433–441. [Google Scholar] [CrossRef]
  65. Nath, M.; Jairath, R.; Eng, G.; Song, X.; Kumar, A. Synthesis, spectral characterization and biological studies of some organotin(IV) complexes of l-proline, trans-hydroxy-l-proline and l-glutamine. Spectrochim. Acta A 2005, 62, 1179–1187. [Google Scholar] [CrossRef] [PubMed]
  66. Gasque, L.; Verhoeven, M.A.; Bernès, S.; Barrios, F.; Haasnoot, J.G.; Reedijk, J. The Added Value of Solid-State Pb NMR Spectroscopy to Understand the 3D Structures of Pb Amino Acid Complexes. Eur. J. Inorg. Chem. 2008, 2008, 4395–4403. [Google Scholar] [CrossRef]
  67. Marandi, F.; Shahbakhsh, N. Synthesis and Crystal Structure of [Pb(phe)2]n: A 2D Coordination Polymer of Lead(II) Containing Phenylalanine. Z. Anorg. Allg. Chem. 2007, 633, 1137–1139. [Google Scholar] [CrossRef]
  68. Marandi, F.; Shahbakhsh, N. Lead(II) complexes of proline. J. Coord. Chem. 2007, 60, 2589–2595. [Google Scholar] [CrossRef]
  69. Nair, R.M.; Dhanya, V.S.; Suma, S.; Sudhadevi Antharjanam, P.K.; Sudarsanakumar, M.R. A novel 2D ladder shaped metal–organic framework based on lead-aspartate system with hydrophobic channels. Main Group Chem. 2015, 14, 91–103. [Google Scholar] [CrossRef]
  70. Preut, H.; Vornefeld, M.; Huber, F. Dimethylgermanium Glycyl-l-methionate. Acta Crystallogr. C 1989, 45, 1504–1506. [Google Scholar] [CrossRef]
  71. Vornefeld, M.; Huber, F.; Preut, H.; Brunner, H. Synthesis and spectroscopic characterization of dimethylgermanium derivatives of dipeptides, crystal structure of dimethylgermanium glycylglycinate and in vivo effects of dimethylgermanium glycylglycinate against murine leukemia P388. Appl. Organomet. Chem. 1989, 3, 177–182. [Google Scholar] [CrossRef]
  72. Giuffrida, S.; Fontana, A.; Magio, F.; Duca, D. Synthesis, characterization and conformational analysis of chloro-bis(glycylglycinate)germanium(IV) chloride. New J. Chem. 2011, 35, 807–819. [Google Scholar] [CrossRef]
  73. Preut, H.; Mundus, B.; Huber, F.; Barbieri, R. Dimethyltin Glyeylmethionate. Acta Crystallogr. C 1986, 42, 536–538. [Google Scholar] [CrossRef]
  74. Nath, M.; Singh, H.; Eng, G.; Song, X. New di- and triorganotin(IV) derivatives of tyrosinylphenylalanine as models for metal–protein interactions: Synthesis and structural characterization. Crystal structure of Me2Sn(Tyr-Phe) · MeOH. J. Organomet. Chem. 2008, 693, 2541–2550. [Google Scholar] [CrossRef]
  75. Girasolo, M.A.; Pizzino, T.; Mansueto, C.; Valle, G.; Stocco, G.C. Spectroscopic Characterization and Biological Activity of L-Methionyl-L Histidinato Complexes of R2Sn(IV) Ions (R = Me, nBu, Ph) and X-ray Structure of Me2SnMetHis·0.5MeOH. Appl. Organomet. Chem. 2000, 14, 197–211. [Google Scholar] [CrossRef]
  76. Stocco, G.; Gulì, G.; Valle, G. Structures of (L-Methionyl-L-methioninato)dimethyltin(IV) and (L-Alanyl-L-histidinato)dimethyltin(IV). A Class of Potential Antitumour Agents. Acta Crystallogr. C 1992, 48, 2116–2120. [Google Scholar] [CrossRef]
  77. Girasolo, M.A.; Pellerito, L.; Stocco, G.C.; Valle, G. Crystal and molecular structure of dimethyl(L-tryptophyl-L-alaninato)tin(IV)-methanol (l/l) and Mössbauer spectroscopy studies of lattice dynamics of diorganotin(IV) dipeptide complexes. J. Chem. Soc. Dalton Trans. 1996, 25, 1195–1201. [Google Scholar] [CrossRef]
  78. Preut, H.; Vornefeld, M.; Huber, F. Bis[diethyl(glycylhistidinato)tinl-Methanol (1/1). Acta Crystallogr. C 1991, 47, 264–267. [Google Scholar] [CrossRef]
  79. Vornefeld, M.; Huber, F.; Preut, H.; Ruisi, G.; Barbieri, R. Synthesis and spectroscopic characterization of diethyltin(1V) derivatives of dipeptides: Crystal and molecular structure of diethyltin glycyltyrosinate. Appl. Organomet. Chem. 1992, 6, 75–82. [Google Scholar] [CrossRef]
  80. Math, M.; Singh, H.; Kumar, P.; Kumar, A.; Song, X.; Eng, G. Organotin(IV) tryptophanylglycinates: Potential non-steroidal anti-inflammatory agents; crystal structure of dibutyltin(IV) tryptophanylglycinate. Appl. Organomet. Chem. 2009, 23, 347–358. [Google Scholar] [CrossRef]
  81. Katsoulakou, E.; Tiliakos, M.; Papaefstathiou, G.; Terzis, A.; Raptopoulou, C.; Geromichalos, G.; Papazisis, K.; Papi, R.; Pantazaki, A.; Kyriakidis, D.; et al. Diorganotin(IV) complexes of dipeptides containing the a-aminoisobutyryl residue (Aib): Preparation, structural characterization, antibacterial and antiproliferative activities of [(n-Bu)2Sn(H–1L)] (LH = H-Aib-L-Leu-OH, H-Aib-L-Ala-OH). J. Inorg. Biochem. 2008, 102, 1397–1405. [Google Scholar] [CrossRef]
  82. Mundus-Glowacki, B.; Huber, F.; Preut, H.; Guisi, G.; Barbieri, R. Synthesis and spectroscopic characterization of dimethyl-, di-n-butyl-, di-t-butyl- and diphenyl-tin(lV) derivatives of dipeptides: Crystal and molecular structure of di-nbutyltin(IV) glycylvalinate. Appl. Organomet. Chem. 1992, 6, 83–94. [Google Scholar] [CrossRef]
  83. Huber, F.; Vornefeld, M.; Preut, H.; von Angerer, E.; Ruisi, G. Synthesis and spectroscopic characterization of dicyclohexyltin derivatives of dipeptides, and in vitro effects against MDA-MB 231 breast cancer cells: Crystal structures of dicyclohexyltin glycylglycinate and gIycylalaninate. Appl. Organomet. Chem. 1992, 6, 597–606. [Google Scholar] [CrossRef]
  84. Preut, H.; Mundus, B.; Huber, F.; Barbieri, R. Di-tert-butyl(glycylglycinato)tin(IV) Monohydrate. Acta Crystallogr. C 1989, 45, 728–730. [Google Scholar] [CrossRef]
  85. Huber, F.; Haupt, H.J.; Preut, H.; Barbieri, R.; Lo Giudice, M.T. Darstellung, Kristall- und Molekülstruktur von Diphenylzinnglycylglycinat, (C6H6)2SnC4H6N2O3. Z. Anorg. Allg. Chem. 1977, 432, 51–57. [Google Scholar] [CrossRef]
  86. Ho, B.Y.K.; Zuckerman, J.J. Trialkyltin derivatives of amino acids and dipeptides. Inorg. Chem. 1973, 12, 1552–1561. [Google Scholar] [CrossRef]
  87. Ho, B.Y.K.; Zuckerman, J.J. Structural variations in trimethyl- and tricyclohexyltin derivatives of amino acids. Preference for association by nitrogen bridging. Inorg. Nucl. Chem. Lett. 1973, 9, 849–855. [Google Scholar] [CrossRef]
  88. Pellerito, L.; Lo Giudice, M.T.; Ruisi, G.; Bertazzi, N.; Barbieri, R.; Huber, F. Complexes of organometallic compounds. XLIV. Synthesis and Mössbauer spectroscopic studies of glylcylglycinato-O,N,N(2−)diorganotin (IV) complexes. Inorg. Chim. Acta. 1976, 17, L21–L22. [Google Scholar] [CrossRef]
  89. Gerlach, D.; Ehlers, A.W.; Lammertsma, K.; Wagler, J. Silicon(IV) Chelates of an (ONN’)-Tridentate Pyrrole-2-Carbaldimine Ligand: Syntheses, Structures and UV/Vis Properties. Z. Naturforsch. B 2009, 64, 1571–1579. [Google Scholar] [CrossRef]
  90. Brendler, E.; Wächtler, E.; Heine, T.; Zhechkov, L.; Langer, T.; Pöttgen, R.; Hill, A.F.; Wagler, J. Stannylene or Metallastanna(IV)ocane: A Matter of Formalism. Angew. Chem. Int. Ed. 2011, 50, 4696–4700. [Google Scholar] [CrossRef]
  91. Hattori, T.; Yamamoto, H. Peptide Bond Formation Between Unprotected Amino Acids: Convergent Synthesis of Oligopeptides. J. Am. Chem. Soc. 2024, 146, 25738–25744. [Google Scholar] [CrossRef]
  92. Hattori, T.; Yamamoto, H. Trimethylaluminum-mediated one-pot peptide elongation. Chem. Sci. 2023, 14, 5795–5801. [Google Scholar] [CrossRef] [PubMed]
  93. Dai, J.; Zhang, J.; Wu, Y. Chin. X-ray diffraction study on crystals of heteroazotricyclic compounds Ⅱ, crystal and molecular structure of 1-chloropropyl-2,8,9-trioxa-5-aza-1-silicon tricyclo [3,3,3,01.5]undecane-3-one (In Chinese). J. Struct. Chem. 1983, 2, 107. [Google Scholar]
  94. Dai, J.; Zhang, J.; Wu, Y.; Wu, G. Chin. X-ray study on azotricyclic compounds Ⅲ. Structural determination of 1-chloromethyl-2,8,9-trioxa-5-aza-1-silicon tricyclo [3,3,3,01,5]decane-3-one (In Chinese). J. Struct. Chem. 1983, 2, 207. [Google Scholar]
  95. Párkányi, L.; Hencsei, P.; Popowski, E. The Crystal Structures of m-Trifluoromethylphenylsilatranone and p-Fluorophenylsilatranone. J. Organomet. Chem. 1980, 197, 275–283. [Google Scholar] [CrossRef]
  96. Párkányi, L.; Hencsei, P.; Csonka, G.; Kovács, I. The molecular structure of 1-phenylsilatranone. J. Organomet. Chem. 1987, 329, 305–311. [Google Scholar] [CrossRef]
  97. Fülöp, V.; Kálmán, A.; Hencsei, P.; Csonka, G.; Kovács, I. Structure of l-Methylsilatranone, CH3Si(OCOCH2)(OCH2CH2)2N. Acta Crystallogr. C 1988, 44, 720723. [Google Scholar] [CrossRef]
  98. Kemme, A.; Bleidelis, J.; Lapsina, A.; Fleisher, M.; Zelcans, G.; Lukevics, E. Molecular and crystal structure of 1-methylsilatrane-3,7-dione. Latv. PSR Zinat. Akad. Vestis Kim. Ser. 1985, 2, 242–245. [Google Scholar]
  99. Lyssenko, K.A.; Korlyukov, A.A.; Antipin, M.Y.; Knyazev, S.P.; Kirin, V.N.; Alexeev, N.V.; Chernyshev, E.A. The nature of the intramolecular transannular Si⋅⋅⋅N interaction in crystalline 1-methylsilatrane, as found from X-ray diffraction data. Mendeleev. Commun. 2000, 10, 88–90. [Google Scholar] [CrossRef]
  100. Párkányi, L.; Simon, K.; Nagy, J. Crystal and molecular structure of [beta]-1-phenylsilatrane, C12H17O3NSi. Acta Crystallogr. B 1974, 30, 2328–2332. [Google Scholar] [CrossRef]
  101. Kupče, E.; Liepiņš, E.; Lapsina, A.; Zelchan, G.; Lukevics, E. Synthesis and 1H, 13C, 15N, 29Si NMR spectra of sil- and germ-atranones. J. Organomet. Chem. 1983, 251, 15–29. [Google Scholar] [CrossRef]
  102. Ilyukhin, A.B.; Shkol’nikova, L.M.; Seifullina, I.I.; Batalova, T.P.; Dyatlova, N.M. Preparation, structure and properties of germanium(IV) with nitrilotriacetic acid [Ge(Nta)(H2O)(OH)]·2H2O. Koord. Khim. 1991, 17, 795–800. [Google Scholar]
  103. Martsinko, E.E.; Seifullina, I.I.; Minacheva, L.K.; Syvak, T.A.; Sergienko, V.S. Synthesis, properties, crystal, and molecular structure of potassium (nitrilotriacetato)dihydroxogermanate(IV) hydrate K[Ge(Nta)(OH)2]·H2O. Zh. Neorg. Khim. 2009, 54, 1428–1434. [Google Scholar]
  104. Mafra, L.; Almeida Paz, F.A.; Shi, F.-N.; Rocha, J.; Trindade, T.; Fernandez, C.; Makal, A.; Wozniak, K.; Klinowski, J. X-ray Diffraction and Solid-State NMR Studies of a Germanium Binuclear Complex. Chem. Eur. J. 2006, 12, 363–375. [Google Scholar] [CrossRef] [PubMed]
  105. Langer, H.G. Solid complexes with tetravalent metal ions and ethylenediamime tetra-acetic acid (EDTA). J. Inorg. Nucl. Chem. 1964, 26, 59–72. [Google Scholar] [CrossRef]
  106. Mizuta, T.; Yoshida, T.; Miyoshi, K. A Novel Structure of Spontaneously Resolved Germanium(IV) Complex with Ethylenediaminetriacetatemonoacetic Acid (Hedta). Inorg. Chim. Acta 1989, 165, 65–71. [Google Scholar] [CrossRef]
  107. Martsinko, E.E.; Seifullina, I.I.; Minacheva, L.K.; Shchur, T.A.; Sergienko, V.S. Diphenylguanidinium (ethylendiaminetetraacetato)hydroxogermanate hydrate (HDphg)[Ge(OH)(Edta)]·H2O: Synthesis, physicochemical characterization, and crystal structure. Zh. Neorg. Khim. 2007, 52, 2023–2029. [Google Scholar]
  108. Zhu, J.; Huang, T.-S.; Huang, Y.-Q.; Hong, M.-S.; Hu, S.-Z. Chin.Synthesis and structure of aminopolycarboxylic acid germanium(IV) chelates [Ge(OH)(Hpdta)] and [Ge(OH)(H2dtpa)]·H2O (In Chinese). J. Struct. Chem. 1995, 14, 191. [Google Scholar]
  109. Chen, M.-D.; Zhu, J.; Hu, S.-Z. Synthesis, crystal structure and quantum chemical calculations of [Ge(OH)(Hcdta)]·H2O (H4cdta = trans-cyclohexane-1,2-diaminetetraacetic acid). Main Group Met. Chem. 1999, 22, 105–109. [Google Scholar] [CrossRef]
  110. Martsinko, E.E.; Seifullina, I.I.; Minacheva, L.K.; Shchur, T.A.; Sergienko, V.S. Synthesis, properties, and crystal structure of {N-(2-hydroxyethyl)ethylenediaminetriacetato}hydroxogermanium(IV) sesquihydrate. Zh. Neorg. Khim. 2007, 52, 1621–1628. [Google Scholar]
  111. Sergienko, V.S.; Martsinko, E.E.; Seifullina, I.I.; Churakov, A.V.; Chebanenko, E.A. Synthesis and the Crystal and Molecular Structure of the Germanium(IV) Complex with Propylene-1,3-diaminetetraacetic Acid [Ge(Pdta)]. Crystallogr. Rep. 2015, 60, 677–681. [Google Scholar] [CrossRef]
  112. Sergienko, V.S.; Aleksandrov, G.G.; Seifullina, I.I.; Martsinko, E.E. Synthesis and the crystal and molecular structures of a germanium(IV)-copper(II) heteronuclear Diethylenetriaminepentaacetate Complex, [Cu(μ-HDtpa)2{Ge(OH)}2]·12H2O. Crystallogr. Rep. 2004, 49, 788–791. [Google Scholar] [CrossRef]
  113. Martsinko, E.E.; Smola, S.S.; Minacheva, L.K.; Seifullina, I.I.; Sergienko, V.S. Neodymium(III) triaquatrihydroxo(2-hydroxy-1,3-diaminopropanetetraacetato)germanium(IV) hydrate [Ge(OH)(μ-HHpdta)(μ-OH)Nd(OH)(H2O)3]·H2O: Synthesis and crystal and molecular structure. Russ. J. Inorg. Chem. 2009, 54, 1041–1048. [Google Scholar] [CrossRef]
  114. Martsinko, E.E.; Minacheva, L.K.; Smola, S.S.; Seifullina, I.I.; Sergienko, V.S. Synthesis and characterization of heteronuclear germanium(IV) lanthanide 1,3-diamino-2-propanoltetraacetates: Crystal and molecular structure of the [Ge(OH)(μ-Hpdta)(μ-OH)Ln(H2O)3]·2H2O complexes (Ln = Tb, Yb). Russ. J. Inorg. Chem. 2011, 56, 1034–1042. [Google Scholar] [CrossRef]
  115. Martsinko, E.E.; Seifullina, I.I.; Chebanenko, E.A.; Dyakonenko, V.V.; Shishkina, S.V.; Odessa, I.I. СИНТЕЗ І СТРУКТУРА БІЯДЕРНОГО КОМПЛЕКСУ ГЕРМАНІЮ(IV)-ТУЛІЮ(III) З 1,3-ДІАМІНО-2-ГІДРОКСИПРОПАН-N,N,N’,N’-ТЕТРАОЦТОВОЮ КИСЛОТОЮ. Mechnikov Natl. Univ. Bull.—Chem. 2020, 25, 24–31. [Google Scholar] [CrossRef]
  116. Martsinko, E.E.; Minacheva, L.K.; Sergienko, V.S.; Chebanenko, E.A.; Seifullina, I.I. A new binuclear germanium(IV) and copper(II) complex with 1,3-diamino-2-propanoltetraacetic acid: Crystal and molecular structure of [(H2O)(OH)Ge(μ-Hpdta)Cu(H2O)]·3H2O. Russ. J. Inorg. Chem. 2010, 55, 1874–1881. [Google Scholar] [CrossRef]
  117. Seifullina, I.I.; Martsinko, E.E.; Chebanenko, E.A.; Dyakonenko, V.V.; Shishkina, S.V.; Pesaroglo, A.G. Synthesis, molecular and crystal structure of a heterometallic Cu(II)-Ge(IV) complex with 2-hydroxy-1,3-diaminopropane-n,n,n′,n′-tetracetic acid and 2,2′′-bipyridine. Vopr. Khim. Khim. Tekhnol. 2020, 159–164. [Google Scholar] [CrossRef]
  118. Seifullina, I.I.; Minacheva, L.K.; Martsinko, E.E.; Sergienko, V.S.; Churakov, A.V. Tetrameric complexes of germanium(IV) and cobalt(II), Nickel(II), or Zinc(II) with 1,3-Diamino-2-propanol-tetraacertic acid: Crystal and molecular structures of [(OH)2Ge2(μ-Hpdta)2Zn2(H2O)4]·12H2O. Russ. J. Inorg. Chem. 2012, 57, 1545–1552. [Google Scholar] [CrossRef]
  119. Hussain, S.; Ali, S.; Shahzadi, S.; Tahir, M.N.; Shahid, M.; Munawar, K.S.; Abbas, S.M. Synthesis, spectroscopy, single crystal XRD and biological studies of multinuclear organotin dicarboxylates. Polyhedron 2016, 117, 64–72. [Google Scholar] [CrossRef]
  120. Mancilla, T.; Carrillo, L.; Zamudio Rivera, L.S.; Camacho, C.; De Vos, D.; Kiss, R.; Darro, F.; Mahieu, B.; Tiekink, E.R.T.; Rahier, H.; et al. Di-n-butyltin(IV) derivatives of bis(carboxymethyl)benzylamines: Synthesis, NMR and X-ray structure characterization and in vitro antitumour properties. Appl. Organomet. Chem. 2001, 15, 593–603. [Google Scholar] [CrossRef]
  121. Lee, F.L.; Gabe, E.J.; Khoo, L.E.; Leong, W.H.; Eng, G.; Smith, F.E. Synthesis and structural studies of diorganotin iminodiacetates. Inorg. Chim. Acta 1989, 166, 257–261. [Google Scholar] [CrossRef]
  122. Aizawa, S.; Natsume, T.; Hatano, K.; Funahashi, S. Complexation equilibria and structures of dimethyltin (IV) complexes with N-methyliminodiacetate, pyridine-2,6-dicarboxylate, ethylenediamine-N,N′-diacetate and ethylenediamine-N,N,N′,N′-tetraacetate. Inorg. Chim. Acta 1996, 248, 215–224. [Google Scholar] [CrossRef]
  123. Di Nicola, C.; Galindo, A.; Hanna, J.V.; Marcetti, F.; Pettinari, C.; Pettinari, R.; Rivarola, E.; Skelton, B.W.; White, A.H. Synthesis and Spectroscopic and X-ray Structural Characterization of R2SnIV−Oxydiacetate and −Iminodiacetate Complexes. Inorg. Chem. 2005, 44, 3094–3102. [Google Scholar] [CrossRef] [PubMed]
  124. Hussain, S.; Ali, S.; Shahzadi, S.; Rizzoli, C.; Shahid, M. Diorganotin(IV) Complexes with Monohydrate Disodium Salt of Iminodiacetic Acid: Synthesis, Characterization, Crystal Structure and Biological Activities. J. Chin. Chem. Soc. 2015, 62, 793–802. [Google Scholar] [CrossRef]
  125. Henling, L.M.; Marsh, R.E. Some more space-group corrections. Acta Crystallogr. C 2014, 70, 834–836. [Google Scholar] [CrossRef] [PubMed]
  126. Fang, X.-N.; Luo, Q.-Y.; Zeng, X.-R.; Sui, Y.; Li, X.-F. Bis{μ-N-[bis(hydroxymethyl)(oxidomethyl)methyl]glycinato}bis[dimethyltin(IV)] dioxane hemisolvate. Acta Crystallogr. E 2005, 61, m2604–m2606. [Google Scholar] [CrossRef]
  127. Meriem, A.; Willem, R.; Meunier-Piret, J.; Gielen, M. Diorganotin(IV) derivatives of N-(2-hydroxyethyl)- and N-(carbamoylmethyl)iminodiacetic acid: Synthesis, spectroscopic characterization, x-ray structure analysis and in vitro antitumor activity. Main Group Met. Chem. 1989, 12, 187–198. [Google Scholar]
  128. Arjmand, F.; Yousuf, I.; Zaidib, Y.; Toupetc, L. Crystal structure determination, spectroscopic characterization and biological profile of a tailored ionic molecular entity, Sn(IV) iminodiacetic acid–piperazinediium conjugate: In Vitro DNA/RNA binding studies, Topo I inhibition activity, cytotoxic and systemic toxicity studies. RSC Adv. 2015, 5, 16250–16264. [Google Scholar] [CrossRef]
  129. Clark, A.M.; Rickard, C.E.F.; Roper, W.R.; Woodman, T.J.; Wright, L.J. Stepwise Conversion of an Osmium Trimethylstannyl Complex to a Triiodostannyl Complex and Nucleophilic Substitution Reactions at the Tin−Iodine Bonds. Organometallics 2000, 19, 1766–1774. [Google Scholar] [CrossRef]
  130. Dakternieks, D.; Dyson, G.; Jurkschat, K.; Tozer, R.; Tiekink, E.R.T. Utilization of hypervalently activated organotin compounds in synthesis. Preparation and reactions of Me2N(CH2)3SnPh3. J. Organomet. Chem. 1993, 458, 29–38. [Google Scholar] [CrossRef]
  131. Ilyukhin, A.B.; Davidovich, R.L.; Logvinova, V.B. Nitrilotriacetate complexes with acid and normal ligands: Crystal structure of Cs[SnIV(Nta)(HNta)]·H2O and K2[Bi(Nta)(HNta)]·H2O. Zh. Neorg. Khim. 1999, 44, 1931–1934. [Google Scholar]
  132. Logvinova, V.B.; Davidovich, R.L.; Ilyukhin, A.B.; Sergienko, V.S.; Tkachev, V.V.; Atovmyan, L.O. Crystal structure of Cat2[M(Nta)2]·2H2O (Cat = K+, or Rb+ and M = Sn or 0.5Sn+0.5Zr). Russ. J. Coord. Chem. 1998, 24, 530–532. [Google Scholar]
  133. Haussühl, E.; Haussühl, S.; Tillmanns, E. Structural and physical properties of bis(guanidinium) X bis(nitrilotriacetate) hydrate, (C[NH2]3)2X(N[CH2COO]3)2·H2O (X = Zr, Sn, Hf) and bis(methylammonium) X bis(nitrilotriacetate dihydrate, (CH3NH3)2X(N[CH2COO]3)2·2 H2O (X = Sn, Zr). Z. Kristallogr.—Cryst. Mater. 2009, 218, 626–635. [Google Scholar] [CrossRef]
  134. Martsinko, E.E.; Ilyukhin, A.B.; Seifullina, I.I.; Chebanenko, E.A.; Sergienko, V.S. Synthesis, properties, and crystal structure of the tin(IV) complex with N-(2-hydroxyethyl)ethylenediaminetriacetic acid [Sn(μ-Hedtra)(μ-OH)SnCl3(H2O)]·3H2O. Russ. J. Coord. Chem. 2013, 39, 505–509. [Google Scholar] [CrossRef]
  135. Van Remoortere, F.P.; Flynn, J.J.; Boer, F.P. Crystal structure of stannic ethylenediaminetetraacetate monohydrate. Inorg. Chem. 1971, 10, 2313–2319. [Google Scholar] [CrossRef]
  136. Poznyak, A.L.; Ilyukhin, A.B. Crystal structure of Ba[Sn(Edta)(OH)]Cl·4H2O. Kristallografiya 1997, 42, 932–933. [Google Scholar]
  137. Ilyukhin, A.B.; Logvinova, V.B.; Davidovich, R.L.; Poznyak, A.L. Crystal structure of ethylenediaminetetraacetato complexes of tetravalent tin. Zh. Neorg. Khim. 2000, 45, 1339–1343. [Google Scholar]
  138. Brouca-Cabarrecq, C.; Marrot, B.; Mosset, A. Crystal structure of aquo (hydroxo) (ethylenediaminetetraacetato) sodium(I) tin(IV), NaSn(OH)(edta)(H2O). J. Chem. Crystallogr. 1996, 26, 503–508. [Google Scholar] [CrossRef]
  139. Fu, Z.; Chen, B.; He, A.; Lv, X.; Wei, X.; Feng, X.; Huang, C.; Xia, C.; Jin, Y. The crystal structure of oxonium chlorido-ethylenediaminetetraactetotin(IV) hydrate, C10H17ClN2O10Sn. Z. Kristallogr.-New Cryst. Struct. 2017, 232, 941–942. [Google Scholar] [CrossRef]
  140. Iyer, R.K.; Deshpande, S.G.; Amirthalingam, V. Preparation, properties and molecular structure of monohydrogen diethylenetriaminepentaacetatostannate(IV) trihydrate–an eight coordinated 1:1 TIN(IV) chelate. Polyhedron 1984, 3, 1099–1104. [Google Scholar] [CrossRef]
  141. Ilyukhin, A.B.; Sergienko, V.S.; Davidovich, R.L.; Logvinova, V.B. Synthesis and structure of dioxonium diethylenetriaminepentaacetatostannate(IV) and hafnate(IV) H5O2[MDtpa]·H2O. Zh. Neorg. Khim. 1997, 42, 1474–1478. [Google Scholar]
  142. Ilyukhin, A.B.; Davidovich, R.L.; Samsonova, I.N.; Teplukhina, L.V. Eightfold-coordinated diethylenetriaminepentaacetates. Crystal structures of K[M(Dtpa)]·3H2O (M = Zr, Hf) and NH4[Sn(Dtpa)]·H2O. Kristallografiya 2000, 45, 45–49. [Google Scholar] [CrossRef]
  143. Van Remoortere, F.P.; Flynn, J.J.; Boer, F.P.; North, P.P. Crystal structure of distannous ethylenediaminetetraacetate dihydrate. Inorg. Chem. 1971, 10, 1511–1518. [Google Scholar] [CrossRef]
  144. Shields, K.G.; Seccombe, R.C.; Kennard, C.H.L. Stereochemistry of flexible-chelate–metal complexes. Part III. Crystal structure of dihydrogen ethylenediaminetetra-acetatostannate(II). J. Chem. Soc. Dalton Trans. 1973, 2, 741–743. [Google Scholar] [CrossRef]
  145. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  146. Mizuta, T.; Yamamoto, T.; Miyoshi, K.; Kushi, Y. The ligand field stabilization effect on the metal-ligand bond distances in octahedral metal complexes with edta-type ligands. Redetermination of the molecular structure of (ethylenediaminetriacetatoacetic acid)-(aqua)iron(III), [Fe(Hedta)(H2O)]. Inorg. Chim. Acta 1990, 175, 121–126. [Google Scholar] [CrossRef]
  147. Solans, X.; Font Altaba, M.; Garcia-Oricain, J. Crystal structures of ethylenediaminetetraacetato metal complexes. V. Structures containing the [Fe(C10H12N2O8)(H2O)] anion. Acta Crystallogr. C 1984, 40, 635–638. [Google Scholar] [CrossRef]
  148. Smith, F.E.; Hynes, R.C.; Ang, T.T.; Khoo, L.E.; Eng, G. Synthesis and structural characterization of a series of pentacoordinate diorganotin(IV) N-arylidene-α-amino acid complexes. Can. J. Chem. 1992, 70, 1114–1120. [Google Scholar] [CrossRef]
  149. Wang, J.; Yang, X.; Cheng, J.; Xu, Y.; Liu, B.; Wang, H.; Zhang, D. Thermodynamic α-CH acidities of amino acid fragments in five-co-ordinate bicycloazastannoxides in Me2SO. J. Chem. Soc. Dalton Trans. 1996, 25, 3889–3891. [Google Scholar] [CrossRef]
  150. Beltrán, H.I.; Zamudio-Rivera, L.S.; Mancilla, T.; Santillan, R.; Farfán, N. One-Step Preparation, Structural Assignment, and X-ray Study of 2,2-Di-nbutyl-and 2,2-Diphenyl-6-aza-1,3-dioxa-2-stannabenzocyclononen-4-ones Derived from Amino Acids. Chem. Eur. J. 2003, 9, 2291–2306. [Google Scholar] [CrossRef]
  151. Tian, L.; Qian, B.; Sun, Y.; Zheng, X.; Yang, M.; Li, H.; Liu, X. Synthesis, structural characterization and cytotoxic activity of diorganotin(IV) complexes of N-(5-halosalicylidene)-α-amino acid. Appl. Organomet. Chem. 2005, 19, 980–987. [Google Scholar] [CrossRef]
  152. Basu Baul, T.S.; Kehie, P.; Duthie, A.; Wang, R.; Englert, U.; Hopfl, H. Synthesis, characterization, crystal structures and supramolecular features of bicycloazastannoxides derived from Schiff bases with L-tyrosine. J. Organomet. Chem. 2017, 828, 96–105. [Google Scholar] [CrossRef]
  153. Basu Baul, T.S.; Kehie, P.; Duthie, A.; Guchhait, N.; Raviprakash, N.; Mokhamatam, R.B.; Manna, S.K.; Armata, N.; Scopelliti, M.; Wang, R.; et al. Synthesis, photophysical properties and structures of organotin-Schiff bases utilizing aromatic amino acid from the chiral pool and evaluation of the biological perspective of a triphenyltin compound. J. Inorg. Biochem. 2017, 168, 76–89. [Google Scholar] [CrossRef] [PubMed]
  154. Basu Baul, T.S.; Kehie, P.; Hoepfl, H.; Duthie, A.; Eng, G.; Linden, A. Organotin(IV) complexes derived from proteinogenic amino acid: Synthesis, structure and evaluation of larvicidal efficacy on Anopheles stephensi mosquito larvae. Appl. Organomet. Chem. 2017, 31, e3547. [Google Scholar] [CrossRef]
  155. Basu Baul, T.S.; Chaurasiya, A.; Rabha, M.; Khatua, S.; Lycka, A.; Schollmeyer, D.; Jurkschat, K. Diorganotin Compounds Containing α-Aminoacidato Schiff Base Ligands Derived from Functionalized 2-Hydroxy-5-(aryldiazenyl)benzaldehyde. Syntheses, Structures and Sensing of Hydrogen Sulfide. Eur. J. Inorg. Chem. 2020, 2020, 1803–1813. [Google Scholar] [CrossRef]
  156. Singh, N.; Kumar, K.; Srivastav, N.; Singh, R.; Kaur, V.; Jasinski, J.P.; Butcher, R.J. Exploration of fluorescent organotin compounds of α-amino acid Schiff bases for the detection of organophosphorous chemical warfare agents: Quantification of diethylchlorophosphate. New J. Chem. 2018, 42, 8756–8764. [Google Scholar] [CrossRef]
  157. Basu Baul, T.S.; Masharing, C.; Willem, R.; Biesemans, M.; Holcapek, M.; Jirasko, R.; Linden, A. Self-assembly of diorganotin(IV) 2-{[(E)-1-(2-oxyaryl)alkylidene]amino}acetates: An investigation of structures by X-ray diffraction, solution and solid-state tin NMR, and electrospray ionization MS. J. Organomet. Chem. 2005, 690, 3080–3094. [Google Scholar] [CrossRef]
  158. Böhme, U.; Weling, G. The crystal structure of trichlorido[N-[(2-oxyphenyl)methylidene]phenylglycinemethylester-κ3 O,N,O′]- tin(IV)–methylene chloride (1/1), C16H14Cl3NO3Sn·CH2Cl2. Z. Kristallogr.-New Cryst. Struct. 2023, 238, 391–393. [Google Scholar] [CrossRef]
  159. Schwarzer, S.; Böhme, U.; Fels, S.; Günther, B.; Brendler, E. (S)-N-[(2-hydroxynaphthalen-1-yl)methylidene]valine—A valuable ligand for the preparation of chiral complexes. Inorg. Chim. Acta 2018, 483, 136–147. [Google Scholar] [CrossRef]
  160. Nath, M.; Goyal, S. Synthesis, Characteristic Spectral Studies and in Vitro Antimicrobial Activity of Organosilicon(IV) Complexes of N-(2-Hydroxynaphthalidene)-Amino Acid Schiff Bases. Phosphorus Sulfur Silicon Relat. Elem. 2002, 177, 447–463. [Google Scholar] [CrossRef]
  161. Warncke, G.; Böhme, U.; Günther, B.; Kronstein, M. Racemization versus retention of chiral information during the formation of silicon and tin complexes with chiral Schiff base ligands. Polyhedron 2012, 47, 46–52. [Google Scholar] [CrossRef]
  162. Schwarzer, A.; Fels, S.; Böhme, U. Two reversible enantiotropic phase transitions in a pentacoordinate silicon complex with an O,N,O’-tridentate valinate ligand. Acta Crystallogr. C 2015, 71, 511–516. [Google Scholar] [CrossRef] [PubMed]
  163. Böhme, U.; Fels, F. A new type of chiral pentacoordinated silicon compounds with azomethine ligands made from acetylacetone and amino acids. Inorg. Chim. Acta 2013, 406, 251–255. [Google Scholar] [CrossRef]
  164. Böhme, U.; Fels, S. {N-[(4-Methoxy-2-oxidophenyl)(phenyl)methylidene]glycinato}diphenylsilicon(IV). IUCrData 2023, 8, x230306. [Google Scholar] [CrossRef] [PubMed]
  165. Böhme, U.; Fels, S. The crystal structure of [N-{[2-(oxy)-4-methoxyphenyl](phenyl)methylidene}alaninato]-diphenyl-silicon(IV) –chloroform (1/1),C29H25NO4Si·CHCl3. Z. Kristallogr.-N. Cryst. Struct. 2023, 238, 603–605. [Google Scholar] [CrossRef]
  166. Böhme, U.; Wiesner, S.; Günther, B. Easy access to chiral penta- and hexacoordinate silicon compounds. Inorg. Chem. Commun. 2006, 9, 806–809. [Google Scholar] [CrossRef]
  167. Schlecht, S.; Finze, M.; Bertermann, R.; Frank, W.; Domann, A.; Braun, M. Cocrystallising Rotamers of a Pentacoordinate Silicon Complex with a Chiral Aminodiol Ligand. Eur. J. Inorg. Chem. 2013, 2013, 1488–1492. [Google Scholar] [CrossRef]
  168. Tasaka, M.; Hirotsu, M.; Kojima, M.; Utsuno, S.; Yoshikawa, Y. Syntheses, Characterization, and Molecular Mechanics Calculations of Optically Active Silatrane Derivatives. Inorg. Chem. 1996, 35, 6981–6986. [Google Scholar] [CrossRef] [PubMed]
  169. Gauchenova, E.V.; Karlov, S.S.; Selina, A.A.; Chernyshova, E.S.; Churakov, A.V.; Howard, J.A.K.; Troitsky, N.A.; Tandura, S.N.; Lorberth, J.; Zaitseva, G.S. Synthesis and characterization of 3- and 4-phenylgermatranes: X-ray crystal structures of N(CH2CH2O)2(CH2CHPhO)GeZ (Z=F, OSiMe3, CCPh) and N(CH2CH2O)2(CHPhCH2O)GeOH. J. Organomet. Chem. 2003, 676, 8–21. [Google Scholar] [CrossRef]
  170. Glowacki-Pallach, B.; Lutter, M.; Schollmeyer, D.; Hiller, W.; Jouikov, V.; Jurkschat, K. Extending Chirality in Group XIV Metallatranes. Inorg. Chem. 2023, 62, 7662–7680. [Google Scholar] [CrossRef] [PubMed]
  171. Rohovec, J.; Lukeš, I.; Vojtíšek, P.; Císařová, I.; Hermann, P. Complexing properties of phosphinic analogues of glycine. J. Chem. Soc. Dalton Trans. 1996, 25, 2685–2691. [Google Scholar] [CrossRef]
  172. Cameron, T.S.; Prout, C.K.; Rossotti, F.J.C.; Steele, D. Crystal and molecular structure of bis(aminomethanesulphonato)copper(II). J. Chem. Soc. Dalton Trans. 1973, 2, 1590–1592. [Google Scholar] [CrossRef]
  173. Lutter, M.; Jurkschat, K. Aryl(dimethylaminomethyl)phosphinic Acid Esters: Syntheses, Structures, and Reactions with Halogen Hydrogen Acids, Tin Halides, and Trimethyl Halosilanes. Eur. J. Inorg. Chem. 2018, 2018, 3481–3490. [Google Scholar] [CrossRef]
  174. Knerr, S.; Brendler, E.; Zuber, J.; Kroke, E.; Wagler, J. A new class of silatranes derived from nitrilotris(methylenephenylphosphinic) acid. J. Organomet. Chem. 2025, 1023, 123402. [Google Scholar] [CrossRef]
  175. Knerr, S.; Brendler, E.; Gericke, R.; Kroke, E.; Wagler, J. Two Modifications of Nitrilotris(methylenephenylphosphinic) Acid: A Polymeric Network with Intermolecular (O=P–O–H)3 vs. Monomeric Molecules with Intramolecular (O=P–O–H)3 Hydrogen Bond Cyclotrimers. Crystals 2024, 14, 662. [Google Scholar] [CrossRef]
  176. Zhou, W.; Zhang, J.; Sun, Z.-G.; Zhu, Y.-Y.; Jiao, C.-Q.; Shi, S.-P.; Dai, L.-L.; Sun, T.; Li, W.-Z.; Ma, M.-X.; et al. Synthesis, crystal structures, and luminescent properties of tin(II) and lead(II) carboxyphosphonates with 3D framework structures. Inorg. Chem. Commun. 2014, 47, 37–41. [Google Scholar] [CrossRef]
  177. Perry, H.P.; Gagnon, K.J.; Law, J.; Teat, S.; Clearfield, A. Divalent metal phosphonate coordination polymers constructed from a dipiperidine-based bisphosphonate ligand. Dalton Trans. 2012, 41, 3985–3994. [Google Scholar] [CrossRef] [PubMed]
  178. Jiao, C.; Sun, Z.; Zhu, Y.; Chen, K.; Zhu, J.; Li, C.; Wang, C.; Sun, S.; Tian, H.; Chu, W.; et al. Hydrothermal syntheses, crystal structures and luminescence properties of three new metal diphosphonates with layered structure. Inorg. Chim. Acta 2012, 387, 186–194. [Google Scholar] [CrossRef]
  179. Lorberth, J.; Shin, S.-H.; Otto, M.; Wocadlo, S.; Massa, W.; Yashina, N.S. {Me2NCH[(EtO)2P=O]2 − Me2SnCl2}-dimer: Crystal structure of a binuclear bis(chelate) complex of tin(IV) with cis-linked octahedra. J. Organomet. Chem. 1991, 407, 313–318. [Google Scholar] [CrossRef]
  180. Khoo, L.E.; Hazell, A. Polymeric [μ3-(N-phosphonomethyl)-glycinato]tris(tri-n-butyltin). Acta Crystallogr. C 2001, 57, 254–256. [Google Scholar] [CrossRef] [PubMed]
  181. An, B.-H.; Zhang, R.-F.; Li, Q.-L.; Du, X.-M.; Ru, J.; Zhang, S.-L.; Ma, C.-L. Syntheses, structures and in vitro cytostatic activity of four novel homochiral organotin(IV) phosphonates. J. Organomet. Chem. 2019, 881, 51–57. [Google Scholar] [CrossRef]
  182. Lorberth, J.; Wocadlo, S.; Massa, W.; Grigoriev, E.V.; Yashina, N.S.; Petrosyan, V.S.; Finocchiaro, P. Synthesis and structure of 1:1 complex of dimethyltin dichloride with meso-1,2-bis(α-diethoxyphosphorylbenzylamino)ethane. J. Organomet. Chem. 1996, 510, 287–290. [Google Scholar] [CrossRef]
  183. Dakternieks, D.; Jurkschat, K.; Tiekink, E.R.T. Crystal structure of 1-methyl-5-methyl-5-(O-tert-butylphosphonic acid)-1-aza-5-stannabicyclo [3.3.01,5]octane, (C8H17N)SnOPO2C(CH3)3. Z. Kristallogr. 1996, 211, 755–756. [Google Scholar] [CrossRef]
  184. Thamyongkit, P.; Bhise, A.D.; Taniguchi, M.; Lindsey, J.S. Alkylthio Unit as an r-Pyrrole Protecting Group for Use in Dipyrromethane Synthesis. J. Org. Chem. 2006, 71, 903–910. [Google Scholar] [CrossRef] [PubMed]
  185. Purdy, A.P.; Gilardi, R.; Luther, J.; Butcher, R.J. Synthesis, crystal structure, and reactivity of alkali and silver salts of sulfonated imidazoles. Polyhedron 2007, 26, 3930–3938. [Google Scholar] [CrossRef]
  186. Rutsch, P.; Renner, G.; Huttner, G.; Sandhöfner, S. Organometallically Protected Indates and Germanates: Synthesis, Structure and Properties of [{(CO)5M}EX3]2- (M = Cr, Mo, W.; E = In; X = Cl, Br), [(CO5Cr)InBr(μ2-Br)]22- and [{(CO)5Cr}E(oxinat)2]n- (E = In, n = 2; E = Ge, n = 1). Z. Naturforsch. B 2002, 57, 757–772. [Google Scholar] [CrossRef]
  187. Kircher, P.; Huttner, G.; Schiemenz, B.; Heinze, K.; Zsolnai, L.; Walter, O.; Jacobi, A.; Driess, A. Divalent Tin and Lead Compounds EX2 Acting as Bridging Ligands in [{(CO)5M}2EX2]2− (E = Sn, Pb; M = Cr, Mo, W): Preparation and Properties. Chem. Ber. 1997, 130, 687–699. [Google Scholar] [CrossRef]
Figure 1. Hypercoordinate Si-, Ge- and Sn-complexes I [26], II [27], III [28], IV [29], V [30], VI [31], VII [32], VIII [33], IX [34], X [35] and XI [36] with mono-anionic (O,N)-bidentate chelating ligands.
Figure 1. Hypercoordinate Si-, Ge- and Sn-complexes I [26], II [27], III [28], IV [29], V [30], VI [31], VII [32], VIII [33], IX [34], X [35] and XI [36] with mono-anionic (O,N)-bidentate chelating ligands.
Molecules 30 00834 g001
Figure 2. Selected (O,N)-chelators, which feature an α-amino acid motif (drawn with bold-style bonds). The fundamental features of N-alkylated amino acids, such as N,N-dimethylaminoacetic acid (a), and NH-bearing amino acids, such as glycine (b), may be contained in or give rise to more complex chelators such as chiral amino acids and peptides (c), amineoligoacetic acids (d) and alkanolamine acetic acids (e). This variety can be extended to amino acid-derived ligands, which are devoid of an amino and/or carboxylic acid group, e.g., in Schiff bases, which originate from α-amino acids (f). In (c) and (f), R represents a variety of substituents, and the asterisk (*) indicates a center of chirality. Related features may be contained in the other classes of ligands as well.
Figure 2. Selected (O,N)-chelators, which feature an α-amino acid motif (drawn with bold-style bonds). The fundamental features of N-alkylated amino acids, such as N,N-dimethylaminoacetic acid (a), and NH-bearing amino acids, such as glycine (b), may be contained in or give rise to more complex chelators such as chiral amino acids and peptides (c), amineoligoacetic acids (d) and alkanolamine acetic acids (e). This variety can be extended to amino acid-derived ligands, which are devoid of an amino and/or carboxylic acid group, e.g., in Schiff bases, which originate from α-amino acids (f). In (c) and (f), R represents a variety of substituents, and the asterisk (*) indicates a center of chirality. Related features may be contained in the other classes of ligands as well.
Molecules 30 00834 g002
Figure 3. Crystallographically characterized hexa- (1) and pentacoordinate (2a2e) silicon complexes, which feature a mono-anion of an α-amino acid among the (O,N)-chelating ligands. For clarity, the formally dative bond between Si and the amino group is represented by a dashed line.
Figure 3. Crystallographically characterized hexa- (1) and pentacoordinate (2a2e) silicon complexes, which feature a mono-anion of an α-amino acid among the (O,N)-chelating ligands. For clarity, the formally dative bond between Si and the amino group is represented by a dashed line.
Molecules 30 00834 g003
Figure 4. Zwitterionic spirosilicates with two di-anions of α-amino acids as (O,N)-chelators (3a3f) [49,50]. The substituents R,R′ at N and the α-C atom in the generic drawings of the diastereomers 3 and 3 may also represent -CH2CH2CH2- and thus account for the di-anion of l-proline as an (O,N)-chelator as well.
Figure 4. Zwitterionic spirosilicates with two di-anions of α-amino acids as (O,N)-chelators (3a3f) [49,50]. The substituents R,R′ at N and the α-C atom in the generic drawings of the diastereomers 3 and 3 may also represent -CH2CH2CH2- and thus account for the di-anion of l-proline as an (O,N)-chelator as well.
Molecules 30 00834 g004
Figure 5. O,N-bis-silylated α-amino acids: Generic cyclic 4 and spirocyclic types 4′ (R may be various substituents, and the amino acid backbone may carry further substituents not shown in this generic sketch), bis-silylated l-tryptophan 5 [52] and structurally characterized Lewis-base adducts of silacycles of l-valine [39], α-amino isobutyric acid [37] and d-phenylglycine [37] (6a, 6b and 6c, respectively).
Figure 5. O,N-bis-silylated α-amino acids: Generic cyclic 4 and spirocyclic types 4′ (R may be various substituents, and the amino acid backbone may carry further substituents not shown in this generic sketch), bis-silylated l-tryptophan 5 [52] and structurally characterized Lewis-base adducts of silacycles of l-valine [39], α-amino isobutyric acid [37] and d-phenylglycine [37] (6a, 6b and 6c, respectively).
Molecules 30 00834 g005
Figure 6. Sn(IV)-complexes of the amino acids glycine (7a) and l-cysteine (7b), as well as Sn(IV) (8a, 8b) and Sn(II) (8c) complexes of cysteine alkyl esters. The Sn–N as well as Sn–O and Sn–S (if applicable) bond lengths (Å, rounded to two decimal places, where their s.u.s for the 2nd decimal place are (3) for 7a and ≤(1) for the other compounds) are listed next to the respective bond. In the drawings of 7a and 7b, the atoms (Sn) and (O) in parentheses represent the respective atoms of adjacent molecules in the crystal structure.
Figure 6. Sn(IV)-complexes of the amino acids glycine (7a) and l-cysteine (7b), as well as Sn(IV) (8a, 8b) and Sn(II) (8c) complexes of cysteine alkyl esters. The Sn–N as well as Sn–O and Sn–S (if applicable) bond lengths (Å, rounded to two decimal places, where their s.u.s for the 2nd decimal place are (3) for 7a and ≤(1) for the other compounds) are listed next to the respective bond. In the drawings of 7a and 7b, the atoms (Sn) and (O) in parentheses represent the respective atoms of adjacent molecules in the crystal structure.
Molecules 30 00834 g006
Figure 7. Crystallographically characterized Si-complexes of di-anions of valinylvaline (9a) and α-aminoisobutyrylglycine (9b), Ge-complexes of di-anions of glycylmethionine (10a) and glycylglycine (10b) and, as a representative example, the related SnMe2 complex of glycylmethionine (11a). Selected bond lengths (Å, rounded to two decimal places, where their s.u.s for the 2nd decimal place are <(1)) are listed next to the respective bond. For compound 9a, the values are listed for the two crystallographically independent molecules. For compound 9b, which contains six molecules in the asymmetric unit, the bond lengths vary in the ranges 1.80–1.82 Å for Si–O, 1.75–1.77 Å for the equatorial and 2.02–2.05 Å for the axial Si–N bond.
Figure 7. Crystallographically characterized Si-complexes of di-anions of valinylvaline (9a) and α-aminoisobutyrylglycine (9b), Ge-complexes of di-anions of glycylmethionine (10a) and glycylglycine (10b) and, as a representative example, the related SnMe2 complex of glycylmethionine (11a). Selected bond lengths (Å, rounded to two decimal places, where their s.u.s for the 2nd decimal place are <(1)) are listed next to the respective bond. For compound 9a, the values are listed for the two crystallographically independent molecules. For compound 9b, which contains six molecules in the asymmetric unit, the bond lengths vary in the ranges 1.80–1.82 Å for Si–O, 1.75–1.77 Å for the equatorial and 2.02–2.05 Å for the axial Si–N bond.
Molecules 30 00834 g007
Figure 8. Generic representation of the Sn-coordination modes encountered in the crystal structures of dipeptide-derived diorganotin(IV) compounds (substituents at the dipeptide backbones and at the Sn-bound C atoms, if any, are omitted, and adjacent molecules are represented by the essential coordinating moiety only). (a) Pentacoordination of Sn (distorted trigonal–bipyramidal) with O(axial)-N(equatorial)-NH2(axial) positioning of the di-anionic dipeptide-derived ligand and Sn–C bonds in equatorial positions. (b) [5 + 2]-coordination, where two remote O⋯Sn interactions with a carboxylate moiety of an adjacent molecule enhance the Sn-coordination number in the idealized Sn(O,N,N) plane. (c) [5 + 1]-coordination, where one remote O⋯Sn interaction with a carbonyl O atom of the carboxylate moiety of an adjacent molecule enhances the Sn-coordination number in the idealized Sn(O,N,N) plane. The other O atom of this carboxylate group may be involved in H-bonding with the NH2 group (as indicated by an additional dashed line). (d) [5 + 1]-coordination, where one remote N⋯Sn interaction with an imidazole N atom of a histidine group of an adjacent molecule enhances the Sn-coordination number in the idealized Sn(O,N,N) plane.
Figure 8. Generic representation of the Sn-coordination modes encountered in the crystal structures of dipeptide-derived diorganotin(IV) compounds (substituents at the dipeptide backbones and at the Sn-bound C atoms, if any, are omitted, and adjacent molecules are represented by the essential coordinating moiety only). (a) Pentacoordination of Sn (distorted trigonal–bipyramidal) with O(axial)-N(equatorial)-NH2(axial) positioning of the di-anionic dipeptide-derived ligand and Sn–C bonds in equatorial positions. (b) [5 + 2]-coordination, where two remote O⋯Sn interactions with a carboxylate moiety of an adjacent molecule enhance the Sn-coordination number in the idealized Sn(O,N,N) plane. (c) [5 + 1]-coordination, where one remote O⋯Sn interaction with a carbonyl O atom of the carboxylate moiety of an adjacent molecule enhances the Sn-coordination number in the idealized Sn(O,N,N) plane. The other O atom of this carboxylate group may be involved in H-bonding with the NH2 group (as indicated by an additional dashed line). (d) [5 + 1]-coordination, where one remote N⋯Sn interaction with an imidazole N atom of a histidine group of an adjacent molecule enhances the Sn-coordination number in the idealized Sn(O,N,N) plane.
Molecules 30 00834 g008
Figure 9. Examples of hexacoordinate Si- and Sn-complexes (XII and XIII, respectively) with two tridentate di-anionic (O,N,N)-chelating ligands and chelation through exclusive formation of five-membered rings.
Figure 9. Examples of hexacoordinate Si- and Sn-complexes (XII and XIII, respectively) with two tridentate di-anionic (O,N,N)-chelating ligands and chelation through exclusive formation of five-membered rings.
Molecules 30 00834 g009
Figure 10. Generic structural motif of silatranes and various derivatives thereof (12), and listing of the crystallographically characterized silatranones (12a [93], 12b [94], 12c [95], 12d [95], 12e [96] and 12f [97]) and silatranedione 12g [98].
Figure 10. Generic structural motif of silatranes and various derivatives thereof (12), and listing of the crystallographically characterized silatranones (12a [93], 12b [94], 12c [95], 12d [95], 12e [96] and 12f [97]) and silatranedione 12g [98].
Molecules 30 00834 g010
Figure 11. Crystallographically characterized germatranone 13 [102], its anionic relative 13′ (in the potassium salt) [103] and 14 [104]. The counter-ion in compound 14 is piperazinium.
Figure 11. Crystallographically characterized germatranone 13 [102], its anionic relative 13′ (in the potassium salt) [103] and 14 [104]. The counter-ion in compound 14 is piperazinium.
Molecules 30 00834 g011
Figure 12. Generic drawing of germanium complexes of type 15 derived from EDTA and related acids with C2-backbone R′ [106,107,108,109,110] (for details of R and R′, see Table 3), complex 16 derived from a diamine tetraacetic acid with a C3 backbone [111] as well as examples of Ge-containing heteronuclear complexes derived from 2-propanol-1,3-diamine tetraacetic acid (17 [103,113,114,115,116] and 17 [117,118]) (for details of MLn and “O”, see Table 4 and descriptions in the Discussion). Note: In each case, only one enantiomer is drawn as a representative example. In compound 16, the Ge–N, Ge–O(trans-N) and Ge–O(trans-O) bond lengths are 2.05, 1.84 and 1.88 Å, respectively.
Figure 12. Generic drawing of germanium complexes of type 15 derived from EDTA and related acids with C2-backbone R′ [106,107,108,109,110] (for details of R and R′, see Table 3), complex 16 derived from a diamine tetraacetic acid with a C3 backbone [111] as well as examples of Ge-containing heteronuclear complexes derived from 2-propanol-1,3-diamine tetraacetic acid (17 [103,113,114,115,116] and 17 [117,118]) (for details of MLn and “O”, see Table 4 and descriptions in the Discussion). Note: In each case, only one enantiomer is drawn as a representative example. In compound 16, the Ge–N, Ge–O(trans-N) and Ge–O(trans-O) bond lengths are 2.05, 1.84 and 1.88 Å, respectively.
Molecules 30 00834 g012
Figure 13. Generic drawing of tin complexes 18 as well as examples 18a and 18c (for details of L, R and R′ in 18a18g, see Table 5), related alcoholate bridged dimeric complex 19 [126] and complexes 20 [127] and 21 [128] with related ligands. The counter-ion in the crystal structure of the anionic complex 21 is piperazinium. Note: The Sn atoms of the bridging Me2SnCl2 moieties in 18a are located in distorted octahedral coordination spheres.
Figure 13. Generic drawing of tin complexes 18 as well as examples 18a and 18c (for details of L, R and R′ in 18a18g, see Table 5), related alcoholate bridged dimeric complex 19 [126] and complexes 20 [127] and 21 [128] with related ligands. The counter-ion in the crystal structure of the anionic complex 21 is piperazinium. Note: The Sn atoms of the bridging Me2SnCl2 moieties in 18a are located in distorted octahedral coordination spheres.
Molecules 30 00834 g013
Figure 14. Molecular drawings of crystallographically characterized compounds with the stannatranetrione motif with a penta- (22) [129] hexa- (23) [130], hepta- (24) and octacoordinated Sn atom (25). The anionic complex 24 was characterized in its Cs salt [131], and the anion 25 was reported for its, for example, K [132], Rb [132], methylammonium [133] and guanidinium salt [133].
Figure 14. Molecular drawings of crystallographically characterized compounds with the stannatranetrione motif with a penta- (22) [129] hexa- (23) [130], hepta- (24) and octacoordinated Sn atom (25). The anionic complex 24 was characterized in its Cs salt [131], and the anion 25 was reported for its, for example, K [132], Rb [132], methylammonium [133] and guanidinium salt [133].
Molecules 30 00834 g014
Figure 15. Molecular drawings of crystallographically characterized tin compounds of the anions of ethylenediaminetetraacetic acid and diethylenetriaminepentaacetic acid. In compounds of type 26, the ligand L may be formally charge-neutral (like OHSnCl3(H2O) [134] or H2O [135]) or anionic (like hydroxide [136,137,138], fluoride [137] or chloride [139]), resulting in non-charged or anionic Sn-complexes, respectively. Compounds of type 27 were reported as the hydrated protio form [140,141] and as the ammonium salt [142]. In the drawing of 28, the atoms (Sn) in parentheses represent the additional Sn(II) sites in the crystal structure, which feature SnO4-coordination spheres.
Figure 15. Molecular drawings of crystallographically characterized tin compounds of the anions of ethylenediaminetetraacetic acid and diethylenetriaminepentaacetic acid. In compounds of type 26, the ligand L may be formally charge-neutral (like OHSnCl3(H2O) [134] or H2O [135]) or anionic (like hydroxide [136,137,138], fluoride [137] or chloride [139]), resulting in non-charged or anionic Sn-complexes, respectively. Compounds of type 27 were reported as the hydrated protio form [140,141] and as the ammonium salt [142]. In the drawing of 28, the atoms (Sn) in parentheses represent the additional Sn(II) sites in the crystal structure, which feature SnO4-coordination spheres.
Molecules 30 00834 g015
Figure 16. Molecular sketches of selected crystallographically characterized Sn-complexes with Schiff-base-type tridentate (O,N,O)-chelating ligands derived from α-amino acids. In compounds XIVf and XIVg, the (Sn) and (O) atoms in parentheses indicate additional coordination with the respective atomic sites of adjacent molecules in the crystal structure.
Figure 16. Molecular sketches of selected crystallographically characterized Sn-complexes with Schiff-base-type tridentate (O,N,O)-chelating ligands derived from α-amino acids. In compounds XIVf and XIVg, the (Sn) and (O) atoms in parentheses indicate additional coordination with the respective atomic sites of adjacent molecules in the crystal structure.
Molecules 30 00834 g016
Figure 17. Molecular sketches of selected crystallographically characterized Ge- and Si-complexes with Schiff-base-type tridentate (O,N,O)-chelating ligands derived from α-amino acids.
Figure 17. Molecular sketches of selected crystallographically characterized Ge- and Si-complexes with Schiff-base-type tridentate (O,N,O)-chelating ligands derived from α-amino acids.
Molecules 30 00834 g017
Figure 18. Pentacoordinate Si- and Ge-compounds with chiral chelating ligands that originate from chiral α-amino acids (XXII [166], XXIII [167], XXIV [168]) and from epoxides (XXV [169] and XXVI [170]). The asterisk (*) denotes the asymmetric C atom.
Figure 18. Pentacoordinate Si- and Ge-compounds with chiral chelating ligands that originate from chiral α-amino acids (XXII [166], XXIII [167], XXIV [168]) and from epoxides (XXV [169] and XXVI [170]). The asterisk (*) denotes the asymmetric C atom.
Molecules 30 00834 g018
Figure 19. Molecular sketches of Sn-, Ge- and Si-compounds with ligands of the α-aminophosphonic acid or α-aminophosphinic acid type. For compounds XXVII, 14 and the set of compounds XXVIII, the E⋯N distances (E = Sn, Ge, Si) in Å are listed. For compounds XXX and XXXI, dashed lines indicate intermolecular Sn⋯O coordination in the crystal structure.
Figure 19. Molecular sketches of Sn-, Ge- and Si-compounds with ligands of the α-aminophosphonic acid or α-aminophosphinic acid type. For compounds XXVII, 14 and the set of compounds XXVIII, the E⋯N distances (E = Sn, Ge, Si) in Å are listed. For compounds XXX and XXXI, dashed lines indicate intermolecular Sn⋯O coordination in the crystal structure.
Molecules 30 00834 g019
Figure 20. Molecular sketches of compounds which feature SO moieties inside five-membered (O,N)-Sn-chelates (XXXII) and which may be considered precursors for SO-functionalized (O,N)-chelating ligands (XXXIII and XXXIV).
Figure 20. Molecular sketches of compounds which feature SO moieties inside five-membered (O,N)-Sn-chelates (XXXII) and which may be considered precursors for SO-functionalized (O,N)-chelating ligands (XXXIII and XXXIV).
Molecules 30 00834 g020
Table 1. Selected bond lengths (Å) and angles (deg.) of the Si-coordination spheres of compounds 1, 2a2e, 3a3f and 6a6c. The values for each column are rounded to the same number of decimal places, and the s.u. on the last decimal place listed is ≤(1).
Table 1. Selected bond lengths (Å) and angles (deg.) of the Si-coordination spheres of compounds 1, 2a2e, 3a3f and 6a6c. The values for each column are rounded to the same number of decimal places, and the s.u. on the last decimal place listed is ≤(1).
AxisAxial AngleSi–N 1,2Si⋯N 1,2Si–O 1,2Si–O 1,2
1H2N-Si-NH2172.1n/a1.89 3n/a1.80 3
O-Si-(NCO)176.8 3
2aO-Si-NH2167.61.69 e1.99 a1.79 a1.71 e
2bO-Si-NH2168.31.69 e1.97 a1.79 a1.71 e
2c 4O-Si-NH2169.0, 170.61.71, 1.71 e1.97, 1.98 a1.80, 1.80 a1.72, 1.72 e
2dO-Si-NH2169.41.70 e1.97 a1.80 a1.70 e
2eO-Si-O164.91.70 e1.88 e1.79 a1.82 a
3aO-Si-O176.31.71, 1.72 en/a1.84, 1.82 an/a
3bO-Si-O178.51.71, 1.72 en/a1.83, 1.81 an/a
3cO-Si-O175.81.73, 1.73 en/a1.84, 1.82 an/a
3d 5O-Si-O173.1–174.51.71–1.72 en/a1.81–1.83 an/a
3eO-Si-O176.81.71, 1.71 en/a1.81, 1.83 an/a
3f 5O-Si-O177.6–179.31.73–1.74 en/a1.81–1.85 an/a
6aO-Si-NH3171.11.72 e2.011.88 an/a
6bO-Si-NMI172.31.71 e2.041.85 an/a
6cO-Si-NMI171.31.72 e2.011.87 an/a
1 The bonds refer to those of the amino acids’ anions in the order Si–N to the di-anionic ligand, Si⋯N to the mono-anionic ligand, Si–O to the di-anionic ligand, Si–O to the mono-anionic ligand. 2 The superscript indices (a) and (e) indicate axial and equatorial positioning, respectively, of the respective donor atom (N or O) in the Si-coordination sphere 3 The two features of the respective entry are identical because of crystallographic symmetry. 4 The two lines contain entries for the two crystallographically independent molecules. 5 Ranges are given for the corresponding data of the multiple independent molecules of the crystal structures.
Table 2. Synthesis routes and selected bond lengths (Å) and angles (deg.) of the Sn-coordination spheres of crystallographically characterized tin compounds (11a11p) of the type (a-b)SnR2, which feature a tridentate di-anionic dipeptide-derived (O,N,N)-chelating ligand, with a and b being the C- and N-termini of the dipeptide-di-anion and R being the Sn-bound hydrocarbyl groups. The values for each column are rounded to the same number of decimal places, and the s.u. on the last decimal place listed is ≤(1).
Table 2. Synthesis routes and selected bond lengths (Å) and angles (deg.) of the Sn-coordination spheres of crystallographically characterized tin compounds (11a11p) of the type (a-b)SnR2, which feature a tridentate di-anionic dipeptide-derived (O,N,N)-chelating ligand, with a and b being the C- and N-termini of the dipeptide-di-anion and R being the Sn-bound hydrocarbyl groups. The values for each column are rounded to the same number of decimal places, and the s.u. on the last decimal place listed is ≤(1).
Ref.Route(a-b)R 2O-Sn-NC-Sn-CSi–OSi–NSi⋯NMotif
11a[73]B(met-gly)Me153.0123.82.162.072.25a
11b[74]C(phe-tyr)Me149.5136.52.202.102.25c 3
11c 1[75]A(his-met)Me147.0124.72.142.122.30a
152.3126.12.182.072.24a
11d[76]A(met-met)Me148.8132.02.192.112.26c 3
11e[76]A(his-ala)Me146.4143.92.222.142.28b 3
11f[77]A(ala-trp)Me151.5123.82.182.062.27a
11g 1[78]A(his-gly)Et151.9128.82.182.102.27a
150.0148.52.242.132.27d 3
11h[79]C(tyr-gly)Et152.2131.42.192.082.29a
11i[80]C(gly-trp)nBu147.9151.92.252.122.29b 3
11j[81]A(leu-aib)nBu148.7127.82.202.112.28a
11k[81]A(ala-aib)nBu147.6129.82.242.122.24c 3
11l[82]A,B,C(val-gly)nBu151.3125.32.142.102.27a
11m[83]A(gly-gly)Cy151.4123.62.172.102.28a
11n[83]A(ala-gly)Cy150.8123.02.162.092.30a
11o[84]C(gly-gly)tBu149.6121.72.202.092.29a
11p[85]A(gly-gly)Ph153.2117.52.162.082.27a
1 Data listed for two independent molecules in the crystallographic asymmetric unit. 2 The abbreviations correspond to ala = alanine, aib = α-amino isobutyric acid, gly = glycine, his = histidine, leu = leucine, met = methionine, phe = phenylalanine, trp = tryptophan, tyr = tyrosine, val = valine. 3 The additional atom distances, which contribute to the remote [5 + 1]- or [5 + 2]-coordination of the Sn atoms in these compounds, are: Sn⋯O 3.12 Å in 11b, Sn⋯O 3.14 Å in 11d, Sn⋯O 3.00 and 3.13 Å in 11e, Sn⋯N 2.79 Å in 11g, Sn⋯O 2.80 and 2.85 Å in 11i and Sn⋯O 2.82 Å in 11k.
Table 3. Bond lengths (Å) of the Ge-coordination spheres of crystallographically characterized Ge-complexes of type 15. The bond labels correspond to Figure 12. The values for each column are rounded to the same number of decimal places.
Table 3. Bond lengths (Å) of the Ge-coordination spheres of crystallographically characterized Ge-complexes of type 15. The bond labels correspond to Figure 12. The values for each column are rounded to the same number of decimal places.
Ref.R1Ge–N (a)Ge–N (b)Ge–O (c)Ge–O (d)Ge–O (e)Ge–O (f)
15a[106]CH2CH22.082.111.881.881.891.77
15b[107]CH2CH22.092.101.891.881.901.76
15c[108]CH2CHMe22.092.101.881.871.891.78
15d[109]1,2-C6H1032.072.131.881.861.891.77
15e 4[110]CH2CH22.092.091.881.901.911.76
2.072.111.871.881.911.77
15f[108]CH2CH22.092.111.891.881.891.75
1 The dangling ligand arms R are -CH2COOH (in 15a, 15c, 15d), -CH2COO as N,N′-diphenylguanidinium salt (in 15b), -CH2CH2OH (in 15e) and a zwitterionic form of CH2CH2N(CH2COOH)2 (in 15f). 2 The Me group of the backbone is bound to the C atom at the tripod-center N-donor atom. 3 trans-1,2-cyclohexanediyl backbone. 4 Data are listed for two crystallographically independent molecules in the asymmetric unit.
Table 4. Bond lengths (Å) of the Ge-coordination spheres of crystallographically characterized heteronuclear complexes of type 17. The bond labels correspond to Figure 12. The values for each column are rounded to the same number of decimal places.
Table 4. Bond lengths (Å) of the Ge-coordination spheres of crystallographically characterized heteronuclear complexes of type 17. The bond labels correspond to Figure 12. The values for each column are rounded to the same number of decimal places.
Ref.MLnGe–N (a)Ge–O (b)Ge–O (b’)Ge–O (c)Ge–O (d)Ge–O (e)
17a[103]La(H2O)42.101.931.931.861.811.80
17b[113]Nd(H2O)42.101.931.921.861.811.80
17c[114]Tb(H2O)32.101.931.921.871.811.81
17d[115]Tm(H2O)32.101.931.911.871.811.81
17e[114]Yb(H2O)32.101.931.921.861.811.81
17f[116]Cu(H2O)2.091.911.911.841.90 11.79
1 In compound 17f, the ligand “O” is H2O.
Table 5. Bond lengths (Å) and C-Sn-C angles (deg.) of the Sn-coordination spheres of complexes 18a18g. The positions of L, R and R′ correspond to 18 in Figure 13. The values for each column are rounded to the same number of decimal places.
Table 5. Bond lengths (Å) and C-Sn-C angles (deg.) of the Sn-coordination spheres of complexes 18a18g. The positions of L, R and R′ correspond to 18 in Figure 13. The values for each column are rounded to the same number of decimal places.
Ref.L,R,RC-Sn-CSn–NSn–O 1Sn–O 2Sn–O(L)Sn⋯O
18a[119]-,H,Me134.32.272.172.17n/an/a
18b[120]-,mTol,nBu132.42.232.132.14n/an/a
18c[121]H2O,H,Me161.12.292.202.372.382.79
18d[122]H2O,Me,Me162.52.362.192.352.362.74
18e[123]MeOH,H,Me159.52.28 32.162.352.372.95
18f[124]H2O,H,nBu162.72.322.212.352.362.77
18g[125]H2O,H,nBu163.42.302.202.332.312.76
1 This bond is to the non-bridging carboxylate O of the (O,N,O)-ligand. 2 This bond is to the bridging carboxylate O of the (O,N,O)-ligand, which is involved in remote O⋯Sn-coordination with an adjacent complex molecule. 3 Average value from two disorder positions of the N atom site.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Seidel, A.; Wagler, J. Anions of α-Amino Acids as (O,N)-Donor Ligands in Si-, Ge- and Sn-Coordination Chemistry. Molecules 2025, 30, 834. https://doi.org/10.3390/molecules30040834

AMA Style

Seidel A, Wagler J. Anions of α-Amino Acids as (O,N)-Donor Ligands in Si-, Ge- and Sn-Coordination Chemistry. Molecules. 2025; 30(4):834. https://doi.org/10.3390/molecules30040834

Chicago/Turabian Style

Seidel, Anne, and Jörg Wagler. 2025. "Anions of α-Amino Acids as (O,N)-Donor Ligands in Si-, Ge- and Sn-Coordination Chemistry" Molecules 30, no. 4: 834. https://doi.org/10.3390/molecules30040834

APA Style

Seidel, A., & Wagler, J. (2025). Anions of α-Amino Acids as (O,N)-Donor Ligands in Si-, Ge- and Sn-Coordination Chemistry. Molecules, 30(4), 834. https://doi.org/10.3390/molecules30040834

Article Metrics

Back to TopTop