Previous Article in Journal
Physical Vapor Deposition of Indium-Doped GeTe: Analyzing the Evaporation Process and Kinetics
Previous Article in Special Issue
Conformational, Electrochemical, and Antioxidative Properties of Conjugates of Different Ferrocene Turn-Inducing Scaffolds with Hydrophobic Amino Acids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thionitrosyl Complexes of Rhenium and Technetium with PPh3 and Chelating Ligands—Synthesis and Reactivity

Institute of Chemistry and Biochemistry, Freie Universität Berlin, Fabeckstr. 34/36, D-14195 Berlin, Germany
*
Author to whom correspondence should be addressed.
Inorganics 2024, 12(8), 210; https://doi.org/10.3390/inorganics12080210
Submission received: 10 July 2024 / Revised: 25 July 2024 / Accepted: 26 July 2024 / Published: 31 July 2024
(This article belongs to the Special Issue Metal Complexes Diversity: Synthesis, Conformations, and Bioactivity)

Abstract

:
In contrast to corresponding nitrosyl compounds, thionitrosyl complexes of rhenium and technetium are rare. Synthetic access to the thionitrosyl core is possible by two main approaches: (i) the treatment of corresponding nitrido complexes with S2Cl2 and (ii) by reaction of halide complexes with trithiazyl chloride. The first synthetic route was applied for the synthesis of novel rhenium and technetium thionitrosyls with the metals in their oxidation states “+1” and “+2”. [MVNCl2(PPh3)2], [MVNCl(PPh3)(LOMe)] and [MVINCl2(LOMe)] (M = Re, Tc; {LOMe} = (η5-cyclopentadienyl)tris(dimethyl phosphito-P)cobaltate(III)) complexes have been used as starting materials for the synthesis of [ReII(NS)Cl3(PPh3)2] (1), [ReII(NS)Cl3(PPh3)(OPPh3)] (2), [ReII(NS)Cl(PPh3)(LOMe)]+ (4a), [ReII(NS)Cl2(LOMe)] (5a), [TcII(NS)Cl(PPh3)(LOMe)]+ (4b) and [TcII(NS)Cl2(LOMe)] (5b). The triphenylphosphine complex 1 is partially suitable as a precursor for ongoing ligand exchange reactions and has been used for the synthesis of [ReI(NS)(PPh3)(Et2btu)2] (3a) (HEt2btu = N,N-diethyl-N′-benzoyl thiourea) containing two chelating benzoyl thioureato ligands. The novel compounds have been isolated in crystalline form and studied by X-ray diffraction and spectroscopic methods including IR, NMR and EPR spectroscopy and (where possible) mass spectrometry. A comparison of structurally related rhenium and technetium complexes allows for conclusions about similarities and differences in stability, reaction kinetics and redox behavior between these 4d and 5d transition metals.

Graphical Abstract

1. Introduction

The two “group 7” elements technetium and rhenium possess radioactive isotopes (99mTc, 186,188Re), which make them more than interesting for nuclear medical applications [1,2,3,4,5,6,7,8,9,10,11]. The metastable nuclide 99mTc (pure γ-emitter, half-life 6 h, γ-energy: 411 keV) is sometimes denoted as the “workhorse” of diagnostic nuclear medicine, since more than 80 percent of clinical applications in this field in more than 10,000 hospitals worldwide are performed with 99mTc-containg pharmaceuticals [12]. This dominating position is mainly due to the favorable nuclear properties of this nuclide: (i) an almost optimal γ-energy without considerable amounts of accompanying particle radiation, (ii) a half-life which is long enough to examine metabolic processes but short enough to avoid a considerable radiation burden to the patient and (iii) the formation of 99mTc in a so-called generator system from 99Mo [13], which ensures a permanent availability of this nuclide in the clinics. A similar generator system starting from 188W as mother nuclide is also available for the supply of 188Re (Eβ,max: 2.12 MeV, Eβ,av: 784 keV, half-life: 16.9 h), a beta-emitting isotope with potential in nuclear medical therapy [14]. For both isotopes, kit-like preparations of the pharmaceuticals have been developed, which allow for a rapid and cost-efficient production of the individual administrations [1,14].
99mTc and 188Re are also under discussion as an almost perfect example of a “matched pair” of nuclides for theragnostic applications [15,16], which is a modern approach in personalized medicine, where an optimized dosage of a therapeutic drug becomes possible by the parallel monitoring of the therapeutic effects by a diagnostic counterpart.
Although the syntheses of 99mTc and 188Re compounds proceed at an approximate nanomolar level due to the low (chemical) concentrations obtained from the corresponding nuclide generators, fundamental chemical and structural studies concerning the formed compounds are commonly conducted with natural rhenium and the long-lived technetium isotope 99Tc, a weak beta emitter (Emax = 297 keV, half-life 2.11 × 105 years), which is also available in macroscopic amounts. During such studies, occasionally at the first glance, unusual solutions for the development of novel radiopharmaceuticals have been found, including several organometallic approaches [17,18,19,20,21,22,23] and drugs on the basis of nitrido [24,25,26,27] or hydrazido compounds [28,29,30]. Finally, the usability of a potential solution is decided by the successful transfer to the low concentration level of 99mTc and the stability of the product under aqueous conditions. Consequently, the potential of a chemical solution for a final medical application cannot easily be predicted by a simple, intuitive evaluation of a substance or a substance class. A good example is “99mTc-sestamibi”, a cationic organotechnetium(I) complex with six isocyanide ligands. As a compound with six technetium-carbon bonds, it was most probably not in the first row for candidates for myocardial imaging. Nowadays, it is one of the most used radiopharmaceuticals worldwide and, up to now, more than 40 million patients have been examined with this drug [17]. Also, nitrosyl complexes of technetium were considered for such purposes and related 99mTc compounds have been tested for their biodistribution [31]. The related structural chemistry was performed with the long-lived isotope 99Tc and starting from these early days in the 1980s, our knowledge about nitrosyltechnetium compounds has significantly increased.
In contrast to the relatively large number of nitrosyl complexes of the elements rhenium [32,33,34,35,36,37,38,39,40] and technetium [40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60], there are only a few reports about thionitrosyls of these two elements [61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80]. This should be mainly due to the lack of a readily available monomeric nitrogen sulfide. Although the thionitrosyl radical has been described spectroscopically and some NS+ salts have been isolated [81,82,83,84], the synthesis of such compounds remains a challenge and they have only occasionally been used as precursors for the synthesis of thionitrosyl complexes [61,83,84]. More convenient are reactions starting from cyclic compounds such as tetrasulfur tetranitride, S4N4, or trithiazyltrichloride, (NSCl)3 (Figure 1), which readily undergo thermal or photochemical decomposition under release of NS+ or NSCl+ building blocks. But it should be noted that N4S4 is a potentially dangerous compound, which tends to explode on heating or upon shock. Thus, it is not a favored agent at least for reactions with the radioactive technetium. Finally, the most promising synthetic approach to thionitrosyl complexes is the addition of a “sulfur atom” to a nitrido ligand. A variety of “sulfur sources” have been used including elemental sulfur, SOCl2, dithionite or NCS ligands. Most successful, however, are reactions of nitrido complexes with S2Cl2, which give low-valent thionitrosyl species in good yields. About reactions following the latter approach, we report in the present paper starting from different rhenium and technetium complexes with the metals in the oxidation states “+5” or “+6”.

2. Results and Discussion

2.1. Triphenylphosphine Complexes

Treatment of the rhenium(V) complex [ReNCl2(PPh3)2] with an excess of disulfur dichloride in refluxing dichloromethane results in a subsequent dissolution of the sparingly soluble starting material and a dark-red, almost black solution is formed, from which the thionitrosyl complexes [Re(NS)Cl3(PPh3)2] (1) and [Re(NS)Cl3(PPh3)(OPPh3)] (2) can be isolated. The parallel formation of phosphine and phosphine oxide complexes is not completely surprising with respect to the oxidizing conditions in the reaction mixture and has been observed previously during similar reactions with [TcNCl2(PMe2Ph)3] [74]. The two products can readily be separated during the crystallization procedure since compound 2 is highly soluble in acetone, while complex 1 is almost insoluble in this solvent. During the slow evaporation of CH2Cl2/acetone solutions of mixtures of 1 and 2, [Re(NS)Cl3(PPh3)2] precipitates first as dark-red crystals, while the orange-yellow crystals of 2 do not deposit until almost all acetone is evaporated. The best method for the isolation of pure [Re(NS)Cl3(PPh3)2], however, is the addition of an excess of PPh3 during the synthesis, which obviously avoids the formation of significant amounts of the phosphine oxide complex (Scheme 1).
The infrared spectra of the complexes show the characteristic ν(NS) stretches at 1213 cm–1 (phosphine complex) and 1230 cm–1 (phosphine oxide complex). These values are in agreement with such recorded for other rhenium complexes and those calculated for similar chromium(I) species [62]. The difference of approximately 20 cm–1 reflects a stronger back-donation in compound 1 compared with 2. Since the oxidation state of rhenium in both compounds is “+2”, the observed difference indicates some changes in the coordination sphere of the transition metal. Particularly the electronic properties of the ligand in the trans position to the thionitrosyl ligand have a direct influence to the donor/acceptor behavior of NS+.
Single-crystal X-ray diffraction confirms that the oxidation of one of the PPh3 ligands is accompanied with a ligand rearrangement and the OPPh3 ligand in 2 occupies the position trans to NS+ instead of Cl in 1. The molecular structures of both complexes are depicted in Figure 2 and selected bond lengths and angles are summarized in Table 1. It becomes evident that the thionitrosyl units are arranged linearly in both compounds with Re–N–S angles between 167° and 178°. Such a bonding situation has been found for all hitherto structurally characterized thionitrosyl complexes. Also, the direction of a formed phosphine oxide ligand in the trans position to a multiply bonded ligand is not without precedence and has been observed before during the formation of [Tc(NS)Cl3(PMe2Ph)(OPMe2Ph) from cis,mer-[TcNCl2(PMe2Ph)3] [74]. More generally, the presence of one triphenylphosphine oxide ligand in the coordination sphere of a transition metal favors the cis coordination to PPh3, since the steric and electronic effects, which direct two PPh3 ligands commonly into trans positions to each other, are significantly lowered.
[Re(NS)Cl3(PPh3)2] (1) as well as [Re(NS)Cl3(PPh3)(OPPh3)] (2) contain rhenium in the oxidation state “+2”. Due to the corresponding 5d5 “low spin” configuration, they are paramagnetic with one unpaired electron. This allows for the detection of resolved EPR spectra in solution. Figure 3 shows the spectra of both compounds at various temperatures. The spectra in liquid solutions (Figure 3a) consist of six hyperfine lines due to interactions of the unpaired electron with 185,187Re, both having a nuclear spin of I = 5/2. Superhyperfine couplings due to interactions of the unpaired electron with the coordinating 31P (I = 1/2), 35,37Cl (I = 3/2) or 14N (I = 1) nuclei could not be resolved. The frozen solution EPR spectra, which are shown in Figure 3b, confirm the presence of essentially axially symmetric, randomly oriented S = ½ spin systems with resolved parallel and perpendicular sets of 185,187Re hyperfine lines. The analysis of the spectral parameters indicates the presence of a small rhombic component in the perpendicular part of the spectrum of complex 1, which has been taken into account during the simulation. The simulated spectra, which have been used to derive the spectral parameters, are depicted in the Supporting Material.
As already mentioned, couplings with the 31P nuclei of the coordinated phosphine ligands could not be resolved in the spectra of 1 and 2. This finding is in contrast to the previously studied technetium(II) complexes [Tc(NS)Cl3(PMe2Ph)2] and [Tc(NS)Cl3(PMe2Ph)(OPMe2Ph)], where the 99Tc hyperfine signals of the parallel part show splittings into triplets (phosphine complex) or doublets (phosphine/phosphine oxide complex) and provide direct information for the occupation of their respective equatorial coordination spheres [74,77,85,86]. The larger line-widths of the EPR spectra obtained for the rhenium complexes prevent the resolutions of such couplings. Nevertheless, a remarkable trend in the EPR spectral parameters should be mentioned: the isotropic 185,187Re coupling constants a0Re show a dependence on the number of chlorido ligands in the equatorial coordination sphere in a way that they increase from compound 1 (2 Cl: 342 × 10–4 cm–1) via compound 2 (3 Cl: 372 × 10–4 cm–1) to the [Re(NS)Cl4] anion (4 Cl: 457 × 10–4 cm–1 [66]). In the same sequence of complexes, the corresponding g0 values decrease. Similar trends are also observed for the g values and coupling constants of the anisotropic spectra, which reflects considerable changes in the MO of the unpaired electron mainly due to the presence or absence of phosphine ligands.

2.2. Complexes with Chelating Ligands

Since the triphenylphosphine complexes 1 and 2 might have potential as precursors for the synthesis of more thionitrosyl complexes of rhenium, we tested some ligand exchange reactions with the chelating ligands shown in Figure 4. HEt2btu is a potentially bidentate chelator. It readily deprotonates after the addition of a supporting base and stable complexes are known with almost all common transition metals. This also includes rhenium and technetium complexes with the metals in several oxidation states [87,88,89,90,91,92,93,94]. The tripodal, organometallic ligand {LOMe} belongs to a classical family of tripodal ligands, which has been designed by Wolfgang Kläui and tested for a plethora of various applications [95,96]. Also, a number of rhenium complexes with this ligand are known and it is the first example of a chelating ligand, which forms stable technetium complexes with the metal in seven different oxidation states [54,95,96,97,98,99,100,101].
Reactions of [Re(NS)Cl3(PPh3)2] with HEt2btu in a mixture of CH2Cl2 and EtOH proceeded at room temperature (Scheme 2). After the addition of a small amount of Et3N as a supporting base, the red reaction mixture immediately turned dark brown and a red-brown solid could be isolated. Single crystals suitable for X-ray diffraction were obtained by the slow evaporation of a CH2Cl2 solution. The deprotonation of HEt2btu and its coordination as an S,O chelate is strongly indicated by the IR spectrum of the product by the disappearance of the NH stretches of the ligands and a strong bathochromic shift of the νC=O band by almost 200 cm−1. This is typical for chelate-bonded R2btu ligands [87]. Reduction of the metal ion is frequently observed during reactions with sulfur-containing ligands and, thus, the product of the described reaction with HEt2btu is the diamagnetic rhenium(I) complex [Re(NS)(PPh3)(Et2btu)2] (3a). This allows for the measurement of NMR spectra. They support the composition of the product as is performed by the mass spectrum.
The results of the spectroscopic studies are confirmed by a single-crystal structure analysis of the product. Figure 5 depicts the molecular structure of compound 3a, which shows the expected linear arrangement of the thionitrosyl ligand (Table 2). The equatorial coordination sphere of rhenium is occupied by an S,O chelate, a PPh3 and the sulfur atom of the second Et2btu ligand. Interestingly, the two Re–O single bonds are almost equal, which indicates similar structural trans influences induced by the NS+ and the PPh3 ligands. The N–S distance in the rhenium(I) complex is markedly longer than in the rhenium(II) complexes 1 and 2, which might be a consequence of a stronger back-donation of the d6 metal ion into anti-bonding orbitals of the thionitrosyl ligand. Unfortunately, the corresponding IR stretch, which is normally more indicative, cannot be assigned unambiguously for 3a due to many overlapping bands in the respective range of the spectrum.
With respect to the clean reaction of [Re(NS)Cl3(PPh3)2] with HEt2btu and the ready formation of the thionitrosylrhenium(I) complex 3a, a surprising result was obtained for an analogous reaction of the corresponding technetium complex [Tc(NS)Cl3(PPh3)2]. It proceeded under complete removal of the NS+ ligand and oxidation of the metal ion. The resulting technetium(III) tris complex [Tc(Et2btu)3] (Scheme 2) has been reported before (i) as the product of the direct reduction of pertechnetate with SnCl2 in the presence of excess ligand [102], and (ii) by a subsequent reduction/ligand exchange procedure starting from (NBu4)[TcOCl4] [87]. The observed lability of the thionitrosyl unit is a considerable drawback for a potential use of 1 as precursor in ligand exchange procedures and is accompanied by another disappointing experience during attempted reactions with Na{LOMe}. Although reactions of an entire series of different technetium and rhenium complexes with the “Kläui ligand” give well-defined and stable products [95,96,97,98,99,100,101], attempted reactions with 1 at room temperature did not proceed. Heating of such reaction mixtures finally gave a small amount of the chelate [Re(NS)Cl(PPh3)(LOMe)], but the yield was low and the product was accompanied by a number of side-products and impurities, which precluded the isolation of a pure compound in reasonable yields. Consequently, we preferred the second general approach to thionitrosyl compounds: the reaction of nitrido complexes with S2Cl2. Corresponding starting materials with {LOMe} ligands, [MNCl(PPh3)(LOMe)] and [MNCl2(LOMe)] complexes with M = Re or Tc, are readily available from simple procedures, and the reactions with S2Cl2 (Scheme 3) give the products [M(NS)Cl(PPh3)(LOMe)]Cl (M = Re: 4a; M = Tc: 4b) and [M(NS)Cl2(LOMe)] (M = Re: 5a; M = Tc: 5b) in reasonable yields and good purities.
The reactions with all four starting complexes are restricted to the nitrido ligands and the remaining coordination spheres of the metals remain unchanged. Cationic thionitrosyls of rhenium(II) and technetium(II) are formed, when [MVNCl(PPh3)(LOMe)] complexes (M = Re, Tc) are used as precursors. The products can be precipitated from the reaction mixture as yellow (Re compound, 4aCl) and red (Tc complex, 4bCl) chloride salts by the addition of n-hexane. Microcrystalline samples were obtained after recrystallization of CH2Cl2/n-hexane mixtures. For the rhenium complex, the PF6 salt was prepared by metathesis with AgPF6. Orange-yellow needles of [Re(NS)Cl(PPh3)(LOMe)](PF6), 4a(PF6), were suitable for X-ray diffraction. The structure of the complex cation is shown in Figure 6a and selected bond lengths and angles are summarized in Table 3. Expectedly, the {LOMe} ligands act as tripods in both complexes, which directs the remaining three ligands into trans positions of their oxygen atoms. The M–N–S angles are close to 180° in both compounds.
Similar reaction patterns are observed for [ReNCl2((LOMe)] and [TcNCl2((LOMe)]. Both precursors contain the transition metals in their oxidation states “+6”, while the products [Re(NS)Cl2(LOMe)] (5a) and [Tc(NS)Cl2(LOMe)] (5b) are Re(II) and Tc(II) complexes (Scheme 3). An X-ray crystal structure determination has been performed for the technetium complex and its molecular structure is shown in Figure 6b. The general structural features found for the cation 4a also apply to technetium complex 5b (Table 3).
As compounds 1 and 2, the {LOMe} complexes of the present study, are paramagnetic with a d5 “low spin” configuration, this results in S = ½ spin systems and allows us to record resolved EPR spectra in liquid and frozen solutions. The spectra have essentially axial symmetry with well-resolved 185,187Re and 99Tc hyperfine couplings in their parallel and perpendicular parts. Figure 7 shows exemplarily the spectra obtained for [Tc(NS)Cl(PPh3)(LOMe)]Cl (4bCl).
99Tc has a nuclear spin of I = 9/2, which results in a ten-line pattern in the spectrum in liquid solution, while two sets of ten 99Tc hyperfine lines (each one in the parallel and one in the perpendicular part of the spectrum) are characteristic for the anisotropic frozen-solution spectrum as can be described by the Spin Hamiltonian (1), where g, g, AM. and AM are the principal values of the 99Tc and 185,187Re hyperfine tensors AM. The spectral parameters are summarized in Table 4.
H ^ s p = β e g B z S ^ z + g B x S ^ x + B y S ^ y + A M S ^ z I ^ z + A M ( S ^ x I ^ x + S ^ y I ^ y )
Superhyperfine interactions due to couplings of the unpaired electron with the 31P nuclei of the PPh3 ligands in compounds 4 are not visible. This prevents us from obtaining direct information about the extent of spin delocalization into the ligand orbitals. Indirect information, however, can be deduced from the 185,187Re coupling parameters. As described for compounds 1 and 2, also for the {LOMe} thionitrosyl complexes, the isotropic and anisotropic ATc and ARe coupling parameters decrease, when phosphine ligands are present in the equatorial coordination spheres of the metals instead of Cl. This can be understood by the transfer of electron density to the π-acceptor PPh3. The experimental line-widths of the spectra, however, do not allow for the resolution of such couplings for the complexes of the present study.

3. Materials and Methods

[ReNCl2(PPh3)2] [103], [TcNCl2(PPh3)2] [104], (NBu4)[TcNCl4] [105], (NBu4)[ReNCl4] [106], [ReNCl(PPh3)(LOMe)] [98], [TcNCl(PPh3)(LOMe)] [58], [ReNCl2(LOMe)] [107], [TcNCl2(LOMe)] [58], [Tc(NS)Cl3(PPh3)2] [108] and HEt2btu [109] were prepared according to literature procedures. All other chemicals were reagent-grade and purchased commercially. Reactions with air- or moisture-sensitive compounds were performed with the standard Schlenk technique. A laboratory approved for the handling of radioactive material was used for the synthesis of the technetium compounds. Normal glassware gives adequate protection against the weak beta radiation of 99Tc as long as small amounts of this isotope as in the present paper are used.

3.1. Syntheses

[Re(NS)Cl3(PPh3)2] (1): [ReNCl2(PPh3)2] (2 g, 2.6 mmol) was suspended in 40 mL CH2Cl2 and an excess of S2Cl2 (0.5 mL, 6.26 mmol) and PPh3 (680 mg, 5.2 mmol) was added. The mixture was heated under reflux for approximately 30 min, which resulted in a slow dissolution of the starting material and the formation of a dark-red solution. The progress of the reaction can be controlled by TLC. After a complete consumption of [ReNCl2(PPh3)2], the solvent and the remaining S2Cl2 were removed in vacuum. The resulting solid was crystallized from a CH2Cl2/acetone (1:1) mixture giving dark-red crystals. Yield: 1.49 g (66%). Elemental analysis: Calcd. For C36H30Cl3NP2ReS: C, 50.1; H, 3.5; N, 1.6; S, 3.7%. Found: C 49.8; H, 3.5; N, 1.6; S, 4.0%. IR (ATR, cm–1): 3057 (w), 1981 (w), 1670 (w), 1586 (w), 1574 (w), 1482 (s), 1434 (s), 1316 (m), 1213 (s), 1187 (s), 1159 (m), 1089 (s), 1074 (m), 1028 (m), 998 (m), 928 (w), 855 (w), 758 (m), 744 (s), 711 (m), 692 (s), 617 (w). EPR (77 K, CH2Cl2): gx = 2.110, gy = 1.991, gz = 1.943; AxRe = 281 × 10–4 cm–1, AyRe = 283 × 10–4 cm–1, AzRe = 524 × 10–4 cm–1. EPR (RT, CH2Cl2): g0 = 2.025; a0Re = 342 × 10–4 cm–1. ESI+ MS (m/z): 885.002 [M+Na]+ (calcd. 885.009), 900.975 [M+K]+ (calcd. 900.983).
[Re(NS)Cl3(PPh3)(OPPh3)] (2): The product was obtained as a side product of the synthesis of [Re(NS)Cl3(PPh3)2] when no additional PPh3 was added. It could be isolated as an orange-red solid following compound 1 during the crystallization procedure. Yield: 0.6 g (26%). IR (ATR, cm–1): 3056 (w), 1589 (w), 1482 (m), 1435 (s), 1231 (s), 1188 (w), 1140 (vs), 1119 (vs), 1026 (m), 998 (m), 924 (w), 852 (w), 747 (s), 723 (vs), 690 (vs). EPR (77 K, acetone): g = 1.923, g = 2.008; ARe = 594 × 10–4 cm–1, ARe = 293 × 10–4 cm–1. EPR (RT, acetone): g0 = 2.017; a0Re = 372 × 10–4 cm–1.
[Re(NS)(PPh3)(Et2btu)2] (3a): [Re(NS)Cl3(PPh3)2] (100 mg, 0.1 mmol) was dissolved in 30 mL CH2Cl2 and a solution of HEt2btu (60 mg, 0.25 mmol) in 20 mL Ethanol containing 3 drops of triethylamine was added. The mixture was stirred at room temperature for 30 min. During this time, the color changed to pale brown. The volatiles were removed in vacuum and the remaining solid was subsequently washed with cold methanol, diethyl ether and n-hexane. The raw product was dissolved in CH2Cl2 for crystallization. Slow evaporation of the solvent gave red-brown crystals. Yield: 31 mg (25%). Elemental analysis: Calcd. for C42H45N5O2PReS3: C, 52.3; H, 4.7; N, 7.3; S, 10.0%. Found: C 52.0; H, 4.9; N, 6.8; S, 10.2%. IR (ATR, cm–1): 3058 (w), 2972 (w), 2927 (w), 1599 (w), 1587 (w), 1528 (m), 1499 (s), 1446 (m), 1416 (s), 1394 (s), 1353 (s),1311 (m), 1294 (m), 1251 (s),1210 (w), 1140 (m), 1092 (m), 1071 (m), 1025 (m), 997 (m), 895 (m), 883 (m), 827 (m), 797 (m), 752 (w), 742 (m), 702 (s), 692 (s), 671 (m), 661 (m), 619 (w). 1H NMR (400 MHz, CDCl3, ppm): 8.10 (d, J = 7.6 Hz, 2H, Ph), 7.65 (d, J = 7.3 Hz, 6H, Ph), 7.37–7.13 (m, 15H), 7.01 (t, J = 7.9 Hz, 2H, Ph), 4.37 (dq, J = 14.2, 7.1 Hz, 1H, CH2), 4.14–3.80 (m, 4H, CH2), 3.73 (q, J = 7.3, 6.8 Hz, 1H, CH2), 3.62 (q, J = 6.3 Hz, 1H, CH2), 3.51 (dd, J = 12.9, 6.5 Hz, 1H, CH2), 1.52 (s, 2H), 1.39 (dd, J = 9.6, 4.2 Hz, 3H, CH3), 1.27 (q, J = 12.3, 9.7 Hz, 3H, CH3), 1.03 (t, J = 6.5 Hz, 3H, CH3), 0.90 (d, J = 6.3 Hz, 3H, CH3). J = 8.0 Hz, 12H, CH3). 31P NMR (CDCl3, ppm): 21.67 (s). ESI+ MS (m/z): = 965.237 [M]+ (calcd. 965.203) 988.225 [M+Na]+ (calcd. 988.192), 1004.202 [M+K]+ (calcd. 1004.166).
[Re(NS)Cl(PPh3)(LOMe)]Cl (4aCl): [ReNCl(PPh3)(LOMe)] (250 mg, 0.25 mmol) was suspended in 30 mL CH2Cl2, an excess of S2Cl2 (0.1 mL) was added and the mixture was heated under reflux for 30 min. The solvent was reduced to 15 mL and n-hexane (150 mL) was added. The resulting precipitate was washed with hexane and diethyl ether and dried. Recrystallization from a CH2Cl2/n-hexane (2:1) mixture gave a pale-yellow solid. Yield: 211 mg (83%). Elemental analysis: Calcd. for C29H38Cl2CoNO9P4ReS: C, 34.3; H, 3.8; N, 1.4; S, 3.2%. Found: C 34.0; H, 4.0; N, 1.4; S, 3.1%. IR (ATR, cm–1): 3056 (w), 2992 (w), 2948 (w), 1634 (w), 1483 (w), 1457 (w), 1436 (m), 1223 (m), 1177 (m), 1036 (vs), 1014 (vs), 943 (s), 912 (s), 853 (s), 793 (s), 746 (vs), 693 (s), 599 (vs). EPR (77 K, CH2Cl2): g = 1.842, g = 1.986; ARe = 644 × 10–4 cm–1, ARe = 330 × 10–4 cm–1. EPR (RT, CHCl3): g0 = 1.995; a0Re = 420 × 10–4 cm–1. ESI+ MS (m/z): 981.013 [M]+ (calcd. 980.979).
[Re(NS)Cl(PPh3)(LOMe)](PF6) (4a(PF6)): [Re(NS)Cl(PPh3)(LOMe)]Cl (50 mg, 0.05 mmol) was dissolved in acetonitrile and an excess of AgPF6 (19 mg, 0.075 mmol) was added. The solution was stirred at room temperature and the formed colorless precipitate was removed by filtration. Orange-yellow needles suitable for X-ray diffraction formed upon the slow evaporation of a solution of the complex in CH2Cl2/n-hexane. Yield: 52 mg (93%). The essential spectroscopic data are identical with those of 4aCl.
[Re(NS)Cl2(LOMe)] (5a): [ReNCl2(LOMe)] (34 mg, 0.05 mmol) was dissolved in 5 mL CH2Cl2 and 3 drops of S2Cl2 were added. The mixture was stirred at room temperature. The progress of the reaction can readily be monitored by EPR spectroscopy and typically after 15 min the signals of the starting material have disappeared. The solvent and residual S2Cl2 were removed in vacuum and the solid residue was washed subsequently with n-hexane and diethyl ether. The product was recrystallized from CH2Cl2/diethyl ether giving a pale-brown, microcrystalline solid. Yield: 20 mg (53%). IR (ATR, cm–1): 3111 (w), 2958 (m), 2874 (w), 1460 (w), 1425 (w), 1381 (w), 1260 (w), 1222 (s), 1175 (s), 1012 (vs), 945 (w), 860 (m), 796 (vs), 743 (s), 595 (m). EPR (77 K, CH2Cl2): g = 1.945, g = 1.760; ARe = 787 × 10–4 cm–1, ARe = 398 × 10–4 cm–1. EPR (RT, CHCl3): g0 = 1.990; a0Re = 455 × 10–4 cm–1. ESI+ MS (m/z): 776.846 [M+Na]+ (ber.: 776.847); 792.821 [M+K]+ (ber.: 792.821).
[Tc(NS)Cl(PPh3)(LOMe)]Cl (4bCl): [TcNCl(PPh3)(LOMe)] (72 mg, 0.08 mmol) was dissolved in 2 mL CH2Cl2 (2 mL) and S2Cl2 (11 mg, 0.08 mmol) was added under permanent stirring. The stirring was continued at room temperature for 1 h and the obtained dark-red mixture was filtered to remove some insoluble solid (most probably elemental sulfur). All volatiles were removed in vacuum and the remaining solid crystallized from a CH2Cl2/n-pentane (1:1) mixture. Storing such a mixture in a refrigerator for several days gave a red, microcrystalline powder. Yield: 11 mg (15%). IR (KBr, cm–1): 2910 (vs), 2723 (s), 1726 (m), 1630 (w) 1479 (vs), 1381 (vs), 1227 (vs), 1171 (m), 1040 (w), 949 (s), 928 (s), 851 (w), 802 (w), 723 (m), 610 (w), 542 (m). EPR (77 K, CH2Cl2): g = 1.966, g = 2.023; ATc = 252 × 10–4 cm–1, ATc = 112 × 10–4 cm–1. EPR (RT, CH2Cl2): g0 = 2.007; a0Tc = 168 × 10–4 cm–1.
[Tc(NS)Cl2(LOMe)] (5b): [TcNCl2(LOMe)] (41 mg, 0.06 mmol) was dissolved in CH2Cl2 (10 mL) and treated with an excess of S2Cl2 (0.3 mL). After stirring for 30 min at room temperature, all volatiles were removed in vacuum. The remaining solid was washed with n-hexane and diethyl ether and dissolved in 1 mL CH2Cl2. Overlayering with n-hexane and storing in a refrigerator for several days gave pale-red crystals. Yield: 32 mg (75%). IR (KBr, cm–1): 3107 (w), 3086 (w), 2990 (w), 2841 (w), 1462 (w), 1423 (w), 1310 (vs), 1236 (vs), 1182 (vs), 1128 (vs), 1098 (vs), 986 (vs), 847 (m), 795 (s), 747 (s), 638 (m), 608 (s), 530 (w), 476 (w), 426 (w). EPR (77 K, CH2Cl2): g = 1.943, g = 2.021, ATc = 285 × 10–4 cm–1, ATc = 129 × 10–4 cm–1. EPR (RT, CH2Cl2) g0 = 1.996; a0Tc = 177 × 10–4 cm–1.

3.2. Spectroscopic and Analytical Methods

Elemental analyses of carbon, hydrogen, nitrogen and sulfur were determined using a Heraeus vario EL elemental analyzer. IR spectra were measured as KBr pellets on a Shimadzu IR Affinity-1 (technetium compounds) or a Thermo Scientific Nicolet iS10 ATR spectrometer (rhenium complexes). The NMR spectra were recorded on JEOL ECS-400 or ECZ-400 400 MHz spectrometers. ESI TOF mass spectra were measured with an Agilent 6210 ESI TOF (Agilent Technologies, Santa Clara, CA, USA). X-Band EPR spectra were recorded in solution with a Magnettech Miniscope MS400 spectrometer at 300 and 77 K. Simulation and visualization of the EPR spectra were conducted with the EasySpin tool box in MatLab [110,111].

3.3. X-ray Crystallography

The intensities for the X-ray determinations were collected on STOE IPDS-2T or Bruker CCD instruments with Mo/Kα radiation. The various temperatures applied are due to the experimental setup of the different diffractometers. Semi-empirical or numerical absorption corrections were carried out by the SADABS or X-RED32 programs [112,113]. Structure solution and refinement were performed with the SHELX programs [114,115] included in the WinGX [116] program package or OLEX2 [117]. Hydrogen atoms were calculated for idealized positions and treated with the “riding model” option of SHELXL. The solvent mask option of OLEX2 was applied to treat diffuse electron density due to disordered CH2Cl2 in the crystals of [Re(NS)Cl3(PPh3)(OPPh3)] (2). Details are given in the Supplementary Materials. The representation of molecular structures was conducted using the program DIAMOND [118].

4. Conclusions

Reactions of rhenium and technetium nitrido compounds with disulfur dichloride give access to novel thionitrosyl complexes of these elements having monodentate and chelating ligands in their coordination spheres. The new compounds are chemically related to corresponding nitrosyl complexes but do not possess their robustness and stability. Their thermal instability and the lack of a suitable one-step synthesis starting from pertechnetate set narrow limits for potential applications in nuclear medical procedures, as is discussed in the introduction.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/inorganics12080210/s1, Table S1.1: Crystallographic data and data collection parameters; Figure S1.1: Ellipsoid representation of [Re(NS)Cl3(PPh3)2] (1). The thermal ellipsoids are set at a 50% probability level. Hydrogen atoms are omitted for clarity; Table S1.2: Selected bond lengths (Å) and angles (°) in the [Re(NS)Cl3(PPh3)2] (1); Figure S1.2: Ellipsoid representation of two crystallographically independent molecules of [Re(NS)Cl3(PPh3)(OPPh3)] (2). The thermal ellipsoids are set at a 50% probability level. Hydrogen atoms are omitted for clarity; Table S1.3.: Selected bond lengths (Å) and angles (°) in [Re(NS)Cl3(PPh3)(OPPh3)] (2); Figure S1.5: Ellipsoid representation of [Re(NS)(PPh3)(Et2btu)2] (3a). The thermal ellipsoids are set at a 35% probability level. Hydrogen atoms are omitted for clarity; Table S1.6: Selected bond lengths (Å) and angles (°) in [Re(NS)(PPh3)(Et2btu)2] (3a); Figure S1.3: Ellipsoid representation of [Re(NS)Cl(PPh3)(LOMe)] (PF6) (4a(PF6)) including the positional disorder between the Re–N–S and Re–Cl bonds. The thermal ellipsoids are set at a 50% probability level. Hydrogen atoms are omitted for clarity; Table S1.4: Selected bond lengths (Å) and angles (°) in the [Re(NS)Cl(PPh3)(LOMe)]+ (4a) cation; Figure S1.4: Ellipsoid representation of two crystallographically independent molecules of [Tc(NS)Cl2(LOMe)] (5b) including the positional disorder for the O atoms of the {LOMe} ligands and between the Tc–N–S and the Tc–Cl bonds. The thermal ellipsoids are set at a 50% probability level; Table S1.5: Selected bond lengths (Å) and angles (°) in [Tc(NS)Cl2(LOMe)] (5b); Figure S2.1: IR (ATR) spectrum of [Re(NS)Cl3(PPh3)2] (1); Figure S2.2: Solution EPR spectra of [Re(NS)Cl3(PPh3)2] (1) in CH2Cl2, (a) at room temperature and (b) at T = 77 K; Figure S2.3: ESI(+) mass spectrum of [Re(NS)Cl3(PPh3)2] (1) in CH3CN; Figure S2.4: IR spectrum (ATR) of [Re(NS)Cl3(PPh3)(OPPh3)] (2); Figure S2.5: Solution EPR spectra of [Re(NS)Cl3(PPh3)(OPPh3)] (2) in acetone; (a) at room temperature and (b) at T = 77 K; Figure S2.6: ESI(+) mass spectrum of [Re(NS)Cl3(PPh3)(OPPh3)] (2) in CH3CN; Figure S2.7: IR spectrum (ATR) of [Re(NS)(PPh3)(Et2btu)2] (3a); Figure S2.8: 1H NMR spectrum of [Re(NS)(PPh3)(Et2btu)2] (3a) in CDCl3; Figure S2.9: 31P NMR spectrum of [Re(NS)(PPh3)(Et2btu)2] (3a) in CDCl3; Figure S2.10: ESI(+) mass spectrum [Re(NS)(PPh3)(Et2btu)2] (3a) in CH3CN; Figure S2.11: IR spectrum (ATR) of [Re(NS)Cl(PPh3)(LOMe)]Cl (4aCl); Figure S2.12: Solution EPR spectra of [Re(NS)Cl(PPh3)(LOMe)]Cl (4aCl) in CH2Cl2; (a) at room temperature and (b) at T = 77 K; Figure S2.13: ESI(+) mass spectrum of [Re(NS)Cl(PPh3)(LOMe)]Cl (4aCl) in CH3CN; Figure S2.14: IR spectrum (ATR) of [Re(NS)Cl(PPh3)(LOMe)](PF6) (4a(PF6)); Figure S2.15: ESI(+) mass spectrum of [Re(NS)Cl(PPh3)(LOMe)](PF6) (4a(PF6)) in CH3CN; Figure S2.16: IR spectrum (ATR) of [Re(NS)Cl2(LOMe)] (5a); Figure S2.17: Solution EPR spectra of [Re(NS)Cl2(LOMe)] (5a) in CH2Cl2; (a) at room temperature and (b) at T = 77 K; Figure S2.18: ESI(+) mass spectrum of [Re(NS)Cl(PPh3)(LOMe)](PF6) (4a(PF6)) in CH3CN; Figure S2.19: IR spectrum (KBr) of [Tc(NS)Cl(PPh3)(LOMe)]Cl (4bCl); Figure S2.20: Solution EPR spectra of [Tc(NS)Cl(PPh3)(LOMe)]Cl (4bCl) in CH2Cl2; (a) at room temperature and (b) at T = 77 K; Figure S2.21: IR spectrum (KBr) of [Tc(NS)Cl2(LOMe)] (5b); Figure S2.22: Solution EPR spectra of [Tc(NS)Cl2(LOMe)] (5b). in CH2Cl2; (a) at room temperature and (b) at T = 77 K.

Author Contributions

Conceptualization, U.A. and D.N.; methodology, D.N. and A.H.; validation, A.H., T.E.S. and D.N.; formal analysis, D.N. and T.E.S.; investigation, D.N.; resources, U.A.; data curation, D.N. and A.H.; writing—original draft preparation, U.A.; writing—review and editing, D.N.; A.H. and T.E.S.; visualization, D.N. and U.A.; supervision, U.A.; project administration, U.A.; funding acquisition, U.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Freie Universität Berlin and the Deutsche Forschungsgemeinschaft (Core Facility BIOSUPRAMOL).

Data Availability Statement

Additional data are contained within the article and Supplementary Materials.

Acknowledgments

We would like to acknowledge the assistance of the Core Facility BioSupraMol supported by the DFG.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Technetium-99m Radiopharmaceuticals: Manufacture of Kits; IAEA Technical Reports Series No. 466; International Atomic Energy Agency: Vienna, Austria, 2008; pp. 126–129.
  2. Abram, U.; Alberto, R. Technetium and rhenium—coordination chemistry and nuclear medical applications. J. Braz. Chem. Soc. 2006, 17, 1486–1500. [Google Scholar] [CrossRef]
  3. Bartholomä, M.D.; Louie, A.S.; Valliant, J.F.; Zubieta, J. Technetium and gallium derived radiopharmaceuticals: Comparing and contrasting the chemistry of two important radiometals for the molecular imaging era. Chem. Rev. 2010, 110, 2903–2920. [Google Scholar] [CrossRef] [PubMed]
  4. Bhattacharyya, S.; Dixit, M. Metallic radionuclides in the development of diagnostic and therapeutic radiopharmaceuticals. Dalton Trans. 2011, 40, 6112–6128. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, S.; Chakraborty, S. 99mTc-centered one-pot synthesis for preparation of 99mTc radiotracers. Dalton Trans. 2011, 40, 6077–6086. [Google Scholar] [CrossRef] [PubMed]
  6. Dilworth, J.R.; Pascu, S.I. The Radiopharmaceutical Chemistry of Technetium and Rhenium. In The Chemistry of Molecular Imaging; Long, N., Wong, W.-T., Eds.; John Wiley & Sons: Chichester, UK, 2015; pp. 163–174. [Google Scholar]
  7. Kuntic, V.; Brboric, J.; Vujic, Z.; Uskokovic-Markovic, S. Radioisotopes Used as Radiotracers in vitro and in vivo Diagnostics: A Review. Asian J. Chem. 2016, 28, 235–411. [Google Scholar] [CrossRef]
  8. Papagiannopoulou, D. Technetium-99m radiochemistry for pharmaceutical applications. J. Labelled Compd. Radiopharm. 2017, 60, 502–520. [Google Scholar] [CrossRef] [PubMed]
  9. Duatti, A. Review on 99mTc radiopharmaceuticals with emphasis on new advancements. Nucl. Med. Biol. 2021, 92, 202–216. [Google Scholar] [CrossRef] [PubMed]
  10. Alberto, R.; Nadeem, Q. 99mTechnetium-Based Imaging Agents and Developments in 99Tc Chemistry. In Metal Ions in Bio-Imaging Techniques; Sigel, A., Freisinger, E., Sigel, K.O., Eds.; De Gruyter: Berlin/Munich, Germany; Boston, MA, USA, 2021; pp. 196–238. [Google Scholar]
  11. Alberto, R. Role of Pure Technetium Chemistry: Are There Still Links to Applications in Imaging? Inorg. Chem. 2023, 62, 20539–20548. [Google Scholar] [CrossRef] [PubMed]
  12. World Nuclear Association. Radioisotopes in Medicine. Available online: https://world-nuclear.org/information-library/non-power-nuclear-applications/radioisotopes-research/radioisotopes-in-medicine (accessed on 1 July 2024).
  13. Dash, A.; Knapp, F.F., Jr.; Pillai, M.R.A. 99Mo/99mTc separation: An assessment of technology options. Nucl. Med. Biol. 2013, 40, 167–176. [Google Scholar] [CrossRef]
  14. Lepareur, N.; Lacœuille, F.; Bouvry, C.; Hindré, F.; Garcion, E.; Chérel, M.; Noiret, N.; Garin, E.; Knapp, F.F. Rhenium-188 Labeled Radiopharmaceuticals: Current Clinical Applications in Oncology and Promising Perspectives. Front. Med. 2019, 6, 132. [Google Scholar] [CrossRef]
  15. Cutler, C.S.; Hennkens, H.M.; Sisay, N.; Huclier-Markai, S.; Jurisson, S. Radiometals for Combined Imaging and Therapy. Chem. Rev. 2013, 113, 858–883. [Google Scholar] [CrossRef]
  16. Price, E.W.; Orvig, C. Matching chelators to radiometals for radiopharmaceuticals. Chem. Soc. Rev. 2014, 43, 260–290. [Google Scholar] [CrossRef]
  17. Kronauge, J.F.K.; Mindiola, D.J. The value of Stable Metal-Carbon Bonds in Nuclear Medicine and the Cardiolite Story. Organometallics 2016, 35, 3432–3435. [Google Scholar] [CrossRef]
  18. Morais, G.R.; Paulo, A.; Santos, I. Organometallic Complexes for SPECT Imaging and/or Radionuclide Therapy. Organometallics 2012, 31, 5693–5714. [Google Scholar] [CrossRef]
  19. Alberto, R. From oxo to carbonyl and arene complexes; A journey through technetium chemistry. J. Organomet. Chem. 2018, 869, 264–269. [Google Scholar] [CrossRef]
  20. Benz, M.; Braband, H.; Schmutz, P.; Halter, J.; Alberto, R. From TcVII to TcI; facile syntheses of bis-arene complexes [99(m)Tc(arene)2]+ from pertechnetate. Chem. Sci. 2015, 6, 165–169. [Google Scholar] [CrossRef]
  21. Meola, G.; Braband, H.; Jordi, S.; Fox, T.; Blacque, O.; Spingler, B.; Alberto, R. Structure and reactivities of rhenium and technetium bis-arene sandwich complexes [M(η6-arene)2]+. Dalton Trans. 2017, 46, 14631–14637. [Google Scholar] [CrossRef]
  22. Nadeem, Q.; Meola, G.; Braband, H.; Bollinger, R.; Blancque, O.; Hernandez-Valdes, D.; Alberto, R. To Sandwich Technetium: Highly Functionalized Bis-Arene Complexes [99mTc(η6-arene)2]+ Directly from Water and [TcO4]. Angew. Chem. Int. Ed. Engl. 2020, 59, 1197–1200. [Google Scholar] [CrossRef] [PubMed]
  23. Nadeem, Q.; Battistin, F.; Blancque, O.; Alberto, R. Naphtalene Exchange in [Re(η6-napht)2]+ with Pharmaceutical Leads to Highly Functionalized Sandwich Complexes [M(η6-pharm)2]+ (M = Re/99mTc). Chemistry 2022, 28, e202103566. [Google Scholar] [CrossRef]
  24. Boschi, A.; Uccelli, L.; Marvelli, L.; Cittanti, C.; Giganti, M.; Martini, P. Technetium-99m Radiopharmaceuticals for Ideal Myocardial Perfusion Imaging: Lost and Found Opportunities. Molecules 2022, 27, 1188. [Google Scholar] [CrossRef]
  25. Pasqualini, R.; Duatti, A.; Bellande, E.; Comazzi, V.; Brucato, V.; Hoffschir, D.; Fagret, D.; Comet, M. Bis (Dithiocarbamato) Nitrido Technetium-99m Radiopharmaceuticals: A Class of Neutral Myocardial Imaging Agents. J. Nucl. Med. 1994, 35, 334–341. [Google Scholar]
  26. Boschi, A.; Uccelli, L.; Bolzati, C.; Duatti, A.; Sabba, N.; Moretti, E.; di Domenico, G.; Zavattini, G.; Refosco, F.; Giganti, M. Synthesis and Biologic Evaluation of Monocationic Asymmetric 99mTc-Nitride Heterocomplexes Showing High Heart Uptake and Improved Imaging Properties. J. Nucl. Med. 2003, 44, 806–814. [Google Scholar] [PubMed]
  27. Salvarese, N.; Carta, D.; Marzano, C.; Gerardi, G.; Melendez-Alafort, L.; Bolzati, C. [99mTc][Tc(N)(DASD)(PNPn)]+ (DASD = 1,4-Dioxa-8-Azaspiro[4,5]Decandithiocarbamate, PNP n = Bisphosphinoamine) for Myocardial Imaging: Synthesis, Pharmacological and Pharmacokinetic Studies. J. Med. Chem. 2018, 61, 11114–11126. [Google Scholar] [CrossRef] [PubMed]
  28. Meszaros, L.K.; Dose, A.; Biagini, S.C.G.; Blower, P.J. Hydrazinonicotinic acid (HYNIC)—Coordination chemistry and applications in radiopharmaceutical chemistry. Inorg. Chim. Acta 2010, 363, 1059–1069. [Google Scholar] [CrossRef]
  29. Abrams, M.J.; Juweid, M.; tenKate, C.I.; Schwartz, D.A.; Hauser, M.M.; Gaul, F.E.; Fuccello, A.J.; Rubin, R.H.; Strauss, H.W.; Fischman, A.J. Technetium-99m-human polyclonal IgG radiolabeled via the hydrazine nicotinamide derivative for imaging focal sites of infection in rats. J. Nucl. Med. 1990, 31, 2022–2028. [Google Scholar] [PubMed]
  30. Jiang, Y.; Tian, Y.; Feng, B.; Zhao, T.; Yu, X.; Zhao, Q. A novel molecular imaging probe [99mTc]Tc-HYNIC-FAPI targeting cancer-associated fibroblasts. Sci. Rep. 2023, 13, 3700. [Google Scholar] [CrossRef] [PubMed]
  31. Cheah, C.T.; Newman, J.L.; Nowotnik, D.P.; Thornback, J.R. Synthesis and biological studies of the [99mTc]tetrachloronitrosyltechnetium(II) anion—An alternative low valent technetium starting material. Int. J. Rad. Appl. Instrum. Nucl. Med. Biol. 1987, 14, 573–575. [Google Scholar] [CrossRef] [PubMed]
  32. Machura, B. Structural and spectroscopic properties of rhenium nitrosyl complexes. Coord. Chem. Rev. 2005, 249, 2277–2307. [Google Scholar] [CrossRef]
  33. Dilworth, J.R. Rhenium chemistry—Then and Now. Coord. Chem. Rev. 2021, 436, 213822. [Google Scholar] [CrossRef]
  34. Mahmood, A.; Akgun, Z.; Peng, Y.; Mueller, P.; Jiang, Y.; Berke, H.; Jones, A.G.; Nicholson, T. The synthesis and characterization of rhenium nitrosyl complexes. The X-ray crystal structures of [ReBr2(NO)(NCMe)3], [Re(NO)(N5)](BPh4)2] and [ReBr2(NO)(NCMe){py-CH2-NHCH2CH2-N(CH2-py)2}]. Inorg. Chim. Acta 2013, 405, 455–460. [Google Scholar] [CrossRef]
  35. Abram, U.; Ortner, K.; Hübener, R.; Voigt, A.; Caballho, R.; Vazquez-Lopez, E. Darstellung, Strukturen und EPR-Spektren der Rhenium(II)-Nitrosylkomplexe [Re(NO)Cl2(PPh3)(OPPh3)(OReO3)], [Re(NO)Cl2(OPPh3)2(OReO3)] und [Re(NO)Cl2(PPh3)3][ReO4]. Z. Anorg. Allg. Chem. 1998, 624, 1662–1668. [Google Scholar] [CrossRef]
  36. Agbossou, F.; O’Connor, E.J.; Garner, C.M.; Quiros Mendez, N.; Fernandez, J.M.; Patton, A.T.; Ramsden, J.A.; Gladysz, J.A. Cyclopentadienyl Rhenium Complexes. Inorg. Synth. 1992, 29, 211–225. [Google Scholar]
  37. Seidel, S.N.; Prommesberger, M.; Eichenseher, S.; Meyer, O.; Hampel, F.; Gladysz, J.A. Syntheses and structural analyses of chiral rhenium containing amines of the formula (η5-C5H5)Re(NO)(PPh3)((CH2)nNRR′) (n = 0, 1). Inorg. Chim. Acta 2010, 363, 533–548. [Google Scholar] [CrossRef]
  38. Bernasconi, C.F.; Bhattacharya, S.; Wenzel, P.J.; Olmstead, M.M. Kinetic and Thermodynamic Acidity of [Cp(NO)(PPh3)Re(2,5-dimethyl-3-thienyl)carbene]+. Transition State Imbalance and Intrinsic Barriers. Organometallics 2006, 25, 4322–4330. [Google Scholar] [CrossRef]
  39. Dilsky, S.; Schenk, W.A. Diastereomeric Halfsandwich Rhenium Complexes Containing Hemilabile Phosphane Ligands. Eur. J. Inorg. Chem. 2004, 2004, 4859–4870. [Google Scholar] [CrossRef]
  40. Nicholson, T.; Chun, E.; Mahmood, A.; Mueller, P.; Davison, A.; Jones, A.G. Synthesis, spectroscopy and structural analysis of Technetium and Rhenium nitrosyl complexes. Commun. Inorg. Chem. 2015, 3, 31–39. [Google Scholar]
  41. Blanchard, S.S.; Nicholson, T.; Davison, A.; Davis, W.; Jones, A.G. The synthesis, characterization and substitution reactions of the mixed technetium(I) nitrosyl complex trans-trans-[(NO)(NCCH3)Cl2(PPh3)2Tc]. Inorg. Chim. Acta 1996, 244, 121–130. [Google Scholar] [CrossRef]
  42. Balasekaran, S.M.; Hagenbach, A.; Drees, M.; Abram, U. [TcII(NO)(trifluoroacetate)4F]2−—Synthesis and reactions. Dalton Trans. 2017, 46, 13544–13552. [Google Scholar] [CrossRef] [PubMed]
  43. Linder, K.E.; Davison, A.; Dewan, J.C.; Costello, C.E.; Melaknia, S. Nitrosyl complexes of technetium: Synthesis and characterization of [TcI(NO)(CNCMe3)5](PF6)2 and Tc(NO)Br2(CNCMe3)3 and the crystal structure of Tc(NO)Br2(CNCMe3)3. Inorg. Chem. 1986, 25, 2085–2089. [Google Scholar] [CrossRef]
  44. Ackermann, J.; Noufele, C.N.; Hagenbach, A.; Abram, U. Nitrosyltechnetium(I) Complexes with 2-(Diphenylphosphanyl)aniline. Z. Anorg. Allg. Chem. 2019, 645, 8–13. [Google Scholar] [CrossRef]
  45. Ackermann, J.; Hagenbach, A.; Abram, U. {Tc(NO)(Cp)(PPh3)}+—A novel technetium(I) core. Chem. Commun. 2016, 52, 10285–10288. [Google Scholar] [CrossRef]
  46. Ackermann, J.; Abdulkader, A.; Scholtysik, C.; Jungfer, M.R.; Hagenbach, A.; Abram, U. [TcI(NO)X(Cp)(PPh3)] Complexes (X = I, I3, SCN, CF3SO3, or CF3COO) and Their Reactions. Organometallics 2019, 38, 4471–4478. [Google Scholar] [CrossRef]
  47. Abdulkader, A.; Hagenbach, A.; Abram, U. [Tc(NO)Cl(Cp)(PPh3)]—A Technetium(I) Compound with an Unexpected Synthetic Potential. Eur. J. Inorg. Chem. 2021, 2021, 3812–3818. [Google Scholar]
  48. Schibli, R.; Marti, N.; Maurer, P.; Spingler, B.; Lehaire, M.-L.; Gramlich, V.; Barnes, C.L. Syntheses and Characterization of Dicarbonyl–Nitrosyl Complexes of Technetium(I) and Rhenium(I) in Aqueous Media:  Spectroscopic, Structural, and DFT Analyses. Inorg. Chem. 2005, 44, 683–690. [Google Scholar] [CrossRef]
  49. Brown, D.S.; Newman, J.L.; Thornback, J.R.; Davison, A. Structure of the tetra-n-butylammoium salt of tetrachloro(methanol)nitrosyltechnetium(II) anion. Acta Cryst. 1987, C43, 1692–1694. [Google Scholar]
  50. Brown, D.S.; Newman, J.L.; Thornback, J.R.; Pearlstein, R.M.; Davison, A.; Lawson, A. The synthesis and characterisation of the trichloronitrosyl(acetylacetonato)technetium(II) anion, a novel technetium(II) complex. Inorg. Chim. Acta 1988, 150, 193–196. [Google Scholar] [CrossRef]
  51. Nicholson, T.; Hirsch-Kuchma, M.; Freiberg, E.; Davison, A.; Jones, A.G. The reaction chemistry of a technetium(I) nitrosyl complex with potentially chelating organohydrazines: The X-ray crystal structure of [TcCl2(NO)(HNNC5H4N)(PPh3)]. Inorg. Chim. Acta 1998, 279, 206–209. [Google Scholar]
  52. De Vries, N.; Cook, J.; Davison, A.; Nicholson, T.; Jones, A.G. Synthesis and characterization of a technetium(III) nitrosyl compound: Tc(NO)(Cl)(SC10H13)3. Inorg. Chem. 1990, 29, 1062–1064. [Google Scholar] [CrossRef]
  53. Nicholson, T.; Mahmood, A.; Limpa-Amara, N.; Salvarese, N.; Takase, M.K.; Müller, P.; Akgun, Z.; Jones, A.G. Reactions of the tridentate and tetradentate amine ligands di-(2-picolyl)(N-ethyl)amine and 2,5-bis-(2-pyridylmethyl)-2,5 diazohexane with technetium nitrosyl complexes. Inorg. Chim. Acta 2011, 373, 301–305. [Google Scholar] [CrossRef]
  54. Roca Jungfer, M.; Ernst, M.J.; Hagenbach, A.; Abram, U. [{TcI(NO)(LOMe)(PPh3)Cl}2Ag](PF6) and [TcII(NO)(LOMe)(PPh3)Cl](PF6): Two Unusual Technetium Complexes with a “Kläui-type” Ligand. Z. Anorg. Allg. Chem. 2022, 648, e202100316. [Google Scholar]
  55. Nicholson, T.L.; Mahmood, A.; Muller, P.; Davison, A.; Storm-Blanchard, S.; Jones, A.G. The synthesis and structural characterization of the technetium nitrosyl complexes [TcCl(NO)(SC5H4N)(PPh3)2] and [Tc(NO)(SC5H4N)2(PPh3)]. Inorg. Chim. Acta 2011, 365, 484–486. [Google Scholar] [CrossRef] [PubMed]
  56. Balasekaran, S.M.; Spandl, J.; Hagenbach, A.; Köhler, K.; Drees, M.; Abram, U. Fluoridonitrosyl Complexes of Technetium(I) and Technetium(II). Synthesis, Characterization, Reactions, and DFT Calculations. Inorg. Chem. 2014, 53, 5117–5128. [Google Scholar] [CrossRef]
  57. Nicholson, T.; Hirsch-Kuchma, M.; Shellenbarger-Jones, A.; Davison, A.; Jones, A.G. The synthesis and characterization of a technetium nitrosyl complex with cis-{2-pyridyl,diphenylphosphine} coligands. The X-ray crystal structure of [TcCl2(NO)(pyPPh2-P,N) (pyPPh2-P)]. Inorg. Chim. Acta 1998, 267, 319–322. [Google Scholar] [CrossRef]
  58. Grunwald, A.C.; Scholtysik, C.; Hagenbach, A.; Abram, U. One Ligand, One Metal, Seven Oxidation States: Stable Technetium Complexes with the “Kläui Ligand”. Inorg. Chem. 2020, 59, 9396–9405. [Google Scholar] [CrossRef] [PubMed]
  59. Nicholson, T.L.; Mahmood, A.; Refosco, F.; Tisato, F.; Müller, P.; Jones, A.G. The synthesis and X-ray structural characterization of mer and fac isomers of the technetium(I) nitrosyl complex [TcCl2(NO)(PNPpr)]. Inorg. Chim. Acta 2009, 362, 3637–3640. [Google Scholar] [CrossRef] [PubMed]
  60. Nicholson, T.; Müller, P.; Davison, A.; Jones, A.G. The synthesis and characterization of a cationic technetium nitrosyl complex: The X-ray crystal structure of [TcCl(NO)(DPPE)2](PF6) × CH2Cl2. Inorg. Chim. Acta 2006, 359, 1296–1298. [Google Scholar] [CrossRef]
  61. Pandey, K.K. Coordination Chemistry of Thionitrosyl (NS), Thiazate (NSO), Disulfidothionitrate (S,N), Sulfur Monoxide (SO), and Disulfur Monoxide (S,O) Ligands. Progr. Inorg. Chem. 1992, 40, 445–502. [Google Scholar]
  62. Døsing, A. The electronic structure and photochemistry of transition metal thionitrosyl complexes. Coord. Chem. Rev. 2016, 306, 544–557. [Google Scholar] [CrossRef]
  63. Anhaus, J.; Siddiqi, Z.A.; Roesky, H.W. Reaction of Tetrasulfurtetranitride with Rhenium(VII)-chloronitride. The Crystal Structure of [Ph4As+]2[Cl4Re(NS)(NSCl)2−] × CH2Cl2. Z. Naturforsch. 1985, 40b, 740–744. [Google Scholar] [CrossRef]
  64. Dirican, D.; Pfister, N.; Wozniak, M.; Braun, T. Reactivity of Binary and Ternary Sulfur Halides towards Transition-Metal Compounds. Chem. Eur. J. 2020, 31, 6945–6963. [Google Scholar] [CrossRef]
  65. Dietrich, A.; Neumüller, B.; Dehnicke, K. (PPh4)2[(SN)ReCl3(μ-N)(μ-NSN)ReCl3(THF)]—Ein Nitrido-Thionitrosyl-Dinitridosulfato-Komplex des Rheniums. Z. Anorg. Allg. Chem. 2000, 626, 1268–1270. [Google Scholar] [CrossRef]
  66. Reinel, M.; Höcher, T.; Abram, U.; Kirmse, R. Ein Beitrag zu Rhenium(II)-, Osmium(II)- und Technetium(II)-Thionitrosylkomplexe vom Typ [M(NS)Cl4py]: Darstellung, Strukturen und EPR-Spektren. Z. Anorg. Allg. Chem. 2003, 629, 853–861. [Google Scholar] [CrossRef]
  67. Voigt, A.; Abram, U.; Kirmse, R. Darstellung, Strukturen und EPR-Spektren der Rhenium(II)-Thionitrosylkomplexe trans-[Re(NS)Cl3(MePh2P)2] und trans-[Re(NS)Br3(Me2PhP)2]. Z. Anorg. Allg. Chem. 1999, 625, 1658–1663. [Google Scholar] [CrossRef]
  68. Hauck, H.-G.; Willing, W.; Müller, U.; Dehnicke, K. [ReCl2(NS)(NSCl)(Pyridin)2], ein Thionitrosyl-chlorthionitrenkomplex des Rheniums. Z. Anorg. Allg. Chem. 1986, 534, 77–84. [Google Scholar] [CrossRef]
  69. Hübener, R.; Abram, U.; Strähle, J. Isothiocyanato complexes of rhenium II. Synthesis, characterization and structures of ReN(NCS)2(Me2PhP)3 and Re(NS)(NCS)2(Me2PhP)3. Inorg. Chim. Acta 1994, 216, 223–228. [Google Scholar] [CrossRef]
  70. Ritter, S.; Abram, U. Gemischtligand-Komplexe des Rheniums. VI. Darstellung und Strukturen der Rhenium Thionitrosyl-Komplexe mer-[Re(NS)Cl2(Me2PhP)3] × CH2Cl2 und trans-[Re(NS)Cl3(Me2PhP)2]. Z. Anorg. Allg. Chem. 1994, 620, 1223–1228. [Google Scholar]
  71. Ritter, S.; Abram, U. Gemischtligandkomplexe des Rheniums. IX. Reaktionen am Nitridoliganden von [ReN(Me2PhP)(Et2dtc)2]. Synthese, Charakterisierung und Kristallstrukturen von [Re(NBCl3)(Me2PhP)(Et2dtc)2], [Re(NGaCl3)(Me2PhP)(Et2dtc)2] und [Re(NS)Cl(Me2PhP)2(Et2dtc)]. Z. Anorg. Allg. Chem. 1995, 622, 965–973. [Google Scholar] [CrossRef]
  72. Ruf, C.; Behrens, U.; Lork, E.; Mews, R. Reactions of halides with trans-[Re(CO)4(MeCN)(NS)][AsF6]2: Syntheses and structure of trans-[Re(CO)4(Cl)(NS)][AsF6] and [(OC)5ReNS–NS–N[Re(CO)5]S{N–SNRe(CO)5}–CH2–CH2][AsF6]2, an unusual trinuclear bis(thiazyl)rhenium complex. Chem. Commun. 1996, 939–940. [Google Scholar] [CrossRef]
  73. Baldas, J.; Bonnyman, J.; Mackay, M.F.; Williams, G.A. Structural studies of technetium complexes. V. The preparation and crystal structure of Dichlorobis(diethyldithiocarbamato)thionitrosyltechnetium(III). Austr. J. Chem. 1984, 37, 751–759. [Google Scholar] [CrossRef]
  74. Kaden, L.; Lorenz, B.; Kirmse, R.; Stach, J.; Behm, H.; Beurskens, P.T.; Abram, U. Synthesis, characterization and x-ray molecular and crystal structure of Tc(NS)Cl3(Me2PhP)(Me2PhPO)-a first example of mixed phosphine/phosphine oxide coordination. Inorg. Chim. Acta 1990, 169, 43–48. [Google Scholar] [CrossRef]
  75. Baldas, J.; Colmanet, S.F.; Williams, G.A. Preparation and Structure of Dibromobis(N,N-diethyldithiocarbamato)-thionitrosyltechnetium(III). Austr. J. Chem. 1991, 44, 1125–1132. [Google Scholar] [CrossRef]
  76. Lu, J.; Clarke, M.J. Modulation of Tc–NX (X = O or S) bonds by π-acceptor ligands. J. Chem. Soc. Dalton Trans. 1992, 1243–1248. [Google Scholar] [CrossRef]
  77. Abram, U.; Schulz Lang, E.; Abram, S.; Wegmann, J.; Dilworth, J.R.; Kirmse, R.; Woolins, J.D. Technetium(V) and rhenium(V) nitrido complexes with bis(diphenyl-thiophosphoryl)amide, N(SPPh2)2. J. Chem. Soc. Dalton Trans. 1997, 623–630. [Google Scholar] [CrossRef]
  78. Hiller, W.; Hübener, R.; Lorenz, B.; Kaden, L.; Findeisen, M.; Stach, J.; Abram, U. Structural and spectroscopic studies on mer-dichlorotris(dimethylphenylphosphine)(thionitrosyl)technetium(I), mer-[Tc(NS)Cl2(Me2PhP)3]. Inorg. Chim. Acta 1991, 181, 161–165. [Google Scholar] [CrossRef]
  79. Lu, J.; Clarke, M.J. Sulfur atom transfer with reduction of a [TcVI≡N]3+ core to a [TcI-N≡S]2+ core. Crystal structure of mer-dichlorotris(4-picoline)(thionitrosyl)technetium. Inorg. Chem. 1990, 29, 4123–4125. [Google Scholar] [CrossRef]
  80. Abram, U.; Hübener, R.; Wollert, R.; Kirmse, R.; Hiller, W. Synthesis, characterization and reactions of [Tc(NS)X4] complexes (X = Cl, Br, NCS). Inorg. Chim. Acta 1993, 206, 9–14. [Google Scholar] [CrossRef]
  81. Dressler, K. Ultraviolett- und Schumannspektren der neutralen und ionisierten Moleküle PO, PS, NS, P2. Helv. Phys. Acta 1955, 28, 563–590. [Google Scholar]
  82. O’Hare, P.A.G. Dissociation Energies, Enthalpies of Formation, Ionization Potentials, and Dipole Moments of NS and NS+. J. Chem. Phys. 1970, 52, 2992–2996. [Google Scholar] [CrossRef]
  83. Mews, R. The Thionitrosyl Cation NS+ as a Synthetic Reagent. Angew. Chem. Int. Ed. Engl. 1976, 15, 691–692. [Google Scholar] [CrossRef]
  84. Clegg, W.; Glemser, O.; Harms, K.; Hartmann, G.; Mews, R.; Noltemeyer, M.; Sheldrick, G.M. Crystal structures of thionitrosyl hexafluoroantimonate(V) and thionitrosyl undecafluorodiantimonate(V) at 293 K and of thionitrosyl undecafluorodiantimonate(V) at 121.5 K: The effect of thermal motion on the apparent NS bond length. Acta Cryst. 1981, 37b, 548–552. [Google Scholar] [CrossRef]
  85. Kaden, L.; Lorenz, B.; Kirmse, R.; Stach, J.; Abram, U. Darstellung und Charakterisierung von Thionitrosylkomplexen des Technetiums(I) und –(II). Z. Chem. 1985, 25, 29–30. [Google Scholar] [CrossRef]
  86. Abram, U.; Kirmse, R.; Köhler, K.; Lorenz, B.; Kaden, L. Tc(NX)Y3(Me2PhP)2 Complexes (X = O or S; Y = Cl or Br). Preparation, Characterization and EPR Studies. Inorg. Chim. Acta 1987, 129, 15–20. [Google Scholar] [CrossRef]
  87. Nguyen, H.H.; Abram, U. Rhenium and Technetium Complexes with N,N-Dialkyl-N’-benzoylthioureas. Inorg. Chem. 2007, 46, 5310–5319. [Google Scholar] [PubMed]
  88. Hayes, T.R.; Powell, A.S.; Benny, P.D. Synthesis and stability of 2 + 1 complexes of N,N-diethylbenzoylthiourea with [MI(CO)3]+ (M = Re, 99mTc). J. Coord. Chem. 2015, 68, 3432–3448. [Google Scholar] [CrossRef]
  89. Abram, U.; Abram, S.; Alberto, R.; Schibli, R. Ligand exchange reactions starting from [Re(CO)3Br3]2−. Synthesis, characterization and structures of rhenium(I) tricarbonyl complexes with thiourea and thiourea derivatives. Inorg. Chim. Acta 1996, 248, 193–202. [Google Scholar] [CrossRef]
  90. Borges, A.P.; Possato, B.; Hagenbach, A.; Machado, A.E.H.; Deflon, V.M.; Abram, U.; Maia, P.I.S. Re(V) complexes containing the phenylimido (NPh2-) core and SNS-thiosemicarbazide ligands. Inorg. Chim. Acta 2021, 516, 120110. [Google Scholar] [CrossRef]
  91. Mukiza, J.; Gerber, T.I.A.; Hosten, E.C.; Betz, R. Crystal structure of facO,S-(Z)-1,1-diethyl-3-(hydroxido(phenyl)methylene) thiourea-(κS’(Z)-1,1-diethyl-3-(hydroxido(phenyl)methylene)-thiourea)-tricarbonyl rhenium(I), C27H31N4O5ReS2. Z. Kristallogr. New Cryst. Struct. 2015, 230, 50–52. [Google Scholar] [CrossRef]
  92. Nguyen, H.H.; Abram, U. Rhenium and technetium complexes with tridentate S,N,O ligands derived from benzoylhydrazine. Polyhedron 2009, 28, 3945–3952. [Google Scholar] [CrossRef]
  93. Salsi, F.; Portapilla, G.B.; Simon, S.; Roca Jungfer, M.; Hagenbach, A.; de Albuquerque, S.; Abram, U. Effect of Fluorination on the Structure and Anti-Trypanosoma cruzy Activity of Oxorhenium(V) Complexes with S,N,S-Tridentate Thiosemicarbazones and Benzoylthioureas. Synthesis and Structures of Technetium(V) Analogues. Inorg. Chem. 2019, 58, 10129–10138. [Google Scholar] [CrossRef]
  94. Roca Jungfer, M.; Elsholz, L.; Abram, U. Technetium Hydrides Revisited: Syntheses, Structures, and Reactions of [TcH3(PPh3)4] and [TcH(CO)3(PPh3)2]. Organometallics 2021, 40, 3095–3112. [Google Scholar] [CrossRef]
  95. Kläui, W. The Coordination Chemistry and Organometallic Chemistry of Tridentate Oxygen Ligands with π-Donor Properties. Angew. Chem. Int. Ed. Engl. 1990, 29, 627–637. [Google Scholar] [CrossRef]
  96. Leung, W.-H.; Zhang, Q.-F.; Yi, X.-Y. Recent developments in the coordination and organometallic chemistry of Kläui oxygen tripodal ligands. Coord. Chem. Rev. 2007, 251, 2266–2279. [Google Scholar] [CrossRef]
  97. Kramer, D.J.; Davison, A.; Jones, A.G. Structural models for [M(CO)3(H2O)3]+ (M = Tc, Re): Fully aqueous synthesis of technetium and rhenium tricarbonyl complexes of tripodal oxygen donor ligands. Inorg. Chim. Acta 2001, 312, 215–220. [Google Scholar] [CrossRef]
  98. Leung, W.-H.; Chan, E.Y.Y.; Lai, T.C.Y.; Wong, W.-T. Synthesis and reactivity of nitrido-rhenium and -osmium complexes with an oxygen tripod ligand. J. Chem. Soc. Dalton Trans. 2000, 51–56. [Google Scholar] [CrossRef]
  99. So, Y.-M.; Chiu, W.-H.; Cheung, W.-M.; Ng, H.-Y.; Lee, H.K.; Sung, H.H.-Y.; Williams, I.D.; Leung, W.H. Heterobimetallic rhenium nitrido complexes containing the Kläui tripodal ligand [Co(η5-C5H5){P(O)(OEt)2}3]. Dalton Trans. 2015, 44, 5479–5487. [Google Scholar] [CrossRef]
  100. Banberry, H.J.; Hussain, W.; Evans, I.G.; Hamor, T.A.; Jones, C.J.; McCleverty, J.A.; Schulte, H.-J.; Engles, B.; Kläui, W. The syntheses of high oxidation state metal complexes containing the tripodal ligand [(η5-C5H5)Co{P(OMe)2(O)}3] and the X-ray crystal structure of [(η5-C5H5)Co{P(OMe)2(O)}3 ReO3]. Polyhedron 1990, 9, 2549–2551. [Google Scholar] [CrossRef]
  101. Dyckhoff, B.; Schulte, H.-J.; Englert, U.; Spaniol, T.P.; Kläui, W.; Schubiger, P.A. Rhenium-Komplexe von Sauerstoffchelatliganden: Ein Weg zu neuen Radiopharmaka? Z. Anorg. Allg. Chem. 1992, 614, 131–141. [Google Scholar] [CrossRef]
  102. Abram, U.; Abram, S. Synthese und Charakterisierung neuartiger Technetiumkomplexe mit 1,1-disubstituierten Benzoylthioharnstoffen. Z. Chem. 1983, 23, 228. [Google Scholar] [CrossRef]
  103. Sullivan, B.P.; Brewer, J.C.; Gray, H.B.; Linebarrier, D.; Mayer, J.M. Nitrido and Oxo Complexes of Rhenium(V). Inorg. Synth. 1992, 29, 146–150. [Google Scholar]
  104. Kaden, L.; Lorenz, B.; Schmidt, K.; Sprinz, H.; Wahren, M. Nitridokomplexe des Technetium(V). Isotopenpraxis 1981, 17, 174–175. [Google Scholar]
  105. Baldas, J.; Boas, J.F.; Bonnyman, J.; Williams, G.A. Studies of technetium complexes. Part 6. The preparation, characterisation, and electron spin resonance spectra of salts of tetrachloro- and tetrabromonitridotechnetate(VI): Crystal structure of tetraphenylarsonium tetrachloronitridotechnetate(VI). J. Chem. Soc. Dalton Trans. 1984, 11, 2395–2400. [Google Scholar] [CrossRef]
  106. Abram, U.; Braun, M.; Abram, S.; Kirmse, R.; Voigt, A. [NBu4][ReNCl4]: Facile synthesis, structure, electron paramagnetic resonance spectroscopy and reactions. J. Chem. Soc. Dalton Trans. 1998, 2, 231–238. [Google Scholar] [CrossRef]
  107. Grunwald, A.C. Metal Complexes with Tripodal and Chelating Thiourea Ligands towards Nuclear Medical Imaging. Doctoral Thesis, Freie Universität Berlin, Berlin, Germany, 2020. Available online: https://refubium.fu-berlin.de/handle/fub188/30387 (accessed on 1 July 2024).
  108. Nowak, D. Thionitrosylkomplexe des Rheniums und Technetiums. Doctoral Thesis, Freie Universität Berlin, Berlin, Germany, 2022. Available online: https://refubium.fu-berlin.de/handle/fub188/36899 (accessed on 1 July 2024).
  109. Kleinpeter, E.; Beyer, L. 1H-NMR-Untersuchung der behinderten Rotation um die C-N-Bindung in 1,1’-Diäthyl-3-benzoylharnstoff-Derivaten. J. Prakt. Chem. 1975, 317, 938–942. [Google Scholar] [CrossRef]
  110. Stoll, S.; Schweiger, A. EasySpin, a comprehensive software package for spectral simulation and analysis in EPR. J. Magn. Res. 2006, 178, 42–55. [Google Scholar] [CrossRef] [PubMed]
  111. MATLAB Version: 9.13.0 (R2022b); The MathWorks Inc.: Natick, MA, USA, 2022.
  112. Sheldrick, G. SADABS, vers. 2014/5; University of Göttingen: Göttingen, Germany, 2014. [Google Scholar]
  113. Coppens, P. The Evaluation of Absorption and Extinction in Single-Crystal Structure Analysis. In Crystallographic Computing; Muksgaard: Copenhagen, Denmark, 1979. [Google Scholar]
  114. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. 2008, A64, 112–122. [Google Scholar] [CrossRef] [PubMed]
  115. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. 2015, C71, 3–8. [Google Scholar]
  116. Farrugia, L.J. WinGX and ORTEP for Windows: An update. J. Appl. Cryst. 2012, 45, 849–854. [Google Scholar] [CrossRef]
  117. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  118. Putz, H.; Brandenburg, K. Diamond—Crystal and Molecular Structure Visualization; Vers. 4.6.8.; Crystal Impact: Bonn, Germany, 2022. [Google Scholar]
Figure 1. Common starting materials for the synthesis of thionitrosyl complexes.
Figure 1. Common starting materials for the synthesis of thionitrosyl complexes.
Inorganics 12 00210 g001
Scheme 1. Syntheses of [Re(NS)Cl3(PPh3)2] (1) and [Re(NS)Cl3(PPh3)(OPPh3)] (2).
Scheme 1. Syntheses of [Re(NS)Cl3(PPh3)2] (1) and [Re(NS)Cl3(PPh3)(OPPh3)] (2).
Inorganics 12 00210 sch001
Figure 2. Molecular structures of (a) [Re(NS)Cl3(PPh3)2] (1) and (b) [Re(NS)Cl3(PPh3)(OPPh3)] (2). Thermal ellipsoids represent 50 percent probability.
Figure 2. Molecular structures of (a) [Re(NS)Cl3(PPh3)2] (1) and (b) [Re(NS)Cl3(PPh3)(OPPh3)] (2). Thermal ellipsoids represent 50 percent probability.
Inorganics 12 00210 g002
Figure 3. X-band EPR spectra of 1 and 2: (a) at room temperature (parameters for 1 in CH2Cl2: g0 = 2.025; a0Re = 342 × 10−4 cm−1; parameters for 2 in acetone: g0 = 2.017; a0Re = 372 × 10−4 cm−1) and (b) in frozen solution at T = 77 K (parameters for 1 in CH2Cl2: gx = 2.110, gy = 1.991, gz = 1.943; AxRe = 281 × 10–4 cm–1, AyRe = 283 × 10−4 cm−1, AzRe = 524 × 10−4 cm−1; parameters for 2 in acetone: g = 1.923, g = 2.008; ARe = 594 × 10−4 cm−1, ARe = 293 × 10−4 cm−1).
Figure 3. X-band EPR spectra of 1 and 2: (a) at room temperature (parameters for 1 in CH2Cl2: g0 = 2.025; a0Re = 342 × 10−4 cm−1; parameters for 2 in acetone: g0 = 2.017; a0Re = 372 × 10−4 cm−1) and (b) in frozen solution at T = 77 K (parameters for 1 in CH2Cl2: gx = 2.110, gy = 1.991, gz = 1.943; AxRe = 281 × 10–4 cm–1, AyRe = 283 × 10−4 cm−1, AzRe = 524 × 10−4 cm−1; parameters for 2 in acetone: g = 1.923, g = 2.008; ARe = 594 × 10−4 cm−1, ARe = 293 × 10−4 cm−1).
Inorganics 12 00210 g003
Figure 4. N,N-Diethyl-N′-benzoylthiourea (HEt2btu) and (η5-cyclopentadienyl)tris(dimethyl phosphito-P)cobaltate(III) ({LOMe}) as typical chelating ligands for ligand exchange reactions.
Figure 4. N,N-Diethyl-N′-benzoylthiourea (HEt2btu) and (η5-cyclopentadienyl)tris(dimethyl phosphito-P)cobaltate(III) ({LOMe}) as typical chelating ligands for ligand exchange reactions.
Inorganics 12 00210 g004
Scheme 2. Ligand exchange reactions starting from [Re(NS)Cl3(PPh3)2] (1).
Scheme 2. Ligand exchange reactions starting from [Re(NS)Cl3(PPh3)2] (1).
Inorganics 12 00210 sch002
Figure 5. Molecular structure of [Re(NS)(PPh3)(Et2btu)2] (3a). Thermal ellipsoids represent 50 percent probability.
Figure 5. Molecular structure of [Re(NS)(PPh3)(Et2btu)2] (3a). Thermal ellipsoids represent 50 percent probability.
Inorganics 12 00210 g005
Scheme 3. Formation of thionitrosyl complexes containing the “Kläui ligand” {LOMe}.
Scheme 3. Formation of thionitrosyl complexes containing the “Kläui ligand” {LOMe}.
Inorganics 12 00210 sch003
Figure 6. Molecular structures of (a) the complex cation of [Re(NS)Cl(PPh3)(LOMe)](PF6) (4a) and (b) [Tc(NS)Cl2(LOMe)] (5b). Thermal ellipsoids represent 50 percent probability.
Figure 6. Molecular structures of (a) the complex cation of [Re(NS)Cl(PPh3)(LOMe)](PF6) (4a) and (b) [Tc(NS)Cl2(LOMe)] (5b). Thermal ellipsoids represent 50 percent probability.
Inorganics 12 00210 g006
Figure 7. Liquid solution EPR spectra of [Tc(NS)Cl(PPh3)(LOMe)] in CH2Cl2 (a) at room temperature and (b) at T = 77 K (assignment of the 99Tc lines to the parallel and perpendicular parts of the spectrum is indicated).
Figure 7. Liquid solution EPR spectra of [Tc(NS)Cl(PPh3)(LOMe)] in CH2Cl2 (a) at room temperature and (b) at T = 77 K (assignment of the 99Tc lines to the parallel and perpendicular parts of the spectrum is indicated).
Inorganics 12 00210 g007
Table 1. Selected bond lengths (Å) and angles (°) in [Re(NS)Cl3(PPh3)2] (1) and (b) [Re(NS)Cl3(PPh3)(OPPh3)] (2).
Table 1. Selected bond lengths (Å) and angles (°) in [Re(NS)Cl3(PPh3)2] (1) and (b) [Re(NS)Cl3(PPh3)(OPPh3)] (2).
Re1–N10N10–S10Re1–Cl1Re1–Cl2Re1–Cl3Re1–P1Re1–P2/O1O1–P2Re1–N10–S10Re1–O1–P2
11.795(5)1.518(5)2.353(1)2.340(1)2.416(1)2.540(2)2.555(2)-174.4(3)-
2 (1)1.77(1)
1.77(1)
1.53(1)
1.52(1)
2.368(4)
2.357(4)
2.255(3)
2.332(3)
2.338(3)
2.366(3)
2.506(4)
2.584(4)
2.100(8)
2.099(8)
1.509(9)
1.497(9)
167.2(7)
178.4(6)
155.7(6)
164.5(5)
(1) Values for two crystallographically independent molecules.
Table 2. Selected bond lengths (Å) and angles (°) [Re(NS)(PPh3)(Et2btu)2] (3a).
Table 2. Selected bond lengths (Å) and angles (°) [Re(NS)(PPh3)(Et2btu)2] (3a).
Re1–N10N10–S10Re1–P1Re1–S1Re1–O1Re1–S11Re1–O11O1–C1C1–N1N1–C2
1.749(4)1.571(4)2.362(1)2.401(1)2.101(3)2.439(1)2.103(4)1.296(6)1.301(7)1.353(7)
C2–N2C2–S1O11–C11C11–N11N11–C12C12–N12C12–S11Re1–N10–S10
1.319(7)1.755(6)1.266(6)1.329(6)1.338(7)1.325(7)1.758(6)176.5(3)
Table 3. Selected bond lengths (Å) and angles (°) in [Re(NS)Cl(PPh3)(LOMe)](PF6) (4a(PF6)) and (b) [Tc(NS)Cl2(LOMe)] (5b).
Table 3. Selected bond lengths (Å) and angles (°) in [Re(NS)Cl(PPh3)(LOMe)](PF6) (4a(PF6)) and (b) [Tc(NS)Cl2(LOMe)] (5b).
M1–N10N10–S10M1–Cl1M1–Cl2M1–P1M1–O2M1–O3M1–O4M1–N10–S10
4a1.752(5)1.554(5)2.337(1)-2.475(1)2.071(3)2.032(3)2.083(3)175.4(3)
5b (1)1.732(7)
1.75(1)
1.540(7)
1.54(1)
2.339(3)
2.330(3)
2.343(2)
2.344(3)
-2.060(6)
2.068(5)
2.047(6)
2.057(6)
2.095(5)
2.088(5)
171.7(6)
172.0(8)
(1) Values for two crystallographically independent molecules.
Table 4. EPR parameters of the Re(II) and Tc(II) complexes with {LOMe} ligands, coupling constants are given in 10–4 cm–1.
Table 4. EPR parameters of the Re(II) and Tc(II) complexes with {LOMe} ligands, coupling constants are given in 10–4 cm–1.
g0a0MggAMAM
[Re(NS)Cl(PPh3)(LOMe)]Cl (4aCl)1.9954201.8421.986644330
[Tc(NS)Cl(PPh3)(LOMe)]Cl (4bCl)2.0071681.9662.023252112
[Re(NS)Cl2(LOMe)] (5a)1.9904551.9451.760787398
[Tc(NS)Cl2(LOMe)] (5b)1.9961771.9432.021285129
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

Nowak, D.; Hagenbach, A.; Sawallisch, T.E.; Abram, U. Thionitrosyl Complexes of Rhenium and Technetium with PPh3 and Chelating Ligands—Synthesis and Reactivity. Inorganics 2024, 12, 210. https://doi.org/10.3390/inorganics12080210

AMA Style

Nowak D, Hagenbach A, Sawallisch TE, Abram U. Thionitrosyl Complexes of Rhenium and Technetium with PPh3 and Chelating Ligands—Synthesis and Reactivity. Inorganics. 2024; 12(8):210. https://doi.org/10.3390/inorganics12080210

Chicago/Turabian Style

Nowak, Domenik, Adelheid Hagenbach, Till Erik Sawallisch, and Ulrich Abram. 2024. "Thionitrosyl Complexes of Rhenium and Technetium with PPh3 and Chelating Ligands—Synthesis and Reactivity" Inorganics 12, no. 8: 210. https://doi.org/10.3390/inorganics12080210

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

Article Metrics

Back to TopTop