Next Article in Journal
(1R,2R,6S)-2(4-(4-Isopropylbenzyl)piperazin-1-yl)-3-methyl-6-(prop-1-en-2-yl)cyclohex-3-enol
Previous Article in Journal
(S)-N1,N3-Dibenzyl-1-cyclohexyl-N1,N3-bis((R)-1-phenylethyl)propane-1,3-diamine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Short Note

24-N-(2-Butynyl)phthalimide)(hexacarbonyl)dicobalt

1
Institut UTINAM UMR 6213 CNRS, Université de Franche-Comté, 16 Route de Gray, 25030 Besançon, France
2
Anorganische Chemie, Technische Universität Dortmund, Otto-Hahn Straße 6, 44227 Dortmund, Germany
3
URCOM, UNILEHAVRE, Normandie University, FR3021, EA 3221, 25 Rue P. Lebon, 76058 Le Havre, France
*
Authors to whom correspondence should be addressed.
Molbank 2023, 2023(1), M1545; https://doi.org/10.3390/M1545
Submission received: 8 December 2022 / Revised: 26 December 2022 / Accepted: 6 January 2023 / Published: 11 January 2023
(This article belongs to the Section Organic Synthesis)

Abstract

:
The reaction of [Co2(CO)8] with an equimolar amount of the internal alkyne N-(2-butynyl)phthalimide (1-Phthalimido-2-butyne) 1 in heptane solution yields the title compound [Co2(CO)6(µ-phthalimidoCH2C≡CMe)] 2. Compound 2 has been characterized using IR, 1H and 13C NMR spectroscopy; the tetrahedrane-type cluster framework has been ascertained using a single-crystal X-ray diffraction study performed at 100 K.

Graphical Abstract

1. Introduction

Since the first reports dating from the 1950s on the reaction of dicobalt octacarbonyl with alkynes yielding dicobaltatetrahedranes [Co2(CO)6(µ-RC≡CR′)] [1,2], there is nowadays a plethora of articles dealing with this organometallic reaction [3]. The interest in this research is not only driven by merely synthetic aspects, i.e., spectroscopic and structural characterization of these cage-like species, but also by the use of these compounds as precursors for further modifications in material sciences [4] and organic syntheses [5,6]. An important example for the latter domain is the Pauson–Khand reaction, in which [Co2(CO)8] reacts with an alkyne, an alkene, and carbon monoxide to form an α,β-cyclolopentenone [7,8,9,10,11]. In addition to the applications using dicobaltatetrahedranes as synthetic tools, this class of dinuclear organometallic compounds is emerging in bio-organometallic chemistry [12,13,14,15,16,17,18,19,20,21]. Some examples of dicobaltatetrahedranes, ligated using terminal functionalized alkynes and displaying biological activity, are depicted in Scheme 1.
Among the organic heterocyclic compounds showing a promising biological and medicinal activity, there are also some representatives based on the phthalimido scaffold [22,23,24]. In 2005, Gust et al. also reported the reaction of propargyl phtalimide with [Co2(CO)8] and determined the cytotoxicity and DNA binding efficiency of the resulting complex N-(2-propynyl)phthalimide]hexacarbonyldicobalt (see compound V of Scheme 1) [21]. In the context of our research on Co–Co carbonyl complexes towards various alkynes producing dicobaltatetrahedranes [25,26], we attempted to synthetize a Co–Co complex ligated using an internal alkyne, namely 1-Phthalimido-2-butyne 1, for upcoming biological studies. Apart from the purely synthetic aspect, one of the objectives was to compare the impact of the replacement of a terminal alkyne by an internal one (Phtal-C≡CH vs. Phtal-C≡CMe).

2. Results and Discussion

The title compound 2 was obtained via treatment of [Co2(CO)8] with an equimolar amount of N-(2-butynyl)phthalimide 1 in heptane solution at 60 °C as shown in Scheme 2. Upon cooling, small orange-red needle-shaped crystals of air-stable 2 were isolated, and elemental analysis confirmed their composition as [Co2(CO)6(µ-phthalimidoCH2C≡CMe)].
The IR spectrum of this product in cyclohexane, shown in Figure 1, reveals the signals of terminal carbonyl with intense ν(CO) vibrations at 2093, 2055, and 2028 cm−1. In addition, two further absorptions at 1775 and 1727 cm−1 are attributed to the carbonyls’ stretching modes (symmetric and antisymmetric) of the imide function [27]. These values fit well with those measured in cyclohexane for [Co2(CO)6(µ-phthalimidoCH2C≡CH)] at 2094, 2057, and 2030 as well as 1775 and 1727 cm−1 (Figure S1). The latter complex has already been described and characterized using IR spectroscopy in solid state as KBr pellets (2095, 2054, 2036, 2016 (Co-CO); 1712 (C=O) cm−1) [21]. The infrared spectrum of 2 in the solid state was also recorded in attenuated total reflectance (ATR) mode and is presented in the Supplementary Materials (Figure S2). Despite being a monocrystalline solid, the carbonyl absorptions were not fully resolved (2093, 2052, 2030, and 2008 as well as 1766 and 1708 cm−1).
The 1H-NMR recorded in CDCl3 reveals a singlet at δ 2.65 due to the acetylenic methyl group; a further singlet at δ 5.03 is attributed to the enantiotopic methylene hydrogens. The remaining resonances at δ7.74 and 7.88 are assigned to the aromatic cycle (Figure 2). In the proton-decoupled 13C-NMR spectrum depicted in Figure 3, only a broadened resonance centered at δ 199.3 could be observed for the six Co-bound carbonyl groups, suggesting a fluxional behavior of the COs. In contrast, the resonance of the two chemically equivalent imido C=O carbonyl groups at δ 167.9, the two acetylenic carbons at δ 93. 2 and 91.8, as well as those of the CH2 and CH3 carbons gave rise to well-resolved signals.
In addition to the spectroscopic characterization of 2 in solution, the complex was also unambiguously analyzed using X-ray diffraction performed at 100 K. Complex 2 crystallizes in the triclinic space group P-1; the asymmetric unit contains two independent molecules with slightly different bond lengths and angles. As shown in Figure 4, in this organometallic species the former alkyne is almost orthogonally µ24-coordinated with respect to the two crystallographically non-equivalent Co1 and Co2 centers, which are linked through a metal–metal bond of 2.4688(14) Å.
The resulting tetrahedral skeleton, reminiscent of that of organic tetrahedranes (https://en.wikipedia.org/wiki/Tetrahedrane, accessed on 26 December 2022), allows for considering this 40-electron species an organometallic cluster of dimetallatetrahedranes, in which two Co(CO)3 units of the parent cluster [{Co42-CO)3(CO)9] are replaced by isolobal C–R fragments, forming the edges of the cluster core. The former alkyne unit lost its linearity; the torsion angle C1–C2–C3–C4 changed to −3.1(18)°. The coordination sphere around each Co atom is completed by three terminal carbonyls. The mean Co–C bond length of the four pseudo-equatorial carbonyls is somewhat longer than that of the two pseudo-axial C≡O ligands (1.815(7) vs. 1.789(7) Å). The C2–C3 bond distance of 2 is much shorter than that of a C–C single bond, but considerably elongated with respect to the C≡C bond reported for [phthalimidoCH2C≡CCH2phthalimido] (CSD refcode ECUFOB) (1.323(9) vs. 1.191(4) Å) [28].
Although there are numerous examples of crystallographically characterized hexacarbonyldicobalt compounds which are µ222 capped by an internal alkyne, a survey of the CSD data base revealed only two other examples containing a -CH2-C≡C-Me motif as is present in 2, namely (trans-5,8,8-trimethyl-trans-1,4,5,7-tetrahydro-cis-4-(2-butynyl)bicyclo(5.1.0)octan-3-one)(hexacarbonyl)dicobalt (A) and (pent-3-yn-1-ol)(hexacarbonyl)dicobalt (B) (Figure 5) [29,30]. The most relevant averaged bond lengths (comprising molecules 1 and 2) of 2 are listed in Table 1 and compared with those of compounds A and B (averaged bond lengths).
In the molecular cell of 2 with Z = 4, there are no intermolecular contacts deserving any discussion (Figure 6).

3. Materials and Methods

N-(2-butynyl)phthalimide 1 was commercially purchased from Aldrich. 1H- and 13C-NMR spectra were recorded on a Brucker AC 400 (Bruker, Wissembourg, France) at 400 and 100.62 MHz, respectively, using CDCl3 as solvent. Infrared spectra were recorded on a Vertex 70 spectrometer (Bruker, Wissembourg, France) in ATR mode or in solution.
24-N-(2-butynyl)phthalimide)(hexacarbonyl)dicobalt: Alkyne 1 (199.2 mg, 1 mmol) was added to a stirred solution of Co2(CO)8 (342.0 mg, 1 mmol) in heptane (10 mL). An immediate gas evolution was observed. The reaction mixture was then heated to 60 °C for 5 h. The solution was cooled to room temperature prior to lowering its temperature to 4 °C. Product 2 crystallized as orange-red needles which were collected via filtration. Yield: 78%. Anal. Calc. for C18H9Co2NO8 (M.W = 485.14 g.mol−1): C, 44.56; H, 1.87; N, 2.89%. Detected: C, 44.64; H, 1.92; N, 2.98 %. 1H-NMR (CDCl3) at 298 K: δ 2.65 (s, CH3), 5.03 (s, NCH2), 7.74 (br, CH Ar), 7.88 (br, CH Ar) ppm. 13C{1H}-NMR (CDCl3) at 298 K: δ 20.3 (CH3), 39.7 (NCH2), 91.8 (Cq), 93.2 (Cq), 123.6 (CH), 132.0 (Cq), 134.4 (CH), 167.9 (C=O), 199.3 (br, CO) ppm.
Since the grown single-crystals of 2 used for the determination of the crystal structure were quite small, CuKα radiation was employed instead of MoKα radiation. Moreover, the crystals were twinned. The B-alert found in the checkcif is due to the high electronic residual around the Co–Co bonds, which could not be further refined due to the low resolution at 138.0°.
Crystal data for C18H9Co2NO8, M = 485.12 g.mol−1, orange-red needles, crystal size 0.377 × 0.194 × 0.15 mm3, triclinic, space group P-1: a = 7.4317(4)Å, b = 13.4351(7) Å, c = 19.2794(11) Å, α = 78.462(3°, β = 84.011(3)°, γ = 89.156(3)°, V = 1875.73(18) Å3, Z = 4, Dcalc = 1.718 g/cm3, T = 100 K, h = −8 ≤ h ≤ 8, k = −15 ≤ k ≤ 16, I = 0 ≤ l ≤ 16, GOF = 1.077, R1 = 0.0855, wR2 = 0.2428 for 6962 reflections with I > = 2σ (I) and 11958 independent reflections. Largest diff. peak/hole/e Å−3 1.35/−1.02. The structure was solved using direct methods and refined using full-matrix least-squares against F2 (SHELXL, 2015 [31,32,33]).
Data were collected using graphite-monochromated CuKα radiation l = 1.54178 Å and were deposited at the Cambridge Crystallographic Data Centre as CCDC 2219351. (Supplementary Materials). The data can be obtained free of charge from the Cambridge Crystallographic Data Centre via http://www.ccdc.cam.ac.uk/getstructures.

4. Conclusions

We have demonstrated that addition of the internal alkyne N-(2-butynyl)phthalimide 1 to [Co2(CO)8] straightforwardly yields the stable complex [Co2(CO)6(µ-phthalimidoCH2C≡CMe)] 2 featuring a dimetallatetrahedrane framework. The investigation of the biological activity of 2 and other related compounds will be the topic of a forthcoming study.

Supplementary Materials

The following are available online, CIF file, Check-CIF report, and ATR-IR spectrum.

Author Contributions

I.J. and T.C. prepared the compound; C.S. and J.-L.K. collected the X-ray data and solved the structure; I.J., M.K. and M.O. designed the study; I.J. and M.K. analyzed the data and wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Région Bourgogne-Franche-Comté, DeCOmAB project.

Data Availability Statement

The X-ray data are at CCDC as stated in the paper.

Acknowledgments

We thank Stéphanie Beffy for recording some IR and NMR spectra. J.-L.K. acknowledges the receipt of a Kekulé scholarship from the Fonds der Chemischen Industrie (VCI).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Greenfield, H.; Sternberg, H.W.; Friedel, R.A.; Wotiz, J.H.; Markby, R.; Wender, I. Acetylenic dicobalt hexacarbonyls. organometallic compounds derived from alkynes and dicobalt octacarbonyl. J. Am. Chem. Soc. 1956, 78, 120–124. [Google Scholar] [CrossRef]
  2. Sly, W.G. The Molecular configuration of dicobalt hexacarbonyl diphenylacetylene. J. Am. Chem. Soc. 1959, 81, 18–20. [Google Scholar] [CrossRef]
  3. Barnes, C.E. Comprehensive Organometallic Chemistry II; Cobalt, Rhodium and Iridium; Elsevier: New York, NY, USA, 1995; Volume 8. [Google Scholar]
  4. Georgi, C.; Hildebrandt, A.; Waechtler, T.; Schulz, S.E.; Gessner, T.; Lang, H. A cobalt layer deposition study: Dicobaltatetrahedranes as convenient MOCVD precursor systems. Mat. J. Chem. 2014, 2, 4676–4682. [Google Scholar] [CrossRef] [Green Version]
  5. Lockwood, R.F.; Nicholas, K.M. Transition metal-stabilized carbenium ions as synthetic intermediates. I. α-[(alkynyl)dicobalt hexacarbonyl] carbenium ions as propargylating agents. Tetrahedron Lett. 1977, 18, 4163–4165. [Google Scholar] [CrossRef]
  6. Nicholas, K.M. Chemistry and synthetic utility of cobalt-complexed propargyl cations. Acc. Chem. Res. 1987, 20, 207–214. [Google Scholar] [CrossRef]
  7. Pauson, P.L.; Khand, I.U. Uses of cobalt-carbonyl acetylene complexes in organic synthesis. Ann. N.Y. Acad. Sci. 1977, 295, 2–14. [Google Scholar] [CrossRef]
  8. Hay, A.M.; Kerr, W.J.; Kirk, G.G.; Middlemiss, D. Highly efficient enantioselective Pauson-Khand reactions. Organometallics 1995, 14, 4986–4988. [Google Scholar] [CrossRef]
  9. Gibson, S.E.; Stevenazzi, A. The Pauson-Khand reaction: The catalytic age is here! Angew. Chem. Int. Ed. 2003, 4, 1800–1810. [Google Scholar] [CrossRef]
  10. Blanco-Urgoiti, J.; Añorbe, L.; Pérez-Serrano, L.; Domínguez, G.; Pérez-Castells, J. The Pauson–Khand reaction, a powerful synthetic tool for the synthesis of complex molecules. Chem. Soc. Rev. 2004, 33, 32–42. [Google Scholar] [CrossRef]
  11. Werner, H. Peter Ludwig Pauson (1925–2013). Angew. Chem. Int. Ed. 2014, 53, 3309. [Google Scholar] [CrossRef]
  12. Ott, I.; Kircher, B.; Dembinski, R.; Gust, R. Alkyne hexacarbonyl dicobalt complexes in medicinal chemistry and drug development. Expert Opin. Ther. Patents. 2008, 18, 327–337. [Google Scholar] [CrossRef]
  13. Minteanu, C.R.; Suntharalingam, K. Advances in cobalt complexes as anticancer agents. Dalton Trans. 2015, 44, 13796–13808. [Google Scholar] [CrossRef]
  14. Jung, M.; Kerr, D.E.; Senter, P.D. Bioorganometallic chemistry: Synthesis and antitumor activity of cobalt carbonyl complexes. Arch. Pharm. 1997, 330, 173–176. [Google Scholar] [CrossRef] [PubMed]
  15. Schmidt, K.; Jung, M.; Keilitz, R.; Schnurr, B.; Gust, R. Acetylenehexacarbonyldicobalt complexes, a novel class of antitumor drugs. Inorg. Chim. Acta. 2000, 306, 6–16. [Google Scholar] [CrossRef]
  16. Ott, I.; Kircher, B.; Gust, R. Investigations on the effects of cobalt-alkyne complexes on leukemia and lymphoma cells:  cytotoxicity and cellular uptake. J. Inorg. Biochem. 2004, 98, 485–489. [Google Scholar] [CrossRef] [PubMed]
  17. Vessieres, A.; Top, S.; Vaillant, C.; Osella, D.; Mornon, J.P.; Jaouen, G. Estradiols modified by metal carbonyl clusters as suicide substrates for the study of receptor proteins: Application to the estradiol receptor. Angew. Chem. Int. Ed. Engl. 1992, 31, 753–755. [Google Scholar] [CrossRef]
  18. Sergeant, C.D.; Ott, I.; Sniady, A.; Meneni, S.; Gust, R.; Rheingold, A.L.; Dembinski, R. Metallo-nucleosides: Synthesis and biological evaluation of hexacarbonyl dicobalt 5-alkynyl-2′-deoxyuridines. Org. Biomol. Chem. 2008, 6, 73–80. [Google Scholar] [CrossRef]
  19. Kaczmarek, R.; Korczyński, D.; Królewska-Golińska, K.; Wheeler, K.A.; Chavez, F.A.; Mikus, A.; Dembinski, R. Organometallic Nucleosides: Synthesis and Biological Evaluation of Substituted Dicobalt Hexacarbonyl 2′-Deoxy-5-oxopropynyluridines. ChemistryOpen 2018, 7, 237–247. [Google Scholar] [CrossRef] [Green Version]
  20. Neukamm, M.A.; Pinto, A.; Metzler-Nolte, N. Synthesis and cytotoxicity of a cobaltcarbonyl-alkyne enkephalin bioconjugate. Chem. Commun. 2008, 2, 232–234. [Google Scholar] [CrossRef]
  21. Ott, I.; Schmidt, K.; Kircher, B.; Schumacher, P.; Wiglenda, T.; Gust, R. Antitumor-active cobalt-alkyne complexes derived from acetylsalicylic acid: Studies on the mode of drug action. J. Med. Chem. 2005, 48, 622–629. [Google Scholar] [CrossRef]
  22. Almeida, M.L.; Oliveira, M.C.V.A.; Pitta, I.R.; Pitta, M.G.R. Advances in Synthesis and Medicinal Applications of Compounds Derived from Phthalimide. COS 2020, 17, 252–270. [Google Scholar] [CrossRef] [PubMed]
  23. Sharma, U.; Kumar, P.; Kumar, N.; Singh, B. Recent Advances in the Chemistry of Phthalimide Analogues and their Therapeutic Potential. Min. Rev. Med. Chem. 2010, 10, 678–704. [Google Scholar] [CrossRef] [PubMed]
  24. Pigeon, P.; Gaschard, M.; Othman, M.; Salmain, M.; Jaouen, G. α-Hydroxylactams as Efficient Entries to Diversely Functionalized Ferrociphenols: Synthesis and Antiproliferative Activity Studies. Molecules 2022, 27, 4549. [Google Scholar] [CrossRef]
  25. Clément, S.; Guyard, L.; Khatyr, A.; Knorr, M.; Rousselin, Y.; Kubicki, M.M.; Mugnier, Y.; Richeter, S.; Gerbier, P.; Strohmann, C. Synthesis, crystallographic and electrochemical study of ethynyl[2.2]paracyclophane-derived cobalt metallatetrahedranes. J. Organomet. Chem. 2012, 699, 56–66. [Google Scholar] [CrossRef]
  26. Jourdain, I.; Knorr, M.; Brieger, L.; Strohmann, C. Synthesis of Tris(arylphosphite)-ligated Cobalt(0) Complexes [Co2(CO)6{P(OAr)3}2], and their Reactivity towards Phenylacetylene and Diphenylacetylene. Adv. Chem. Res. 2020, 2. [Google Scholar] [CrossRef]
  27. Hase, Y. The infrared and Raman spectra of phthalimide, N-d-phthalimide and potassium phthalimide. J. Mol. Struct. 1978, 48, 33–42. [Google Scholar] [CrossRef]
  28. Chen, H.; Ai, Z.; Guo, L.; Yao, L.; Li, Y.; Gu, B.; Zhang, Y.; Liu, Y.A.; Tian, B.; Liao, X. Nickel-catalyzed enantioselective domino Heck/Sonogashira coupling for construction of C(sp)-C(sp3) bond-substituted quaternary carbon centers. Tetrahedron Chem. 2022, 2, 100021–100031. [Google Scholar] [CrossRef]
  29. Saha, M.; Muchmore, S.; van der Helm, D.; Nicholas, K.M. Cobalt-Mediated Cyclopentenone Annulation: An Approach to the Synthesis of Cyclocolorenone. J. Org. Chem. 1986, 51, 1960–1966. [Google Scholar] [CrossRef]
  30. Carniato, F.; Gatti, G.; Gervasio, G.; Marabello, D.; Sappa, E.; Secco, A. Reactions of Co2(CO)8 and of Co2(CO)6L (L = 3-pentyn-1-ol, 1,4-butyn-diol or 2-methyl-3-butyn-2-ol) with 2(diphenylphosphino)ethyl-triethoxysilane and tris(hydroxymethyl)phosphine for applications to new sol–gel materials. J. Organomet. Chem. 2009, 694, 4241–4249. [Google Scholar] [CrossRef]
  31. 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. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  32. Sheldrick, G. SHELXT-Integrated space-group and crystal-structure determination. Acta Crystallogr. A 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Sheldrick, G. Crystal structure refinement with SHELXL. Acta Crystallogr. C 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Scheme 1. Examples of some dicobaltatetrahedranes displaying biological activity (I, IIIV: antiproliferative activity against cancer cells; II: hormonally active compound).
Scheme 1. Examples of some dicobaltatetrahedranes displaying biological activity (I, IIIV: antiproliferative activity against cancer cells; II: hormonally active compound).
Molbank 2023 m1545 sch001
Scheme 2. Synthesis of the title compound 2.
Scheme 2. Synthesis of the title compound 2.
Molbank 2023 m1545 sch002
Figure 1. IR spectrum of compound 2 recorded in cyclohexane.
Figure 1. IR spectrum of compound 2 recorded in cyclohexane.
Molbank 2023 m1545 g001
Figure 2. 1H-NMR spectrum (400 MHz, CDCl3) of compound 2 at 25 °C. The asterisk * denotes the chloroform resonance.
Figure 2. 1H-NMR spectrum (400 MHz, CDCl3) of compound 2 at 25 °C. The asterisk * denotes the chloroform resonance.
Molbank 2023 m1545 g002
Figure 3. 13C{1H}-NMR spectrum (100.62 MHz, CDCl3) of compound 2 at 25 °C. The asterisk * denotes the CDCl3 resonance.
Figure 3. 13C{1H}-NMR spectrum (100.62 MHz, CDCl3) of compound 2 at 25 °C. The asterisk * denotes the CDCl3 resonance.
Molbank 2023 m1545 g003
Figure 4. Molecular structure of one of two independent molecules of 2. All H atoms are omitted for clarity. Selected bond lengths (Å) and angles (deg): Co1–Co2 2.4688(14), Co1–C2 1.980(6), Co1–C3 1.952(6), Co2–C2 1.984(6), Co2–C3 1.954(7), Co1–C16 1.812(7), Co1–C17 1.784(8), Co1–C18 1.831(7), Co2–C13 1.794(7), Co2–C14 1.812(7), Co2–C15 1.805(7), C2–C3 1.323(9); Co1–Co2–C3 50.75(15), Co2–Co1–C2 50.83(19), Co1–C2–Co2 77.0(2), Co1–C3–Co2 78.4(2).
Figure 4. Molecular structure of one of two independent molecules of 2. All H atoms are omitted for clarity. Selected bond lengths (Å) and angles (deg): Co1–Co2 2.4688(14), Co1–C2 1.980(6), Co1–C3 1.952(6), Co2–C2 1.984(6), Co2–C3 1.954(7), Co1–C16 1.812(7), Co1–C17 1.784(8), Co1–C18 1.831(7), Co2–C13 1.794(7), Co2–C14 1.812(7), Co2–C15 1.805(7), C2–C3 1.323(9); Co1–Co2–C3 50.75(15), Co2–Co1–C2 50.83(19), Co1–C2–Co2 77.0(2), Co1–C3–Co2 78.4(2).
Molbank 2023 m1545 g004
Figure 5. Presentation of the two structurally characterized dicobaltatetrahedranes (A) and (B) incorporating a -CH2-C≡C-Me motif.
Figure 5. Presentation of the two structurally characterized dicobaltatetrahedranes (A) and (B) incorporating a -CH2-C≡C-Me motif.
Molbank 2023 m1545 g005
Figure 6. View of the molecular cell of 2 showing the two pairwise arranged independent molecules.
Figure 6. View of the molecular cell of 2 showing the two pairwise arranged independent molecules.
Molbank 2023 m1545 g006
Table 1. Comparison of relevant bond lengths (Å) and angles (°) in 2 and the crystallographically characterized [Co2(CO)6(µ-RCH2C≡CCH3)] tetrahedranes A and B.
Table 1. Comparison of relevant bond lengths (Å) and angles (°) in 2 and the crystallographically characterized [Co2(CO)6(µ-RCH2C≡CCH3)] tetrahedranes A and B.
2AB
Co–Co2.4679(14)2.4688(9)2.4690(9)
C–Calkyne1.330(9)1.325(6)1.326(6)
Calkyne–CMe1.481(9)1.486(6)1.492(6)
Calkyne–CH21.499(9)1.500(7)1.493(6)
Co–Calkyne1.972(6)1.974(5)1.972(4)
Co–Cpseudo-equatorial1.815(7)1.823(5)1.820(5)
Co–Cpseudo-axial1.797(7)1.785(6)1.787(5)
CSD referenceThis workDOTKEC [29]HUKCIA [30]
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

Jourdain, I.; Knorr, M.; Charenton, T.; Strohmann, C.; Kirchhoff, J.-L.; Othman, M. (µ24-N-(2-Butynyl)phthalimide)(hexacarbonyl)dicobalt. Molbank 2023, 2023, M1545. https://doi.org/10.3390/M1545

AMA Style

Jourdain I, Knorr M, Charenton T, Strohmann C, Kirchhoff J-L, Othman M. (µ24-N-(2-Butynyl)phthalimide)(hexacarbonyl)dicobalt. Molbank. 2023; 2023(1):M1545. https://doi.org/10.3390/M1545

Chicago/Turabian Style

Jourdain, Isabelle, Michael Knorr, Tom Charenton, Carsten Strohmann, Jan-Lukas Kirchhoff, and Mohamed Othman. 2023. "(µ24-N-(2-Butynyl)phthalimide)(hexacarbonyl)dicobalt" Molbank 2023, no. 1: M1545. https://doi.org/10.3390/M1545

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