O slideshow foi denunciado.
Seu SlideShare está sendo baixado. ×

Reductive Cyclotrimerization of Carbon Monoxide to the Deltate Dianion by an Organometallic Uranium Complex

Anúncio
Anúncio
Anúncio
Anúncio
Anúncio
Anúncio
Anúncio
Anúncio
Anúncio
Anúncio
Anúncio
Anúncio
Carregando em…3
×

Confira estes a seguir

1 de 3 Anúncio

Mais Conteúdo rRelacionado

Diapositivos para si (20)

Anúncio

Semelhante a Reductive Cyclotrimerization of Carbon Monoxide to the Deltate Dianion by an Organometallic Uranium Complex (20)

Mais recentes (20)

Anúncio

Reductive Cyclotrimerization of Carbon Monoxide to the Deltate Dianion by an Organometallic Uranium Complex

  1. 1. below the È10-Mpc scales calculated here would tell us about density perturbations in photons (and baryons) in the early universe, even at scales smaller than the diffusion scale at recombination. In this sense, the magnetic field generated by this mechanism can be regarded as a fossil of density perturbations in the early universe, whose signature in photons and baryons has been lost. Therefore, this result provides the possibility of probing observations on how density perturbations in photons have evolved and been swept away at these small scales, where no one can, in principle, probe directly through photons. The amplitude of the cosmologically gen- erated magnetic fields is too small to be ob- served directly through polarization effects or synchrotron emission. However, magnetic fields with such small amplitude may be detected by gamma ray burst observations (25) through the delay of the arrival time of gamma ray photons due to the magnetic deflection of high-energy electrons responsible for such gamma ray photons (26, 27). We suggest that, because the weak magnetic fields should inevitably be generated from cosmological perturbations as presented here, and because they should exist all over the universe even in the intercluster fields, then the weak fields should be detectable by future high-energy gamma ray experiments, such as GLAST (the Gamma Ray Large Area Space Telescope) (28). Although the power of B increases as k2 on small scales (Fig. 2), the diffusion due to Coulomb scattering between electrons and protons damps the magnetic fields around k È 1012 Mpcj1 E(1 þ z)/(104)^7/4. Therefore, the energy density in magnetic fields remains finite. Because magnetic fields at smaller scales result from density perturbations in the early universe, one needs to take into consideration the high- energy effects neglected in the collision term: e.g., relativistic corrections for the energy of electrons and weak interactions along with the Compton scattering. These effects would be- come important when the temperature of the universe was above È1 MeV, which corre- sponds to the comoving wave number È105 Mpcj1. These effects can be safely neglected at the scales considered here. References and Notes 1. L. Biermann, A. Schlu¨ter, Phys. Rev. 82, 863 (1951). 2. L. Mestel, I. W. Roxburgh, Astrophys. J. 136, 615 (1962). 3. H. Hanayama et al., Astrophys. J. 633, 941 (2005). 4. L. M. Widrow, Rev. Mod. Phys. 74, 775 (2003). 5. N. Y. Gnedin, A. Ferrara, E. G. Zweibel, Astrophys. J. 539, 505 (2000). 6. R. M. Kulsrud, R. Cen, J. P. Ostriker, D. Ryu, Astrophys. J. 480, 481 (1997). 7. G. Davies, L. M. Widrow, Astrophys. J. 540, 755 (2000). 8. A. H. Guth, Phys. Rev. D 23, 347 (1981). 9. K. Sato, Mon. Not. R. Astron. Soc. 195, 467 (1981). 10. B. Ratra, Astrophys. J. 391, L1 (1992). 11. D. Lemoine, M. Lemoine, Phys. Rev. D 52, 1955 (1995). 12. K. Bamba, J. Yokoyama, Phys. Rev. D 69, 043507 (2004). 13. M. S. Turner, L. M. Widrow, Phys. Rev. D 37, 2743 (1988). 14. C. Caprini, R. Durrer, Phys. Rev. D 65, 023517 (2002). 15. E. R. Harrison, Mon. Not. R. Astron. Soc. 147, 279 (1970). 16. H. Lesch, M. Chiba, Astron. Astrophys. 297, 305 (1995). 17. Z. Berezhiani, A. D. Dolgov, Astropart. Phys. 21, 59 (2004). 18. S. Matarrese, S. Mollerach, A. Notari, A. Riotto, Phys. Rev. D 71, 043502 (2005). 19. K. Takahashi, K. Ichiki, H. Ohno, H. Hanayama, Phys. Rev. Lett. 95, 121301 (2005). 20. C. L. Bennett et al., Astrophys. J. 148, (suppl.), 1 (2003). 21. A.-C. Davis, M. Lilley, O. Tornkvist, Phys. Rev. D 60, 021301 (1999). 22. A. A. Starobinsky, Phys. Lett. B 117, 175 (1982). 23. H. Maki, H. Susa, Astrophys. J. 609, 467 (2004). 24. J. Silk, Astrophys. J. 151, 459 (1968). 25. R. Plaga, Nature 374, 430 (1995). 26. E. Waxman, P. Coppi, Astrophys. J. 464, L75 (1996). 27. S. Razzaque, P. Meszaros, B. Zhang, Astrophys. J. 613, 1072 (2004). 28. N. Gehrels, P. Michelson, Astropart. Phys. 11, 277 (1999). 29. K. Yamamoto, N. Sugiyama, H. Sato, Astrophys. J. 501, 442 (1998). 30. K.T. and K.I. are supported by a Grant-in-Aid for the Japan Society for the Promotion of Science Fellows and are research fellows of the Japan Society for the Promotion of Science. N.S. is supported by a Grant-in-Aid for Scientific Research from the Japanese Ministry of Education (no. 17540276). This work was supported in part by the Kavli Institute for Cosmological Physics through the grant NSF PHY-0114422. Supporting Online Material www.sciencemag.org/cgi/content/full/1120690/DC1 SOM Text References 28 September 2005; accepted 12 December 2005 Published online 5 January 2006; 10.1126/science.1120690 Include this information when citing this paper. Reductive Cyclotrimerization of Carbon Monoxide to the Deltate Dianion by an Organometallic Uranium Complex Owen T. Summerscales,1 F. Geoffrey N. Cloke,1 * Peter B. Hitchcock,1 Jennifer C. Green,2 Nilay Hazari2 Despite the long history of the Fischer-Tropsch reaction, carbon monoxide has proven remarkably resistant to selective homologation under mild conditions. Here, we find that an organouranium(III) complex induces efficient reductive trimerization of carbon monoxide at room temperature and pressure. The result is a triangular, cyclic C3O3 2–, or deltate, dianion held between two uranium(IV) units. The bonding within the C3O3 2– unit and its coordination to the two U centers have been analyzed by x-ray diffraction and density functional theory computational studies, which show a stabilizing C-C agostic interaction between the C3 core and one U center. Solution nuclear magnetic resonance studies reveal a rapid equilibration of the deltate unit between the U centers. C arbon monoxide, usually obtained from coal or natural gas, is an im- portant industrial feedstock for the production of hydrocarbons and oxygenates, especially in times of limited crude oil sup- ply, through the Fischer-Tropsch process. This process uses CO/H2 as the feedstock and is catalyzed by both homogeneous and heterogeneous systems (1). Metal-catalyzed coupling of CO with unsaturated organic sub- strates (e.g., hydroformylation catalysis) repre- sents a further industrially important method of C-C bond formation (2). The concept of reductively homologating CO directly to form larger units of (CO)n is an appealing one; however, the strength of the CO triple bond has made this route difficult to achieve effec- tively. The cyclic, aromatic, oxocarbon anions CnOn 2– (n 0 3 to 6 in Scheme 1) have fas- cinated organic chemists for many years (3, 4) and, in principle, are accessible by such a route. This would provide a facile synthesis of fused carbon ring systems as building blocks for further elaboration to more complex or- ganic molecules (e.g., possible ring expansion of A for prostaglandin synthesis). The formation of salts of the croconate and rhodizonate dianions—C and D—from the reduction of CO by molten alkali metals Ewhich also affords salts of the linear C2O2 2– ethyne- diolate dianion (5)^ was first reported in the early part of the 19th century (6), and there is some nuclear magnetic resonance (NMR) evidence for formation of trace amounts of the deltate dianion A in a complex mixture of products from the reaction of CO with Na-K alloy in tetrahydrofuran (THF) (7). However, salts of the trimer A have eluded crystallo- graphic characterization or selective metal- mediated synthesis. The squarate dianion B has been generated by electrochemical meth- ods using very high pressures of CO (8), and surface catalysis routes using alkaline earth oxides yield mixtures of various (CO)n 2– (n 0 2 to 6) species (9). The idea of using low-valent f-element complexes, which combine high reduction po- 1 Department of Chemistry, School of Life Sciences, Univer- sity of Sussex, Brighton BN1 9QJ, UK. 2 Inorganic Chemistry Laboratory, South Parks Road, Oxford OX1 3QR, UK. *To whom correspondence should be addressed. E-mail: f.g.cloke@sussex.ac.uk REPORTS www.sciencemag.org SCIENCE VOL 311 10 FEBRUARY 2006 829
  2. 2. tentials with solubilizing ligands, to give iso- lable (CO)n 2– species was demonstrated by Evans et al., who reported that ESm(h-Cp*)2(THF)2^ (Cp* is pentamethylcyclopentadienyl) will react under a 90 pounds per square inch over- pressure of CO to give a doubly charged ketene carboxylate unit (O2CCCO)2– bound to samarium (10). Here, we extend that idea to selectively synthesize and structurally char- acterize the deltate dianion A directly from CO under very mild conditions. Specifically, reduc- tive cyclotrimerization of CO is induced by an organometallic U(III) complex. There is considerable current interest in the binding of small molecules to U(III) centers, e.g., CO (11), CO2 (12), and N2 (13). As part of our continuing investigation into reduction chemistry by low-valent uranium complexes, we recently reported the reversible binding and concomitant double reduction of dinitrogen by the pentalene complex EU(h-C8H4 .)(h-Cp*)^ (14) E. is 1,4-bis(tri-isopropylsilyl)^. We ex- plored the reactivity of the related silylated cyclooctatetraene (COT) complex with CO; the use of bulky tri-isopropylsilyl substituents on the COT ring provides steric protection for reactive metal centers and imparts high crys- tallinity to the derived metal complexes. The THF adduct EU(h-COT.)(h-Cp*)(THF)^ 1 was obtained by a route similar to that employed for the unsilylated analog EU(h-COT)(h-Cp*)(THF)^ (15); reaction of UI3 with KCp* in THF yielded EUI2Cp*(THF)3^, which was treated in situ with 0.8 equivalent of K2ECOT.^ in THF to give 1 as a dark purple, crystalline material in 39% overall yield after recrystallization from pen- tane at –50-C (Scheme 2). Compound 1 has been structurally characterized by x-ray diffrac- tion (16). Compound 1 reacts with ambient pressures (1 bar) of CO in pentane to give the dimeric U(IV) complex 2 in 40% isolated yield. Crys- tallization from diethyl ether at –50-C gave dark red needles of 2 suitable for x-ray diffrac- tion, which revealed a reductively homologated CO trimer held between the two uranium(IV) centers (Scheme 2 and Fig. 1) (17). The structure shows a cyclic (C3O3) moiety bound through the oxygen atoms in a bridg- ing h1:h2 fashion to the two uranium cen- ters. Metal-ligand bond lengths within the EU(h-COT.)(h-Cp*)^ fragments reflect the steric hindrance around the respective uranium cen- ters, i.e., average U-COT. and U-Cp* ring carbon distances are slightly longer around the h2-bound U1 center (2.696 ) and 2.775 ), respectively) than the h1-bound U2 (2.656 ) and 2.740 )). We propose an oxidation state of (IV) for both U centers in 2 on the basis of the density functional theory (DFT) computational studies below. Oxidation from U(III) in 1 to U(IV) in 2 is not, however, reflected in the structural parameters; such changes were similarly lacking in the conversion of the pentalene complex EU(h-C8H4 .)(h-Cp*)^ to EU(h-C8H4 .)(h-Cp*)^2(m-h2:h2-N2) (14), presum- ably due to steric congestion in the uranium(IV) dimers. Because the other bond distances and angles associated with the two EU(h-COT.)(h-Cp*)^ fragments are unexceptional, we now focus on the U(C3O3)U core. The (C3O3) unit is planar, with the two uranium centers lying slightly above (U2, 0.0906 )) and below (U1, 0.1747 )) this plane. The h1-uranium-oxygen distance (U2-O3) of 2.183(3) ) is slightly longer than those found in U(IV) aryloxides Ae.g., 2.120 ) in EU(h-Cp*)3(OPh)^Z (Ph, is phenyl) (18), but U1-O1 E2.516(3) )^ and U1-O2 E2.484(3) )^ are significantly longer. The C-O bond lengths Ewhich lie between typical values for single C-O (1.43 )) and double C0O (1.21 )) bonds^ show complementary variations to the U-O bond lengths: C3-O3 is somewhat longer than C1-O1 and C2-O2 (Fig. 1). The triangular C3 skeleton in 2 is also noticeably distorted (Fig. 1), with two short C-C bonds and one long one, the latter showing some interaction with U1. We probed the bonding in this deltate diuranium com- plex using DFT. The unsubstituted compound EU2(h-COT)2(h-Cp)2(m-h1:h2-C3O3)^ II was used as a model for compound 2. The ex- perimental coordinates determined from the solid state structure of 2 were used to de- scribe the initial geometry of II, and a geom- etry optimization was performed with no symmetry constraints. In addition, a fragment analysis was conducted in which II was divided into two U(h-COT)(h-Cp) fragments and a C3O3 fragment. It is often difficult to achieve self-consistent field (SCF) convergence in calculations involving open-shell actinides, because there are a large number of orbitals close to the Fermi surface of the molecule. In this case, the geometry optimization was carried out as a spin-restricted calculation, with electron density smeared over the frontier orbitals to assist SCF convergence (16). As a result, the frontier orbitals have fractional occupancy; however, the molec- ular geometry is typically not very sensitive to the exact description of the f electrons (19). There was good agreement between the experimental and calculated bond distances and angles in the U(C3O3)U core of II (table S1), demonstrating that DFT can accurately predict the structure of 2. The calculated struc- ture reproduced the distortions in the C3 core, with two short C-C bonds (C2-C3, 1.41 ); C1-C3, 1.41 )) and one long bond (C1-C2, 1.47 )). The C3O3 unit was planar, and the C-O and U-O bond distances followed the pattern observed in the solid-state structure. Scheme 1. Fig. 1. (Left) The mo- lecular structure of 2 (thermal ellipsoids at 50%). (Right) View of core structure (dis- tances in A˚; values in parentheses are errors in the last significant digit). REPORTS 10 FEBRUARY 2006 VOL 311 SCIENCE www.sciencemag.org830
  3. 3. There were minor differences between the calculated and experimental U-COT and U-Cp bond lengths and angles, suggesting that these interactions are controlled by steric inter- actions; however, the discrepancy is not suf- ficiently great to invalidate the model. The calculations indicate that each U is best de- scribed as having two electrons localized in 5f orbitals. Thus, the U configuration is consistent with U(IV); chemical plausibility and theoret- ical studies (20) also suggest that the deltate structure is not likely to be stable in a neutral state, whereas it is known in a dianionic 2-state. The COT and Cp ligands bind to the U centers as expected, with primarily a p interaction be- tween U and Cp and a d interaction between U and COT. Previous work has shown that the highest occupied molecular orbitals (MOs) of the re- laxed (C3O3)2– anion (in D3h symmetry) are a p orbital of a2µ symmetry (which confers aro- maticity on the dianion) and a degenerate pair of C-C bonding orbitals (e¶), which have C-O antibonding character (21). On binding to the two U fragments, the symmetry of the (C3O3)2– anion is lowered, but the MOs of the distorted fragment are clearly related to the MOs of the relaxed parent fragment in D3h symmetry. Fragment analysis indicates that four MOs of the deltate anion have a notable interaction with the two U-containing fragments (table S2 and fig. S1). The primary interaction involves donation of electron density from the oxygen atoms of the (C3O3)2– ligand to the U atoms. However, a more complex interaction occurs between the lengthened C-C bond and a U f orbital. Figure 2 shows a representation of this parent (C3O3)2– MO, which shows that the orbital primarily consists of the lone pairs on O1 and O2 and a bonding interaction between C1 and C2. The parent orbital is bonding be- tween C1 and C2 and is antibonding with re- spect to the lone pairs on both O1 and O2. This orbital interacts with an f orbital on U1 to form an MO of II (which is also shown in Fig. 2). This MO comprises not only a bonding inter- action between U1 and O1, O2 but also (and of more importance) a bonding interaction be- tween U1 and C1, C2. The net effect of this second interaction is to weaken the C1-C2 bond, because some of the electron density between C1 and C2 is shared with U1, and to strengthen the C1-O1 and C2-O2 bonds, because the antibonding electron density is de- creased. Thus, a C-C agostic interaction ap- pears to be responsible for the distortion in the C3 core observed in 2. The highly nodal char- acteristics of the f orbital on U enable such an interaction and, thus, actinides may be par- ticularly suited to stabilize such CO homologs. This finding suggests that tuning the ligand environment around U(III) may give access to higher homologs of the deltate dianion. 1H NMR solution studies of 2 in d8-toluene do not reflect the asymmetry present in the solid-state structure (Fig. 1). Only single en- vironments for the Cp* ligands and the C8H6 . ligands are observed at 25-C, indicating a rapid (milliseconds on the NMR time scale) fluxional process leading to equilibration of the deltate unit between U1 and U2 at this temperature. This finding was confirmed by a study of the isotopically labeled complex EU(h-COT.)(h-Cp*)^2 (m-h1 :h2 -1 3 C3 O3 ) 2-13CO, prepared by reaction of 1 with 13CO: 2-13CO displays a singlet (n1/2 12 Hz) at d 225 parts per million for the deltate carbons in the 13C NMR spectrum at 25-C. At –100-C, both 1H and 13C spectra for 2-13CO display the expected changes in chemical shifts with tem- perature for a paramagnet, but, apart from considerable broadening of the signals, there is no evidence for freezing out of the fluxional process at this temperature. Hence, these data do not allow us to distinguish between the two likely mechanisms for the equilibration of the C3O3 unit in 2, i.e., a Bwindshield wiper[ mo- tion of O1 (or O2) between U1 and U2, versus a rotational Bpropeller[ mechanism involving all three oxygen atoms. References and Notes 1. C. K. Rofer-DePoorter, Chem. Rev. 81, 447 (1981). 2. B. Cornils, W. A. Herrmann, Applied Homogenous Catalysis with Orgaonmetallic Compounds (VCH, Weinheim, Germany, 1996). 3. R. West, R. Eggerding, J. Am. Chem. Soc. 98, 3641 (1976). 4. G. Seitz, P. Imming, Chem. Rev. 92, 1227 (1992). 5. W. Buchner, Helv. Chim. Acta 46, 2111 (1963). 6. L. Gmelin, Ann. Phys. Chem. 4, 31 (1825). 7. P. W. Lednor, P. C. Versloot, J. Chem. Soc. Chem. Commun. 285 (1983). 8. G. Silvestri, S. Gambino, G. Filardo, G. Spadaro, L. Palmisano, Electrochim. Acta 23, 413 (1978). 9. S. Coluccia, E. Garrone, E. Guglielminotti, A. Zecchina, J. Chem. Soc., Faraday Trans. 1 77, 1063 (1981). 10. W. J. Evans, J. W. Grate, L. A. Hughes, H. Zhang, J. L. Atwood, J. Am. Chem. Soc. 107, 3728 (1985). 11. I. Castro-Rodriguez, K. Meyer, J. Am. Chem. Soc. 127, 11242 (2005). 12. I. Castro-Rodriguez, H. Nakai, L. N. Zakharov, A. L. Rheingold, K. Meyer, Science 305, 1757 (2004). 13. W. J. Evans, S. A. Kozimor, J. W. Ziller, J. Am. Chem. Soc. 125, 14264 (2002). 14. F. G. N. Cloke, P. B. Hitchcock, J. Am. Chem. Soc. 124, 9352 (2002). 15. A. R. Schake et al., Organometallics 12, 1497 (1993). 16. Materials and methods are available as supporting material on Science Online. 17. Data collection was performed at 173(2) K on an Enraf-Nonius Kappa charge-coupled device diffractometer with graphite-monochromated MoKa radiation (l 0 0.71073 A˚). The molecular structure was solved by direct methods and refined on F2 by full-matrix least squares techniques. For 2.Et2O: triclinic, P¯1111 (No. 2), a 0 10.7193(2) A˚, b 0 16.6281(2) A˚, c 0 23.2496(4) A˚, a 0 99.768(1)-, b 0 95.816(1)-, g 0 91.500(1)-, V 0 4058.97(11) A˚3, Z 0 2, R1 0 0.034, wR2 0 0.070. 18. M. R. Spirlet, J. Rebizant, C. Apostolidis, G. van den Bossche, B. Kanellakopolus, Acta Crystallogr., Sect. C, Struct. Commun. C46, 2318 (1990). 19. F. G. N. Cloke, J. C. Green, N. Kaltsoyannis, Organometallics 23, 832 (2004). 20. P. v. R. Schleyer, K. Najafian, B. Kiran, H. Jiao, J. Org. Chem. 65, 426 (2000). 21. R. West, D. Eggerding, J. Perkins, D. Handy, E. C. Tuazon, J. Am. Chem. Soc. 101, 1710 (1979). 22. We thank the Engineering and Physical Sciences Research Council for financial support of this work. Metrical data for 1 and 2 are available from the Cambridge Crystallographic Data Centre under reference numbers CCDC-292279 and CCDC-292280, respectively. Supporting Online Material www.sciencemag.org/cgi/content/full/311/5762/829/DC1 Materials and Methods SOM Text Fig. S1 Tables S1 to S3 References 25 October 2005; accepted 12 December 2005 10.1126/science.1121784 Fig. 2. (A) Key orbital of the distorted deltate di- anion involved in an agostic interaction with the U cen- ter. Its parentage is one of the degenerate e¶ C-C bond- ing orbitals of (C3O3)2–. (B) MO showing the agostic interaction between the deltate dianion and an f orbital on U(I). Details of these orbitals are given in table S2. Scheme 2. REPORTS www.sciencemag.org SCIENCE VOL 311 10 FEBRUARY 2006 831

×