Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 21
Filtrar
Mais filtros

Base de dados
Tipo de documento
Assunto da revista
Intervalo de ano de publicação
1.
Angew Chem Int Ed Engl ; 61(8): e202116565, 2022 Feb 14.
Artigo em Inglês | MEDLINE | ID: mdl-34904338

RESUMO

The crystal structures of [NH3 F]+ [CF3 SO3 ]- , [NH2 F2 ]+ [SbF6 ]- , and [N2 F3 ]+ [Sb3 F16 ]- have been determined, representing the first structural characterizations of these simple fluoro-nitrogen cations. The influences of the hybridization of the central nitrogen atom and of the number of fluorine substituents on the N-F bond lengths are evaluated for the series N2 F+ , N2 F3 + , NF2 O+ , NH3 F+ , NH2 F2 + , and NF4 + . It is shown that the N-F bond length decreases from 1.40 Što 1.26 Šwith increasing fluorine substitution and increasing s-character of the nitrogen atom, and that unusual N-F bond lengths reported in the previous literature are caused by disorder problems.

2.
Inorg Chem ; 49(15): 6823-33, 2010 Aug 02.
Artigo em Inglês | MEDLINE | ID: mdl-20465274

RESUMO

N(2)F(+) salts are important precursors in the synthesis of N(5)(+) compounds, and better methods are reported for their larger scale production. A new, marginally stable N(2)F(+) salt, N(2)F(+)Sn(2)F(9)(-), was prepared and characterized. An ordered crystal structure was obtained for N(2)F(+)Sb(2)F(11)(-), resulting in the first observation of individual N[triple bond]N and N-F bond distances for N(2)F(+) in the solid phase. The observed N[triple bond]N and N-F bond distances of 1.089(9) and 1.257(8) A, respectively, are among the shortest experimentally observed N-N and N-F bonds. High-level electronic structure calculations at the CCSD(T) level with correlation-consistent basis sets extrapolated to the complete basis limit show that cis-N(2)F(2) is more stable than trans-N(2)F(2) by 1.4 kcal/mol at 298 K. The calculations also demonstrate that the lowest uncatalyzed pathway for the trans-cis isomerization of N(2)F(2) has a barrier of 60 kcal/mol and involves rotation about the N=N double bond. This barrier is substantially higher than the energy required for the dissociation of N(2)F(2) to N(2) and 2 F. Therefore, some of the N(2)F(2) dissociates before undergoing an uncatalyzed isomerization, with some of the dissociation products probably catalyzing the isomerization. Furthermore, it is shown that the trans-cis isomerization of N(2)F(2) is catalyzed by strong Lewis acids, involves a planar transition state of symmetry C(s), and yields a 9:1 equilibrium mixture of cis-N(2)F(2) and trans-N(2)F(2). Explanations are given for the increased reactivity of cis-N(2)F(2) with Lewis acids and the exclusive formation of cis-N(2)F(2) in the reaction of N(2)F(+) with F(-). The geometry and vibrational frequencies of the F(2)N=N isomer have also been calculated and imply strong contributions from ionic N(2)F(+) F(-) resonance structures, similar to those in F(3)NO and FNO.

3.
Chem Commun (Camb) ; (47): 4992-4, 2007 Dec 21.
Artigo em Inglês | MEDLINE | ID: mdl-18049730

RESUMO

New fluorinated polyhedral oligomeric silsesquioxane (F-POSS) structures possessing a high degree of hydrophobicity have been prepared via a facile corner-capping methodology.

4.
Chem Commun (Camb) ; (16): 1712-3, 2002 Aug 21.
Artigo em Inglês | MEDLINE | ID: mdl-12196962

RESUMO

The conversion of ketones into geminal bis(difluoramines) can be achieved under mild two-phase reaction conditions by employing triphenylmethyldifluoramine as an in situ source of difluoramine.

5.
Inorg Chem ; 36(13): 2730-2745, 1997 Jun 18.
Artigo em Inglês | MEDLINE | ID: mdl-11669905

RESUMO

Acyclic tertiary amines such as triethylamine and tri-n-propylamine used as HCl scavengers in nucleophilic substitution reactions of cyclic chlorocarbaphosphazenes [N(3)PC(2)Cl(4) (I) and N(3)P(2)CCl(5) (II)] with (CF(2))(n)()(CF(2)CH(2)OH)(2) [n = 0 (III) or 1 (IV)] are found to undergo a facile C-N bond cleavage with the regiospecific substitution of the dialkylamino groups on the ring carbon atoms of the carbaphosphazene. In the case of cyclic amines such as 1-methylpiperidine and 4-methylmorpholine, the cleavage was found to occur regiospecifically at the N-CH(3) bond, resulting in the ring substitution of the cyclic secondary amino group on the dicarbaphosphazene ring carbon atoms. The polyfluoro diol III forms a spirocyclic ring exclusively at the phosphorus site in compounds [CF(2)CH(2)O](2)PN(3)C(2)[N(C(2)H(5))(2)](2) (1), [CF(2)CH(2)O](2)PN(3)C(2)[N(C(3)H(7))(2)](2) (2), [CF(2)CH(2)O](2)PN(3)C(2)[NCH(2)(CH(2))(3)CH(2)](2) (3), and [CF(2)CH(2)O](2)PN(3)C(2)[N(CH(2))(2)O(CH(2))(2)](2) (5) along with the formation of carbon-substituted carbaphosphazenes, Cl(2)PN(3)C(2)[NCH(2)(CH(2))(3)CH(2)](2) (4) and Cl(2)PN(3)C(2)[N(CH(2))(2)O(CH(2))(2)](2) (6). Reaction of II with III in the presence of triethylamine affords the dispiro product [CF(2)CH(2)O)(2)](2)P(2)N(3)C[N(C(2)H(5))(2)] (7), which crystallizes in a polar orthorhombic space group, Cmc2(1). Upon refluxing of I or II with R(3)N (R = C(2)H(5), n-C(3)H(7)) in toluene, the amino-substituted carbaphosphazenes, Cl(2)PN(3)C(2)[N(C(2)H(5))(2)](2) (8), Cl(4)P(2)N(3)C[N(C(2)H(5))(2)] (9), and Cl(4)P(2)N(3)C[N(n-C(3)H(7))(2)] (10) are obtained in good yields. Hydrolysis of 8 leads to the formation of Cl(O)PN(H)N(2)C(2)[N(C(2)H(5))(2)](2) (11). When lithium salts of the fluoro diols III and IV are reacted with I or II in diethyl ether, the P-Cl bond is selectively substituted, yielding the spirocyclic [CF(2)CH(2)O](2)PN(3)C(2)Cl(2) (12), [CF(2)(CF(2)CH(2)O)(2)]PN(3)C(2)Cl(2) (13), [CF(2)CH(2)O](2)P(2)N(3)CCl(3) (14), and [(CF(2)CH(2)O)(2)](2)P(2)N(3)CCl (15). The C-Cl bonds in 12 and 14 were easily substituted by their reaction with 4-FC(6)H(4)XNa (X = O or S) to form [CF(2)CH(2)O](2)PN(3)C(2)[4-FC(6)H(4)O](2) (16) and [CF(2)CH(2)O](2)PN(3)C(2)[4-FC(6)H(4)S](2) (17). Reactions of 12 and 13 with (CH(3))(3)SiN(CH(3))(2) under mild conditions result in the elimination of (CH(3))(3)SiCl along with the formation of [CF(2)CH(2)O](2)PN(3)C(2)[N(CH(3))(2)](2) (18). The X-ray analyses of 13 and 18 represent the first examples of eight-membered spirocyclic phosphazenes. The thermal behavior of II, 9, 10, 14, and15 has also been investigated at 120 degrees C. Single-crystal X-ray diffraction studies were carried out for 1, 2, 7, 9-14, and 16-19, and these compounds are also characterized using IR, (1)H, (13)C, (19)F and (31)P NMR spectroscopy, MS, and elemental analysis.

6.
Inorg Chem ; 35(6): 1590-1601, 1996 Mar 13.
Artigo em Inglês | MEDLINE | ID: mdl-11666378

RESUMO

Di-tert-butyl peroxide initiated free radical addition of THF to various fluorinated alkenes (CF(2)=CH(2), CF(2)=CFH, CH(2)=CHCF(3), CF(2)=CFCF(3), CF(2)=CFC(5)F(11), CF(2)=CFOCF(2)CF(CF(3))OCF(2)CF(2)SO(2)F) gives either bidirectional addition products [CH(2)CH(2)CH(2)OCH(CF(2)CH(3)) (1), CH(2)CH(2)CH(2)OCH(CH(2)CHF(2)) (2), CH(2)CH(2)CH(2)OCH(CF(2)CH(2)F) (3), and CH(2)CH(2)CH(2)OCH(CFHCHF(2)) (4)] or unidirectional products [CH(2)CH(2)CH(2)OCH(CH(2)CH(2)CF(3)) (5), CH(2)CH(2)CH(2)OCH(CF(2)CHFCF(3)) (6), CH(2)CH(2)CH(2)OCH(CF(2)CHFC(5)F(11)) (7), and CH(2)CH(2)CH(2)OCH(CF(2)CHFOCF(2)CF(CF(3))OCF(2)CF(2)SO(2)F) (8)] depending on the structure of the alkene. Reaction of dioxane with CF(2)=CFOCF(2)CF(CF(3))OCF(2)CF(2)SO(2)F gives a single product, CH(2)OCH(2)CH(2)OCH(CF(2)CHFOCF(2)CF(CF(3))OCF(2)CF(2)SO(2)F) (9). In the case of hexafluoropropene or perfluoroallylbenzene, reaction with an excess of tetrahydrofuran gives only the monosubstituted products CH(2)CH(2)CH(2)OCH(CF(2)CHFCF(3)) (6) and CH(2)CH(2)CH(2)OCH(CF(2)CFHCF(2)C(6)F(5)) (11) respectively. When tetrahydrofuran is reacted with a 3:1 molar excess of the same two perfluoroalkenes, the disubstituted products CH(2)CH(2)CH(CF(2)CHFCF(3))OCH(CF(2)CHFCF(3)) (10) and CH(2)CH(2)CH(CF(2)CFHCF(2)C(6)F(5))OCH(CF(2)CFHCF(2)C(6)F(5)) (12) are formed respectively. When 18-crown-6 is reacted in the same way with fluoroalkenes in a 1:1 molar ratio, the monosubstituted products 18-crown-6-CH(2)CH(2)CF(3) (13), 18-crown-6-CF(2)CHFCF(3) (14), 18-crown-6-CF(2)CFHCF(2)C(6)F(5) (15), and 18-crown-6-CF(2)CHFOCF(2)CF(CF(3))OCF(2)CF(2)SO(2)F (16) are obtained. Polyfluorinated 18-crown-6 products containing three and two polyfluroalkyl/aryl groups are prepared when 18-crown-6 is reacted with a 3:1 molar excess of perfluoropropene and perfluoroallylbenzene to give 18-crown-6-(CF(2)CHFCF(3))(3) (17) and 18-crown-6-(CF(2)CFHCF(2)C(6)F(5))(2) (18), respectively. (Pentafluorophenoxy)trimethylsilane reacts with 11 in the presence of a catalytic amount of cesium fluoride to give compound 19, CH(2)CH(2)CH(2)OCH(CF(2)CFHCF(2)C(6)F(4)OC(6)F(5)). Two molecules of 11 are bridged by reaction with Me(3)SiOCH(2)CF(2)CF(2)CF(2)CH(2)OSiMe(3) to give CH(2)CH(2)CH(2)OCHCF(2)CFHCF(2)C(6)F(4)OCH(2)CF(2)CF(2)CF(2)CH(2)OC(6)F(4)CF(2)CHFCF(2)CHOCH(2)CH(2)CH(2) (20), while 12 forms the macroheterocycle OCHCH(2)CH(2)CHCF(2)CFHCF(2)C(6)F(4)OCH(2)CF(2)CF(2)CF(2)CH(2)OC(6)F(4)CF(2)CFHCF(2) (21) under similar reaction conditions. The lanthanum triflate complexes of 18-crown-6 (22) and 18-crown-6(CF(2)CFHCF(3)) (23) were prepared and the structures were obtained via single-crystal X-ray analysis. Although crystals suitable for single-crystal X-ray analysis could not be formed, lanthanum triflate complexes were formed with polyfluorinated ethers 15 and 16 to give the fluorinated complexes La(OSO(2)CF(3))(3)(18-crown-6-CF(2)CFHCF(2)C(6)F(5))(H(2)O) (24) and [La(OCH(2)CH(2))(5)OCH(2)CHCF(2)CFHOCF(2)C(CF(3))FOCF(2)CF(2)SO(2)F](3+)[CF(3)SO(3)(-)](3) (25) respectively. The acid salt La[N(SO(2)CF(3))(2)](3) (26) was also prepared and characterized, and reacted with dibenzo-18-crown-6 to give the complex dibenzo-18-crown-6-La[N(SO(2)CF(3))(2)](3) (27).

7.
Inorg Chem ; 37(15): 3734-3746, 1998 Jul 27.
Artigo em Inglês | MEDLINE | ID: mdl-11670473

RESUMO

THF/toluene solutions of phosphonic or phosphoric acids were reacted with (t)Bu(3)Ga at low temperature to yield the cyclic dimers [(t)Bu(2)GaO(2)P(OH)R](2) (R = Ph, Me, (t)Bu, H, OH; 1-5). Poor crystallinity and variable thermal stabilities of 1-5 necessitated derivatization with Me(3)SiNMe(2) to yield [(t)Bu(2)GaO(2)P(OSiMe(3))R](2) (R = Ph, Me, (t)Bu, H, OSiMe(3); 6-10), which were more amenable to purification and characterization. In solution, trans isomers were predominant for 6 and 7 at ambient temperature, whereas the cis isomer of 8 was predominant. NMR spectroscopy demonstrated cis-trans interconversion for 6-8 and crossover experiments showed interconversion to occur by, or be accompanied with, an intermolecular exchange process. Thermolysis of 3 in refluxing toluene yielded the cluster [((t)BuGa)(2)((t)Bu(2)Ga)(O(3)P(t)Bu)(2){O(2)P(OH)(t)Bu}] (11), which was converted to [((t)BuGa)(2)((t)Bu(2)Ga)(O(3)P(t)Bu)(2){O(2)P(OSiMe(3))(t)Bu}] (12) with Me(3)SiNMe(2). Thermolysis of 1-3 in refluxing diglyme, or solid-state pyrolysis at 250 degrees C in vacuo, yielded [(t)BuGaO(3)PR](4) (R = Ph, (t)Bu, Me; 13-15). The gallophosphate [(t)BuGaO(3)P(OSiMe(3))](4) (16) was similarly obtained by reaction of (t)Bu(3)Ga with H(3)PO(4) in refluxing diglyme, followed by trimethylsilylation with Me(3)SiNMe(2). Compounds 13-16 possess cuboidal Ga(4)P(4)O(12) cores analogous to double-four-ring secondary building units in the gallophosphates cloverite, gallophosphate-A, and ULM-5. The thermal, hydrolytic, and oxidative stabilities of 13-16 are discussed, as are observed intermolecular exchange processes. In addition to characterization of 1-16 by multinuclear ((1)H, (13)C, (31)P) NMR spectroscopy, infrared spectroscopy, mass spectrometry, and elemental analysis, molecular structures for compounds 6, 8, 10, 12, 14, 15, and 16 were determined by X-ray crystallography.

8.
Angew Chem Int Ed Engl ; 37(24): 3394-3397, 1998 Dec 31.
Artigo em Inglês | MEDLINE | ID: mdl-29711302

RESUMO

An unprecedented coupling reaction of heteroatom-containing tripyrranes leads to the formation of core-modified sapphyrins 1 and 2, which self-assemble in the solid state to form supramolecular ladders. Weak C-H⋅⋅⋅S and C-H⋅⋅⋅Se hydrogen-bonding interactions in addition to C-H⋅⋅⋅N hydrogen bonds are responsible for the observed structures.

13.
Science ; 323(5920): 1461-4, 2009 Mar 13.
Artigo em Inglês | MEDLINE | ID: mdl-19286551

RESUMO

Halogen bonding has increasingly facilitated the assembly of diverse host-guest solids. Here, we show that a well-known class of organic salts, bis(trimethylammonium) alkane diiodides, can reversibly encapsulate alpha,omega-diiodoperfluoroalkanes (DIPFAs) through intermolecular interactions between the host's I- anions and the guest's terminal iodine substituents. The process is highly selective for the fluorocarbon that forms an I-...I(CF2)mI...I- superanion that is matched in length to the chosen dication. DIPFAs that are 2 to 12 carbons in length (common industrial intermediates) can thereby be isolated from mixtures by means of crystallization from solution upon addition of the dissolved size-matched ionic salt. The solid-state salts can also selectively capture the DIPFAs from the vapor phase, yielding the same product formed from solution despite a lack of porosity of the starting lattice structure. Heating liberates the DIPFAs and regenerates the original salt lattice, highlighting the practical potential for the system in separation applications.

14.
Inorg Chem ; 47(9): 3617-24, 2008 May 05.
Artigo em Inglês | MEDLINE | ID: mdl-18345619

RESUMO

Ionic liquid azides from azidoethyl, alkyl, and alkenyl substituted derivatives of 1,2,4- and 1,2,3-amino-triazoles were prepared and examined for the first time to investigate their structural and physical properties. All reported salts possess melting points below 100 degrees C. The unique character of these newly discovered ionic liquid azides is based upon the fact that these molecules are not simple protonated salts like previously reported substituted hydrazinium azides. The presence of quaternary nitrogen confers both thermal stability and negligible volatility.

15.
Inorg Chem ; 43(5): 1659-66, 2004 Mar 08.
Artigo em Inglês | MEDLINE | ID: mdl-14989658

RESUMO

The first example of a mononuclear diphosphanidoargentate, bis[bis(trifluoromethyl)phosphanido]argentate, [Ag[P(CF(3))(2)](2)](-), is obtained via the reaction of HP(CF(3))(2) with [Ag(CN)(2)](-) and isolated as its [K(18-crown-6)] salt. When the cyclic phosphane (PCF(3))(4) is reacted with a slight excess of [K(18-crown-6)][Ag[P(CF(3))(2)](2)], selective insertion of one PCF(3) unit into each silver phosphorus bond is observed, which on the basis of NMR spectroscopic evidence suggests the [Ag[P(CF(3))P(CF(3))(2)](2)](-) ion. On treatment of the phosphane complexes [M(CO)(5)PH(CF(3))(2)] (M = Cr, W) with [K(18-crown-6)][Ag(CN)(2)], the analogous trinuclear argentates, [Ag[(micro-P(CF(3))(2))M(CO)(5)](2)](-), are formed. The chromium compound [K(18-crown-6)][Ag[(micro-P(CF(3))(2))Cr(CO)(5)](2)] crystallizes in a noncentrosymmetric space group Fdd2 (No. 43), a = 2970.2(6) pm, b = 1584.5(3) pm, c = 1787.0(4), V = 8.410(3) nm(3), Z = 8. The C(2) symmetric anion, [Ag[(micro-P(CF(3))(2))Cr(CO)(5)](2)](-), shows a nearly linear arrangement of the P-Ag-P unit. Although the bis(pentafluorophenyl)phosphanido compound [Ag[P(C(6)F(5))(2)](2)](-) has not been obtained so far, the synthesis of its trinuclear counterpart, [K(18-crown-6)][Ag[(micro-P(C(6)F(5))(2))W(CO)(5)](2)], was successful.

16.
Inorg Chem ; 41(24): 6397-403, 2002 Dec 02.
Artigo em Inglês | MEDLINE | ID: mdl-12444783

RESUMO

The large discrepancies between the calculated and observed structures for BrF(4)(+) and IF(4)(+) (Christe, K. O.; Zhang, X.; Sheehy, J. A.; Bau, R. J. Am. Chem. Soc. 2001, 123, 6338) prompted a redetermination of the crystal structures of BrF(4)(+)Sb(2)F(11)(-) (monoclinic, P2(1)/c, a = 5.2289(6) A, b = 14.510(2) A, c = 14.194(2) A, beta = 90.280(1) degrees, Z = 4) and IF(4)(+)SbF(6)(-) (orthorhombic, Ibca, a = 8.2702(9) A, b = 8.3115(9) A, c = 20.607(2) A, Z = 8). It is shown that for BrF(4)(+), the large differences were mainly due to large errors in the original experimental data. For IF(4)(+)SbF(6)(-), the geometry previously reported for IF(4)(+) was reasonably close to that found in this study despite a very large R-factor of 0.15 and a refinement in an incorrect space group. The general agreement between the calculated and the redetermined geometries of BrF(4)(+) and IF(4)(+) is excellent, except for the preferential compression of one bond angle in each ion due to the influence of interionic fluorine bridges. In BrF(4)(+), the fluorine bridges are equatorial and compress this angle. In IF(4)(+), the nature of the fluorine bridges depends on the counterion, and either the axial (in IF(4)(+)SbF(6)(-)) or the equatorial (in IF(4)(+)Sb(2)F(11)(-)) bond angle is preferentially compressed. Therefore, the geometries of the free ions are best described by the theoretical calculations.

17.
Chemistry ; 9(12): 2839-44, 2003 Jun 16.
Artigo em Inglês | MEDLINE | ID: mdl-12866559

RESUMO

Metathetical processes were used to convert N5SbF6 into N5[B(CF3)4] and (N5)2SnF6. The latter salt is especially noteworthy because it contains two N5+ ions per anion, thus demonstrating that salts with touching polynitrogen cations can be prepared. This constitutes an important milestone towards our ultimate goal of synthesizing a stable, ionic nitrogen allotrope. The stepwise decomposition of (N5)2SnF6 yielded N5SnF5. Multinuclear NMR spectra show that in HF the SnF5- ion exists as a mixture of Sn2F(10)(2-) and Sn4F(20)(4-) ions. Attempts to isolate FN5 from the thermolysis of (N5)2SnF6 were unsuccessful, yielding only the expected decomposition products, FN3, N2, trans-N2F2, NF3, and N2.

18.
Chemistry ; 10(2): 508-17, 2004 Jan 23.
Artigo em Inglês | MEDLINE | ID: mdl-14735519

RESUMO

The highly explosive molecules As(N(3))(3) and Sb(N(3))(3) were obtained in pure form by the reactions of the corresponding fluorides with (CH(3))(3)SiN(3) in SO(2) and purification by sublimation. The crystal structures and (14)N NMR, infrared, and Raman spectra were determined, and the results compared to ab initio second-order perturbation theory calculations. Whereas Sb(N(3))(3) possesses a propeller-shaped, pyramidal structure with perfect C(3) symmetry, the As(N(3))(3) molecule is significantly distorted from C(3) symmetry due to crystal packing effects.

19.
Inorg Chem ; 43(10): 3189-99, 2004 May 17.
Artigo em Inglês | MEDLINE | ID: mdl-15132626

RESUMO

The new tin(IV) species (CH(3))(2)SnCl(OTeF(5)) was prepared via either the solvolysis of (CH(3))(3)SnCl in HOTeF(5) or the reaction of (CH(3))(3)SnCl with ClOTeF(5). It was characterized by NMR and vibrational spectroscopy, mass spectrometry, and single crystal X-ray diffraction. (CH(3))(2)SnCl(OTeF(5)) crystallizes in the monoclinic space group P2(1)/n (a = 5.8204(8) A, b =10.782(1) A, c =15.493(2) A, beta = 91.958(2) degrees, V = 971.7(2) A(3), Z = 4). NMR spectroscopy of (CH(3))(3)SnX, prepared from excess Sn(CH(3))(4) and HX (X = OTeF(5) or N(SO(2)CF(3))(2)), revealed a tetracoordinate tin environment using (CH(3))(3)SnX as a neat liquid or in dichloromethane-d(2) (CD(2)Cl(2)) solutions. In acetone-d(6) and acetonitrile-d(3) (CD(3)CN) solutions, the tin atom in (CH(3))(3)SnOTeF(5) was found to extend its coordination number to five by adding one solvent molecule. In the strong donor solvent DMSO, the Sn-OTeF(5) bond is broken and the (CH(3))(3)Sn(O=S(CH(3))(2))(2)(+) cation and the OTeF(5)(-) anion are formed. (CH(3))(3)SnOTeF(5) and (CH(3))(3)SnN(SO(2)CF(3))(2) react differently with water. While the Te-F bonds in the OTeF(5) group of (CH(3))(3)SnOTeF(5) undergo complete hydrolysis that results in the formation of [(CH(3))(3)Sn(H(2)O)(2)](2)SiF(6), (CH(3))(3)SnN(SO(2)CF(3))(2) forms the stable hydrate salt [(CH(3))(3)Sn(H(2)O)(2)][N(SO(2)CF(3))(2)]. This salt crystallizes in the monoclinic space group P2(1)/c (a = 7.3072(1) A, b =13.4649(2) A, c =16.821(2) A, beta = 98.705(1) degrees, V = 1636.00(3) A(3), Z = 4) and was also characterized by NMR and vibrational spectroscopy.

20.
J Am Chem Soc ; 126(3): 834-43, 2004 Jan 28.
Artigo em Inglês | MEDLINE | ID: mdl-14733558

RESUMO

Ab initio molecular orbital theory has been used to calculate accurate enthalpies of formation and adiabatic electron affinities or ionization potentials for N3, N3-, N5+, and N5- from total atomization energies. The calculated heats of formation of the gas-phase molecules/ions at 0 K are DeltaHf(N3(2Pi)) = 109.2, DeltaHf(N3-(1sigma+)) = 47.4, DeltaHf(N5-(1A1')) = 62.3, and DeltaHf(N5+(1A1)) = 353.3 kcal/mol with an estimated error bar of +/-1 kcal/mol. For comparison purposes, the error in the calculated bond energy for N2 is 0.72 kcal/mol. Born-Haber cycle calculations, using estimated lattice energies and the adiabatic ionization potentials of the anions and electron affinities of the cations, enable reliable stability predictions for the hypothetical N5(+)N3(-) and N5(+)N5(-) salts. The calculations show that neither salt can be stabilized and that both should decompose spontaneously into N3 radicals and N2. This conclusion was experimentally confirmed for the N5(+)N3(-) salt by low-temperature metathetical reactions between N5SbF6 and alkali metal azides in different solvents, resulting in violent reactions with spontaneous nitrogen evolution. It is emphasized that one needs to use adiabatic ionization potentials and electron affinities instead of vertical potentials and affinities for salt stability predictions when the formed radicals are not vibrationally stable. This is the case for the N5 radicals where the energy difference between vertical and adiabatic potentials amounts to about 100 kcal/mol per N5.

SELEÇÃO DE REFERÊNCIAS
DETALHE DA PESQUISA