Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 38
Filter
Add more filters










Publication year range
1.
Inorg Chem ; 45(14): 5351-7, 2006 Jul 10.
Article in English | MEDLINE | ID: mdl-16813398

ABSTRACT

The oxorhenium(V) dimer {MeReO(edt)}2 (1; where edt = 1,2-ethanedithiolate) catalyzes S atom transfer from thiiranes to triarylphosphines and triarylarsines. Despite the fact that phosphines are more nucleophilic than arsines, phosphines are less effective because they rapidly convert the dimer catalyst to the much less reactive catalyst [MeReO(edt)(PAr3)] (2). With AsAr3, which does not yield the monomer, the rate law is given by v = k[thiirane][1], independent of the arsine concentration. The values of k at 25.0 degrees C in CDCl3 are 5.58 +/- 0.08 L mol(-1) s(-1) for cyclohexene sulfide and ca. 2 L mol(-1) s(-1) for propylene sulfide. The activation parameters for cyclohexene sulfide are deltaH(double dagger) = 10.0 +/- 0.9 kcal mol(-1) and deltaS(double dagger) = -21 +/- 3 cal K(-1) mol(-1). Arsine enters the catalytic cycle after the rate-controlling release of alkene, undergoing a reaction with the Re(VII)(O)(S) intermediate that is so rapid in comparison that it cannot be studied directly. The use of a kinetic competition method provided relative rate constants and a Hammett reaction constant, rho = -1.0. Computations showed that there is little thermodynamic selectivity for arsine attack at O or S of the intermediate. There is, however, a large kinetic selectivity in favor of Ar3AsS formation: the calculated values of deltaH(double dagger) for attack of AsAr3 at Re=O vs Re=S in Re(VII)(O)(S) are 23.2 and 1.1 kcal mol(-1), respectively.


Subject(s)
Arsenicals/chemistry , Organophosphorus Compounds/chemistry , Rhenium/chemistry , Sulfides/chemistry , Catalysis , Kinetics , Magnetic Resonance Spectroscopy , Models, Molecular , Organometallic Compounds/chemistry , Thermodynamics
2.
Inorg Chem ; 45(3): 1053-8, 2006 Feb 06.
Article in English | MEDLINE | ID: mdl-16441113

ABSTRACT

Reinvestigation of the reaction between title reagents in aqueous acidic triflate and perchlorate media revealed an unusual difference: the reaction is strictly first-order with respect to the concentration of Fe(phen)3(2+) (phen = 1,10-phenanthroline) in the triflate medium but shows an additional, but we believe artifactual, higher-order term in the perchlorate medium. We postulate that the apparent orders with respect to [Fe(phen)3(2+)] in (H/Li)ClO4 do not indicate the actual chemical mechanism but, in whole or in part, the orders, particularly the higher-order component, reflect an interaction specific to Fe(phen)3(2+) or Fe(phen)3(3+) and ClO4(-) in solution. Data in (H/Li)O3SCF3 solutions indicate that, in the absence of added Fe(phen)3(3+), the first of the three sequential electron-transfer steps is rate controlling. Reactions started in the presence of the product Fe(phen)3(3+) occur somewhat more slowly, suggesting the first electron transfer is reversible. This finding allows the relative rate constants for Cr(V) oxidation and reduction to be evaluated, with limited precision, by two methods of analysis. The dependences on [Cr(VI)] can be resolved into contributions from the species HCrO4(-) and Cr2O7(2-), each of which in turn depends on [H+]. The reaction mechanism is discussed in light of the data obtained in the triflate medium. Further, the rate constants for certain steps can be considered in light of E0 for the Cr(VI)/(V) couple.

3.
Inorg Chem ; 44(23): 8426-32, 2005 Nov 14.
Article in English | MEDLINE | ID: mdl-16270981

ABSTRACT

Computational estimates have been made for the P=S and As=S bond strengths in triphenylphosphine sulfide and triphenylarsine sulfide, on the basis of G3 calculations for the methyl analogues and isodesmic-exchange reactions. Also, with the performance of the G3 method level for related compounds taken into consideration, the best estimates are 82 and 68 kcal/mol, respectively. While the value for triphenylarsine sulfide is within 2 kcal/mol of the single experimental estimate, that for triphenylphosphine sulfide is lower by 6 kcal/mol. (Capps, K. B.; Wixmerten, B.; Bauer, A.; Hoff, C. D. Inorg. Chem. 1998, 37, 2861-2864.) Despite virtually identical electronegativities of P and As, it is found that there is greater charge separation in the P=S bond. It is found that S atom transfer from thiiranes to arsines is exothermic.

4.
J Am Chem Soc ; 127(43): 15006-7, 2005 Nov 02.
Article in English | MEDLINE | ID: mdl-16248622

ABSTRACT

Kinetic evidence suggests the possibility of a dicationic intermediate in the title reaction. Thus the linkage isomerization reaction, PNC+ = PCN+, is described by the rate law, nu = 3/2k[PNC+]3/2, which can be interpreted by a chain mechanism with the propagation reaction PNC+ + P2+ --> P2+ + PCN+. Such propagation is unusual in that the intermediate regenerates itself in this single step rather than forming a different intermediate for a second propagation step. Cyanide ions inhibit the rate because they participate in the termination step, P2+ + CN- --> PCN+. The rate constant in CD3CN at 100 degrees C is 3/2k = 7.2 +/- 0.6 x 10-5 L1/2 mol-1/2 s-1; 3/2k represents the composite (kinit/kterm)1/2 kprop. When the reaction is carried out in the presence of PBr+, however, the reaction becomes much faster and is described by the rate law, nu = kBr[PBr+][PNC+]; because [PBr+] remains at constant concentration, the time-course experiments follow first-order kinetics.

5.
J Am Chem Soc ; 127(32): 11447-53, 2005 Aug 17.
Article in English | MEDLINE | ID: mdl-16089474

ABSTRACT

The rate constants of para-/orthohydrogen (p-/o-H2) nuclear spin isomerization have been measured by means of 1H NMR in deuterated solvents at 298.2 K. The indicated reaction is catalyzed by paramagnetic complex ions giving rate constants that are proportional to the concentrations of the catalysts. The second-order rate constants are directly proportional to the squares of the magnetic moments for the solvated metal complexes for two classifications: M(solv)m2+, M = 3d transition metals; Ln(solv)n3+, where in 1:9 D2O-CD3CN the aqua complexes are the predominant species, Ln = lanthanides. The other 3d transition metal complexes with different ligands show rate constants that also depend on the sizes of ligands. Whereas the correlation between the second-order rate constants and magnetic moments is consistent with Wigner's theory, the size of catalyst shows a more modest effect on the rate constants than expected. The effective collision radii of the complexes, calculated from the rate constants, proved to be approximately constant for each series of solvated metal complexes.

6.
J Am Chem Soc ; 127(29): 10436-47, 2005 Jul 27.
Article in English | MEDLINE | ID: mdl-16028958

ABSTRACT

A series of rhenium(V) complexes, [(X)(ReO)(dt)(PPh(3))] and [(o-SC(6)H(4)PPh(2))(ReO)(mtp)], were prepared to explore electronic effects on the C-S cleavage reaction that occurs upon reaction with PAr(3) at ambient temperature [where X = S(C(6)H(4)-p-Z) (Z = OMe, Me, H, F, Cl), OPh, Cl, and SC(2)H(5), and dt is the chelating dithiolate ligand derived from 2-(mercaptomethyl)thiophenol, 1,2-ethanedithiol, 1,3-propanedithiol, 1,3-butanedithiol, and 2,4-pentanedithiol]. The scope and selectivity of the C-S activation were examined. The C-S bond cleavage to form metallacyclic Re(V) complexes with a ReS core occurs only for the complexes with mtp and pdt frameworks and X = SAr and SC(2)H(5). The difference in reactivity is due to the different donating abilities of ancillary and dithiolate ligands, especially their pi-donating ability, which plays a critical role in C-S activation. The kinetics of the C-S activation process was determined; nucleophilic attack of PPh(3) on the oxo group of the Re(V)O core appears to be the rate-controlling step. The reaction is accelerated by electron-poor ArS ligands, but is unaffected by the substituents on phosphines. A detailed mechanistic study is presented. The results represent a rare example of migration of alkanethiolate leading to the formation of alkylthiolato complexes.

7.
Inorg Chem ; 44(15): 5514-22, 2005 Jul 25.
Article in English | MEDLINE | ID: mdl-16022550

ABSTRACT

The reaction between vanadium(III) and hydrogen peroxide in aqueous acidic solutions was investigated. The rate law shows first-order dependences on both vanadium(III) and hydrogen peroxide concentrations, with a rate constant, defined in terms of -d[H(2)O(2)]/dt, of 2.06 +/- 0.03 L mol(-)(1) s(-)(1) at 25 degrees C; the rate is independent of hydrogen ion concentration. The varying reaction stoichiometry, the appreciable evolution of dioxygen, the oxidation of 2-PrOH to acetone, and the inhibition of acetone formation by the hydroxyl radical scavengers, dimethyl sulfoxide and sodium benzoate, point to a Fenton mechanism as the predominant pathway in the reaction. Methyltrioxorhenium(VII) does not appear to catalyze this reaction. A second-order rate constant for the oxidation of V(3+) by OV(O(2))(+) was determined to be 11.3 +/- 0.3 L mol(-)(1) s(-)(1) at 25 degrees C. An overall reaction scheme consisting of over 20 reactions, in agreement with the experimental results and literature reports, was established by kinetic simulation studies.


Subject(s)
Hydrogen Peroxide/chemistry , Vanadium Compounds/chemistry , Vanadium/chemistry , Hydrogen-Ion Concentration , Kinetics , Molecular Structure , Oxidation-Reduction , Solutions/chemistry , Time Factors , Water/chemistry
8.
Inorg Chem ; 44(10): 3690-9, 2005 May 16.
Article in English | MEDLINE | ID: mdl-15877453

ABSTRACT

The new compound {(PhS)(2)ReO(mu-SPh)}(2), 1, was synthesized from Re(2)O(7) and PhSH and then used as the synthon for a number of hitherto unknown oxorhenium(V) compounds. Reactions between dithiols and 1 (2:1 ratio) afford {PhSReO(dt)}(2), where the dithiols, dtH(2), are 1,2-ethanedithiol (edtH(2)), 1,3-propanedithiol (pdtH(2)), 1,3-butanedithiol (pdtMeH(2)), 1,2-benzenedithiol (bdtH(2)), 2-(mercaptomethyl)thiophenol (mtpH(2)), and 2-mercaptoethyl sulfide (mesH(2)). Similar reactions carried out with a 3:1 ratio of dtH(2) to 1 afford [(ReO)(2)(dt)(3)], dt = edt, pdt. When NEt(3) was introduced prior to the 3:1 reaction between edtH(2) and 1, a compound containing an anionic complex was isolated, [PPh(4)][ReO(edt)(2)]. The new compounds were characterized analytically, spectroscopically, and crystallographically. The Re-O groups in two of the compounds, 1 and {ReO(mu-SPh)(bdt)}(2), exist in rare anti orientations; the others adopt the more familiar syn geometry, as discussed. Selected monomerization reactions of {PhSReO(dt)}(2) were also carried out: {PhSReO(dt)}(2) + 2L = 2[PhSReO(dt)L]. The rate for L = 4-phenylpyridine is given by v = {k(a)[L] + k(b)[L](2)} x [{PhSReO(dt)}(2)], as it is for the reactions of {MeReO(dt)}(2); for all of these compounds, the reaction proceeds nearly entirely by the third-order pathway. Values of k(b)/L(2) mol(-2) s(-1) at 25.0 degrees C are 5.8 x 10(2) (mtp), 2.97 x 10(3) (pdt), 4.62 x 10(5) (edt), and 3.87 x 10(5) (bdt). The rate law for the reactions of {PhSReO(dt)}(2) with L = PAr(3) is v = k(a)[L]/{1 + kappa[L]} x [{PhSReO(dt)}(2)]. For PPh(3), values at 25.0 degrees C of k(a)/L mol(-1) s(-1) (kappa/L mol(-1)) for {PhSReO(dt)}(2) are 9.64 x 10(-2) (1.87) for mtp, 3.43 x 10(-2) (0.492) for pdt, 1.91 (1.42) for edt, 1.84 x 10(-2) (0.82) for bdt, and 1.14 x 10(3) (10.6) for 1. Mechanisms are proposed that are consistent with the data obtained and with earlier work.

9.
Inorg Chem ; 44(7): 2465-71, 2005 Apr 04.
Article in English | MEDLINE | ID: mdl-15792485

ABSTRACT

Although neither vanadium(V) ions nor hydrogen peroxide efficiently oxidize the title substrates, they do so in combination, with vanadium(V) as the catalyst in acidic aqueous acetonitrile. The kinetic data show that, of the two peroxovanadium species present, OV(O2)+ and OV(O2)2-, only the latter reacts at a detectable rate. This unanticipated result can be attributed to the weaker O-O and V-O bonds in the diperoxo complex. The rate constants for both series of substrates follow the Hammett correlation, with rhoP = -1.35 and rhoS = -0.83. To analyze properly the kinetic data for the Ar3P compounds, account must be taken of the protonation to Ar3PH+ in acidic solution. In retrospect, our earlier study [Abu-Omar, M. M.; Espenson, J. H. J. Am. Chem. Soc. 1995, 117, 272-280] of phosphine oxidation catalyzed by MeReO3 failed to do so, and the reaction constant must be corrected from the originally reported value to -1.56.

10.
Inorg Chem ; 44(7): 2560-5, 2005 Apr 04.
Article in English | MEDLINE | ID: mdl-15792495

ABSTRACT

The synthesis and characterization of a new oxorhenium(V) compound is reported; it is [MeReO(edt)(bpym)], 8, where edt = 1,2-ethanedithiolate and bpym = 2,2'-bipyrimidine. Compound 8 was characterized by NMR spectroscopy and single-crystal X-ray analysis. It exists as a six-coordinate Re(V) compound comparable to the previously known [MeReO(edt)(bpy)] and [MeReO(mtp)(bpy)]. Compound 8 catalyzes the oxygen-atom-transfer reaction PicO + PZ3 --> Pic + Z3PO, whereas the other two do not. The kinetics of this reaction with catalyst 8 follows the rate law -d[PicO]/dt = k[8][PicO]/(1 + c[PZ3]). With different phosphines, the rate law has the same k value, 4.17 L mol(-1) s(-1), but different c values. For tritolylphosphine, c = 67.5 L mol(-1) in benzene at 25 degrees C. A mechanism has been proposed to account for these findings. The data establish that an open coordination site on rhenium is necessary for oxygen-atom-transfer reactions.

11.
Inorg Chem ; 44(3): 489-95, 2005 Feb 07.
Article in English | MEDLINE | ID: mdl-15679376

ABSTRACT

The title reactions occur stepwise, the first and fastest being MeReO3 + Eu2+ --> Re(VI) + Eu3+ (k298 = 2.7 x 10(4) L mol(-1) s(-1)), followed by rapid reduction of Re(VI) by Eu2+ to MeReO2. The latter species is reduced by a third Eu2+ to Re(IV), a metastable species characterized by an intense charge transfer band, epsilon410 = 910 L mol(-1) cm(-1) at pH 1; the rate constant for its formation is 61.3 L mol(-1) s(-1), independent of [H+]. Yet another reduction step occurs, during which hydrogen is evolved at a rate v = k[Re(IV)][Eu2+][H+](-1), with k = 2.56 s(-1) at mu = 0.33 mol L(-1). The 410 nm Re(IV) species bears no ionic charge on the basis of the kinetic salt effect. We attribute hydrogen evolution to a reaction between H-ReVO and H3O+, where the hydrido complex arises from the unimolecular rearrangement of Re(III)-OH in a reaction that cannot be detected directly. Chromium(II) ions do not evolve H2, despite E(Cr) degrees approximately E(EU) degrees. We attribute this lack of reactivity to the Re(IV) intermediate being captured as [Re(IV)-O-Cr(III)]2+, with both metals having substitutionally inert d3 electronic configurations. Hydrogen evolution occurs in chloride or triflate media; with perchlorate present, MeReO2 reduces perchlorate to chloride, as reported previously [Abu-Omar, M. M.; Espenson, J. H. Inorg. Chem. 1995, 34, 6239-6240].

12.
J Org Chem ; 70(1): 238-43, 2005 Jan 07.
Article in English | MEDLINE | ID: mdl-15624928

ABSTRACT

Kinetic data have been obtained for three distinct types of reactions of phthalimide N-oxyl radicals (PINO(.)) and N-hydroxyphthalimide (NHPI) derivatives. The first is the self-decomposition of PINO(.) which was found to follow second-order kinetics. In the self-decomposition of 4-methyl-N-hydroxyphthalimide (4-Me-NHPI), H-atom abstraction competes with self-decomposition in the presence of excess 4-Me-NHPI. The second set of reactions studied is hydrogen atom transfer from NHPI to PINO(.), e.g., PINO(.) + 4-Me-NHPI <=> NHPI + 4-Me-PINO(.). The substantial KIE, k(H)/k(D) = 11 for both forward and reverse reactions, supports the assignment of H-atom transfer rather than stepwise electron-proton transfer. These data were correlated with the Marcus cross relation for hydrogen-atom transfer, and good agreement between the experimental and the calculated rate constants was obtained. The third reaction studied is hydrogen abstraction by PINO(.) from p-xylene and toluene. The reaction becomes regularly slower as the ring substituent on PINO(.) is more electron donating. Analysis by the Hammett equation gave rho = 1.1 and 1.8 for the reactions of PINO(.) with p-xylene and toluene, respectively.

14.
Inorg Chem ; 43(13): 3854-62, 2004 Jun 28.
Article in English | MEDLINE | ID: mdl-15206866

ABSTRACT

Compounds that contain the anion [MeReO(edt)(SPh)](-) (3-) were synthesized with the countercations 2-picolinium (PicH+3-) and 2,6-lutidinium (LutH+3-), where edt is 1,2-ethanedithiolate. Both PicH+3- and MeReO(edt)(tetramethylthiourea) (4) were crystallographically characterized. The rhenium atom in each of these compounds exists in a five-coordinate distorted square pyramid. In the solid state, PicH+3- contains an anion with a short (d(SH) = 232 pm) and nearly linear hydrogen-bonded (N-H.S) interaction to the cation. Ligand substitution reactions were studied in chloroform. Displacement of PhSH by PPh(3) follows second-order kinetics, d[MeReO(edt)(PPh(3))]/dt = k[PicH+3-][PPh3], whereas with pyridines an unusual form was found, d[MeReO(edt)(Py)]/dt = k[PyH+3-][Py](2), in which the conversion of PicH+3- to PyH+3- has been incorporated. Further, added Py accelerates the formation of [MeReO(edt)(PPh3)], v = k.[PicH+3-].[PPh3].[Py]. Compound 4, on the other hand, reacts with both PPh(3) and pyridines, L, at a rate given by d[MeReO(edt)(L)]/dt = k.[4].[L]. When PicH+3- reacts with pyridine N-oxides, a three-stage reaction was observed, consistent with ligand replacement of SPh(-) by PyO, N-O bond cleavage of the PyO assisted by another PyO, and eventual decomposition of MeRe(O)(edt)(OPy) to MeReO(3). Each of first two steps showed a large substituent effect; Hammett analysis gave rho(1) = -5.3 and rho(2) = -4.3.

15.
Inorg Chem ; 42(25): 8122-4, 2003 Dec 15.
Article in English | MEDLINE | ID: mdl-14658859

ABSTRACT

The compound MeRe(S)(mtp)(PPh3), 2, where mtpH2 is 2-(mercaptomethyl)thiophenol, was used to catalyze the reaction between pyridine N-oxides, PyO, and triphenylphosphine. The rate law is -d[PyO]/dt=kc'[2].[PyO](1/2), with kc' at 25.0 degrees C in benzene=0.68 (4-picoline N-oxide) and 3.5x10(-3) dm(3/2) mol(-1/2) s(-1) (4-NO2-pyridine N-oxide). A chain mechanism with three steady-state thiorhenium species as chain carriers is implicated.

16.
J Org Chem ; 68(24): 9364-70, 2003 Nov 28.
Article in English | MEDLINE | ID: mdl-14629158

ABSTRACT

The phthalimide N-oxyl (PINO) radical was generated by the oxidation of N-hydroxyphthalimide (NHPI) with Pb(OAc)4 in acetic acid. The molar absorptivity of PINO* is 1.36 x 10(3) L mol(-1) cm(-1) at lambda(max) 382 nm. The PINO radical decomposes slowly with a second-order rate constant of 0.6 +/- 0.1 L mol(-1) s(-1) at 25 degrees C. The reactions of PINO(*) with substituted toluenes, benzaldehydes, and benzyl alcohols were investigated under an argon atmosphere. The second-order rate constants were correlated by means of a Hammett analysis. The reactions with toluenes and benzyl alcohols have better correlations with sigma+ (rho = -1.3 and -0.41), and the reaction with benzaldehydes correlates better with sigma (rho = -0.91). The kinetic isotope effect was also studied and significantly large values of k(H)/k(D) were obtained: 25.0 (p-xylene), 27.1 (toluene), 27.5 (benzaldehyde), and 16.9 (benzyl alcohol) at 25 degrees C. From the Arrhenius plot for the reactions with p-xylene and p-xylene-d(10), the difference of the activation energies, E(a)(D) - E(a)(H), was 12.6 +/- 0.8 kJ mol(-1) and the ratio of preexponential factors, A(H)/A(D), was 0.17 +/- 0.05. These findings indicate that quantum mechanical tunneling plays an important role in these reactions.

17.
Inorg Chem ; 42(18): 5735-42, 2003 Sep 08.
Article in English | MEDLINE | ID: mdl-12950224

ABSTRACT

An oxorhenium(V) dimer, [PMeReO(mtp)](2), D, where mtpH(2) is 2-(mercaptomethyl)thiophenol, catalyzes oxygen atom transfer reaction from methyl phenyl sulfoxide to triarylphosphines. Kinetic studies in benzene-d(6) at 23 degrees C indicate that the reaction takes place through the formation of an adduct between D and sulfoxide. The equilibrium constants, K(DL), for adduct formation were determined by spectrophotometric titration, and the values of K(DL) for MeS(O)C(6)H(4)-4-R were obtained as 14.1(2), 5.7(1), and 2.1(1) for R = Me, H, and Br, respectively. Following sulfoxide binding, oxygen atom transfer occurs with either internal or external nucleophilic assistance. Because [MeReO(mtp)](2) is a much more reactive catalyst than its monomerized form, MeReO(mtp)PPh(3), loss of the active catalyst during the time course of the reaction must be taken into account as a part of the kinetic analysis. As it happens, sulfoxide catalyzes monomerization. Monomerization by triarylphosphines was also studied in the presence of sulfoxide, and a mechanism for that reaction was also proposed. Both the phosphine-assisted monomerization and the phosphine-assisted pathway for oxygen atom transfer involve transition states with ternary components, D, sulfoxide, and phosphine, which we suggest are structural isomers of one another.

19.
Chem Commun (Camb) ; (10): 1162-3, 2003 May 21.
Article in English | MEDLINE | ID: mdl-12778715

ABSTRACT

Exposure to visible light increases the rate of oxidation of chlorinated phenols by hydrogen peroxide in aqueous solution in either the presence or the absence of iron-based catalysts, which may be explained by the aqueous photoreactions of chloroquinone intermediates.

20.
Inorg Chem ; 42(7): 2362-7, 2003 Apr 07.
Article in English | MEDLINE | ID: mdl-12665371

ABSTRACT

Four new methyloxorhenium(V) compounds were synthesized with these tridentate chelating ligands: 2-mercaptoethyl sulfide (abbreviated HSSSH), 2-mercaptoethyl ether (HSOSH), thioldiglycolic acid (HOSOH), and 2-(salicylideneamino)benzoic acid (HONOH). Their reactions with MeReO(3) under suitable conditions led to these products: MeReO(SSS), 1, MeReO(SOS), 2, MeReO(OSO)(PAr(3)), 3, and MeReO(ONO)(PPh(3)), 4. These compounds were characterized spectroscopically and crystallographically. Compounds 1 and 2 have a five-coordinate distorted square pyramidal geometry about rhenium, whereas 3 and 4 are six-coordinate compounds with distorted octahedral structures. The kinetics of oxidation of 2 and 3 in chloroform with pyridine N-oxides follow different patterns. The oxidation of 2 shows first-order dependences on the concentrations of 2 and the ring-substituted pyridine N-oxide. The Hammett analysis of the rate constants gives a remarkably large and negative reaction constant, rho = -4.6. The rate of oxidation of 3 does not depend on the concentration or the identity of the pyridine N-oxide, but it is directly proportional to the concentration of water, both an accidental and then a deliberate cosolvent. The mechanistic differences have been interpreted as reflecting the different steric demands of five- and six-coordinate rhenium compounds.

SELECTION OF CITATIONS
SEARCH DETAIL
...