A PART I SPECTROSCOPIC STUDIES OF FOUR-AND FIVE-COORDINATE TRANSITION METAL COMPLEXES WITH TERTIARY PHOSPHORUS LIGANDS PART II REACTIONS OF SOME CYANOPHOSPHINES WITH DIBORANE AND BORON TRIFLUORIDE _ Thesis for the‘Degree of Ph. D. MICHIGAN STATE UNIVERSITY EDWARD JOSEPH LUKOSIUS 1 9 7 2 4-4.; ‘AL . a LIBRHRY Michigan State University '7' This is to certify that the thesis entitled Part I. Spectroscopic Studies of Four- and Five-Coordinate Transition Metal Complexes with Tertiary Phosphorus Ligands Part II. Reactions of Some Cyanophosphines with Diborane and Boron Trifluoride presented by Edward Joseph Lukosius has been accepted towards fulfillment of the requirements for Ph . D . degree in Chem; 5&1! Major professor Date July 28. 1972 0-7639 .uéhnay NOAH & SUNS’ 800K BINDERY INC. LIBRARY amozns :.. :I’:::::-‘::T. Inn-m ABSTRACT PART I SPECTROSCOPIC STUDIES OF FOUR-AND FIVE-COORDINATE TRANSITION METAL COMPLEXES WITH TERTIARY PHOSPHORUS LIGANDS BY Edward Joseph Lukosius Infrared, electronic, and proton nuclear magnetic resonance data are presented and discussed for the following complexes: NiL3(CN)3, NiL2(CN)3 (L = (CH3)3P, (CH3)2POCH3. cnap(ocna),, P(OCH3)3), [NiL5](C104)3 (L = (CH3)21>ocn3. CHSP(OCH3)2, P(OCH3)3). [C0L51C104 and [RhL5]B(C3H5)4 (L = P(0CH3)3), and {Ni[(CH3)3P]4](BF4)2. The five-coordinate complexes are believed to have a trigonal bipyramidal struc- ture, both in the solid state and in solution, on the basis of their infrared and electronic spectra. Equilibria of the following type are found in solutions of the five-co— ordinate complexes: NiL3(CN)2 _3_ NiL3(CN)2 + L < + [ML51n > [ML4]n+ + L < Isosbestic points have been observed in the ultraviolet- visible Spectra of these solutions. Electronic factors which favor the formation of five-coordinate complexes over Edward Joseph Lukosius four—coordinate complexes have been discussed on the basis of the electronic spectra. The temperature dependence of the proton magnetic resonance spectra of the above equilibria have been investi- gated, and are discussed in terms of ligand exchange. The proton nmr spectra are similar for all complexes; a doublet at fast exchange rates, sharp singlet at intermediate rates, and multiplet at slow exchange rates. The doublets at fast ligand exchange rates are simply 31P-1H coupling, however, the coupling constants and chemical shifts are an average of complexed and free ligand in solution. Decoupling of phos— phorus and hydrogen nuclei at intermediate exchange rates where singlets are Obtained is believed due to a spin-ex- change mechanism for which strong 31P-31P coupling is necessary. Multiplets at slow exchange rates are due to 31P-31P and 31P-1H coupling through the metal. JPCH is found to be of opposite sign in complexed and free ligand, while JPOCH has the same sign. Qualitative rates of ligand exchange have been determined and follow the order: (CH3)3P > P(OCH3)3 > CHaP(OCH3)3 > (CH3)2POCH3. In the metal complexes of the typeML5n+ where L = P(OCH3)3, the rate of ligand exchange follows the order: Co+ > Rh+ > Ni2+. Edward Joseph Lukosius PART II REACTIONS OF SOME CYANOPHOSPHINES WITH DIBORANE AND BORON TRIFLUORIDE The following adducts have been prepared by the reaction of RgPCN (R - CH3. cnao. C3H5O. cana, N(CH3)2) with diborane and boron trifluoride; [R2PCN]BH3 (R = all of the groups listed above), [R3PCN13F3 (R = CH3, CH3O, canal and [RaPCNlBF3(Bfia) (R = CH3. CH30). Nuclear magnetic reso- nance (118, 31P, 19F, 1H) and infrared data (v are pre- CN) sented and discussed. 0n the basis of these data, it is concluded that the borane molecule is coordinated to the phosphorus atom and the BF3 molecule is bound to the nitro- gen atom of the cyanide group in all of the above adducts. The borane hyperconjugation model of the B-P bond has been shown to be inconsistent with the nmr data. However, a correlation between JPB and phosphine basicity has been observed in the borane adducts. Infrared data have suggested that the CN bond in the free cyanophosphines is inter- mediate between a triple and double bond. Experimental and spectroscopic results have indicated that the cyanide group decreases the basicity of the phosphorus atom to which it is bound in these RgPCN ligands,as compared to the corre- sponding PR3 ligands. PART I SPECTROSCOPIC STUDIES OF FOUR-AND FIVEoCOORDINATE TRANSITION METAL COMPLEXES WITH TERTIARY PHOSPHORUS LIGANDS PART II REACTIONS OF SOME CYANOPHOSPHINES WITH DIBORANE AND BORON TRIFLUORIDE BY Edward Joseph Lukosius A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1972 To Susan ACKNOWLEDGMENT It is with sincere appreciation that I acknowledge the encouragement and direction of Dr. Kenneth J. Coskran during this investigation. I am deeply grateful to my wife, Susan, for her under- standing and inspiration. I wish, also, to thank our parents, Mr. and-Mrs. Ignatius F. Lukosius of Boston, Massachusetts and.Mr. and Mrs. John W. Cotter of Dedham, Massachusetts for their assistance, advice and encouragement. iii TABLE OF CONTENTS INTRODUCTION . . . . . EXPERIMENTAL . . . . . RESULTS AND DISCUSSION Infrared Spectra Absorption Spectra Proton Nmr Spectra INTRODUCTION . . . . . EXPERIMENTAL . . . . . Materials . . . . Preparation Preparation Reaction of Nmr Spectra Infrared Spectra of [R RP(CN3 RESULTS . O O O O O 0 DISCUSSION . . . . . . [RaPCN]BH3 Adducts mammal?3 and [RgPCN1BF3(BH3) Adducts CONCLUSION . . . . . . BIBLIOGRAPHY . . . . . PART I 2 With BF3 iv of [R3PCNlBH3 Adducts Preparation of [R2PCN1BF3 Adducts PCN]BF:(BH3) Adducts Page 11 19 19 41 102 107 109 Table II. III. IV. V. VIII. IX. XI- XII- XIII. LIST OF TABLES PART I Five-coordinate Ni(II) complexes with mono- dentate ligands . . . . . . . . . . . . . . . Infrared data, cyanide stretching frequencies Absorption spectral data for Ni(CN)3 complexes K for NiL3(CN)2 complexes at 25°. . . . . inst Absorption spectral data for Ni(ClO4)2 and Ni(BF4) complexes . . . . . . . . . . . . . . Temperatures of singlet maxima and magnetic susceptibility data . . . . . . . . . . . . . PART II Experimental data for 3H3 and BF3 adduct formation . . . . . . . . . . . . . . . . . . Nmr and ir data for [RaPCN]BH3 adducts . . . Nmr and ir data for [RzPCNlBFa and [RgPCN]BF3(BH3) adducts . . . . . . . . . . . Values of JPCH' JPOCH' and JPNCH in R2 PCN and [R2PCN18H3 O o o c o o o o o o o o PhOSphorus chemical shift data in R2 PCN and [R2 PCN] BHS o o o o o o o o o o o o 61H(R) in RaPCN and [RzPCN]BH3 . . . . . . . Change in 631P in RzPCN and [RaPCN]BH8 upon BF3 adduct formation . . . . . . . . . . . . Page 20 33 34 64 80 86 87 95 99 101 105 LIST OF FIGURES PART I Figure Page 1. (I) Trigonal bipyramidal coordination Sphere surrounding Ni(II) in [Ni(rsp)c1]c1o4 (TSP = tris(gamethylthiophenyl)phosphine). (II) Tetragonal pyramidal arran ement of arsenic atoms around Ni(II) in Ni diars)(triars)](ClO4)2 (diars = g-phenylenebis dimethylarsine): triars ='bis(g-dimethylarsinophenyl)methyl arsine) . 2 2. Effect of distortion on d orbital energy levels in Dsh symmetry . . . . . . . . . . . 5 3. Splitting of d orbital energy levels in d8 complexes of C4v symmetry . . . . . . . . . 6 4. Pr0posed dimeric structure of NiL2(CN)3 (L I P(OCH8)3)018 o o o o o o o o o o o o o o 21 5. Ultraviolet-visible spectrum of: (a) Ni[P(0CH3)3]3(CN), in cnzclz: (b). solution in (a) with excess ligand . . . . . . . . . . 25 6. Absorption spectrum of (a) Ni[(CH3)3P]3(CN) in CH3c1a; (b) NiI$C33;3P]2ECN;2 in CH3C12: (c solid state Ni[ CH3 3P]3 CN 7. Effect of temperature change on the ultraviolet- visible spectrum of Ni[P(OCH3)3]3(CN)2 in CHaCJ-z o o o o o o o o o o o o o o o o o O o o 31 2 o o o o o o o o 27 8. Isosbestic points obtained for {Ni[P(OCH3)3]5}- (Cl-0‘ )2 in CHgCla o o o o o o o o o o o o o o 37 9. 1H nmr spectrum of Ni[(cnagp]2(cn)2 in CH3C12 at various temperatures . . .-. . . . . . . . 44 10. Portion of ultraviolet-visible spectrum of Ni[(m3)3P]3(CN)3 in CH3C12 o o o o o a o o o 47 11. 1H nmr spectrum of Ni[(CH3)3P]3(CN)2 in Cflgcla at various temperatures . . . . . . . . 52 vi LIST or FIGURES (Cont.) Figure Page 12. 1H nmr Spectrum of NiL3(CN)3 (L - (CH3)3POC33. CH3P(OCH3)3) in CH3C13 at various temperatures 54 13. 1H nmr spectrum of:(a) Ni[P(OCH3)3]3(CN) in CH C13: (b) solution in (a) with added P OCH3)3: (c solution in (b) when cooled . . . .~. . . 57 14. 1H nmr Spectrum of Ni[(CH3) P]3(CN)2 upon successive addition of (C3333P.. . . . . . . . 60 15. 1H nmr spectrum of Ni[CH3P(0CH3)2yJ(CN)3 in CH3613 upon addition of CH3P(OCH3 2 . . . . . 63 16. 1H nmr spectrum of {Rh[P(0CH3)3]5}B(C5H5)4 in C33C13 at various temperatures . ... . . . . . 69 PART II 17. Structure of [M(CO)4L]2 com lexes (M ' Cr or Mo: R - CH3; 0C3H5. C5115. N(CH3§2)75 9 o o o o o o 75 18. Nmr spectra of [(CH3)3PCN]BH : (a) Time-averaged 13 spectrum at 60.0 MHz. (b3 1H spectrum at 60.0 MHz. (c) 113 spectrum at 60.0 MHz. (d) 31F spectrum at 24.3 MHz . . . . . . . . . . . . . 85 19. 113 nmr spectrum of [(cnao) PCN]BF3(BH3) at 32.1 MHz: (a) Spectrum at -25 ; (b) Spectrum at -50° 90 20. The ligand cone angle for tertiary phosphorus compounds35. . . . . . . . . . . . . . . . . . 97 21. Possible structures of [RCNlBF351 . . . . . . 102 vii INTRODUCTION Four-coordinate complexes of Ni(II) have been known for some time.1 In recent years, a large number of five- coordinate Ni(II) complexes with polydentate ligands have been prepared}!3 The ability of these complexes to attain pentacoordination, as well as their resulting structures, are determined primarily by the restrictive geometry of these polydentate ligands."6 This is exemplified in Figure 1, where the structures of two such complexes (1 and II), as determined by X-ray studies,"v8 are pictured. The restrictive ligand geometry in Ni[(TSP)Cl]ClO4(I) forces the phosphorus and sulfur atoms into four positions of a trigonal bipyramid around Ni(II), allowing the fifth un- hindered position to be filled by a chloride ion.9 In [Ni(diars)(triars)](C104)z(II), the triarsine ligand geometry forces an arsenic atom into the fifth coordination site of a tetragonal pyramid.10 For monodentate ligands this type of ligand restrictive geometry is not possible, and indeed, few five—coordinate Ni(II) complexes with monodentate ligands have been reported. P IA: N 5 (4h! / As [-— s/ I A5/ '\A3/ X I' 'II Figure 1. (I) Trigonal bipyramidal coordination Sphere surrounding Ni(II) in [Ni(TSP)Cl]ClO4 (TSP = tris (g-methylthiophenyl)phosphine). (II) Tetragonal pyramidal arrangement of arsenic atoms around Ni(II) in [Ni(diars)(triars)](C104)2 (diars - g-phenylenebis(dimethylarsine); triars = bis(g—dimethylarsinophenyl)methyl arsine). This latter group consists of two main types; NiL3X2 (L = tertiary phosphorus ligand; X = CN, Cl, Br, I), and [NiL51X2 (L - tertiary phosphite; x - c104, BF4). In 1960 the first five—coordinate Ni(II) complex with monodentate ligands was prepared,11 Ni[P(C2H5)2C5H5]3(CEOCGH5)2 , which dissociated reversibly in solution to free ligand and the corresponding four-coordinate complex (Equation 1). Five— NiIP(C2H5)2CSHSIS(CECCSHS)2 < 1 P(C2H5)2C6H5 + NiIPICzfls)2CsHalz(CECC6H5)2 ( ) coordinate complexes that have been isolated since that time are listed in Table I. The stereochemistry adopted by these complexes is not determined by ligand steric restrictions, as in complexes with polydentate ligands. However, with regard to ligand- ligand repulsion, it has been shown that the trigonal Table I. Five-coordinate Ni(II) complexes with mono- dentate ligands. Compound Reference NiL3x3 L x HP(C5H5 )2 Clo Br. I [12] PC14H13 Cl, Br, I [13] RPC33H3 (R = CH3. C3H3) C1, Br. I [14] c6H5P(CH3)2 I. CN [15] (CH3)3P Br. I [16] P C4H9) CN [17] P 0CH3 CN [18] P 0C3H3§3 CN [19] P 0C3H 3 CN [18] P OCH )3CCH3 CN [18] C3H3P 0CH3) CN [19] C3H3P 0C3H )3 CN [19] C3H3P C3H ): CN [20] C5F5P CH3 2 CN [15] (C6H5)2PCH8 CN [15] [NiL31x3 L x P OCHs 3 C104 [21] P OCH, 3CCH3 C104 [21] P 0CH3 3CCH3CH3 C104 [21] C(CH3) 4CH3 C10, [21] POCH) 3(CH3) )3 PF, [21] Ni(CN)33 [221,[23] Nisr3[(CH3)3P]2 [24] 4 bipyramid is the most stable structure, being favored over the tetragonal pyramid.25 Indeed, X—ray structure studies of the complexes Ni[C5H5P(OC2H5)2]3(CN)3,25 Ni[C3H3P(CH3)3]3(CN)3,27 Ni(HP(C3H3)3]3I3.28 and {Ni[P(OCH)3(CH2)3]5)(Clo‘)229 have demonstrated a near trigonal bipyramidal arrangement of ligands about the Ni(II) atom, with Erggg_apical cyanides in the appropriate com- plexes. Only one square pyramidal arrangement of monoden- tate ligands about the Ni(II) atom has been determined by X-ray structure studies. In [Cr(NH3CH2CH2NH2)3]Ni(CN)5' 1.5H30,3° the Ni(CN)5-3 ions exist in both a square pyram- idal. and distorted trigonal bipyramidal structure. Electronic Spectra of these five-coordinate complexes support the assignment of a trigonal bipyramidal structure in solid and solution. For a low spin, d8, trigonal bi- pyramidal complex of D3h symmetry, two electronic transi— tions are expected as the result of metal d orbital Split- ting.31 v.[(e">4(e')4 -> 44 -> (e")3(e')4(a1)1] (Figure 2). Transition v1 is dipole allowed while v2 is forbidden.32 The electronic spectra of a number of the complexes in Table I exhibit these transi- tions; an intense band at low energies (v1) and a weak band at higher energies (v3).14v1903° However, in a number of these complexes the transition v1 has been observed to split into two distinct bands, in both the solid and solu- tion spectra.15'18'19 In the solid state this splitting has been attributed to a ground state distortion of the 5 ligand field levels of these complexes because of lattice packing interactions and the effects of ligand conformations or because of a second order JahndTeller effect (see Figure 2).33 Thus the symmetry of the complex is reduced and v1 will appear as two bands. The splitting of v1 is expected to be larger than that of v3 since the dx3_y3 and dxy orbitals are more sensitive to distortion than the dxz and dyz orbitals.33 a] d.2" .. x), 93 1 ”I II 42 . .d.2.,2- -° . '°dxy-- . O” a , I '. dxz. . O . 03.1 (or C3V) C2v(or Cs) Figure 2. Effect of distortion on d orbital energy levels in D symmetry. ah The solution electronic spectra of these complexes exhibit the same splitting of VI; however, at low tempera- tures, the two absorptions which constitute v1 are Observed to coalesce and at ~v77°K, a single, very symmetrical band is observed. Venanzi and coworkers 33 have proposed a dynamic JahndTeller distortion of excited states to explain 6 the splitting of VI at high temperatures. These authors have also preposed that a ground state distortion, in which different degrees of distortion could exist at different temperatures, would account for the splitting of VI. A tetragonal pyramidal structure does not account for the Observed spectra, Since three bands should be observed at all temperatures34 (Figure 3). Thus, five-coordinate Ni(II) complexes with monodentate ligands are believed to have a trigonal bipyramidal structure in solid and solution as a result of studies of their absorption Spectra. bl 7} de2.y2 0] — dzz b2 *‘* dxy e L‘fi gdxztdyz Gav Figure 3. Splitting of d orbital energy levels in d8 complexes of C4v symmetry. Infrared studies support this assignment of a trigonal bipyramidal structure. One CN stretching frequency appears in both the solid and solution (excess phOSphine is added 7 to prevent dissociation) spectra for NiL3(CN)3 complexes, indicating a Eggggfsubstituted complex.14:19 Taking into account x-ray and electronic studies mentioned above, this is most easily assigned to cyanides in apical positions of a trigonal bipyramid. Few studies have appeared on the nuclear magnetic resonance spectra of these complexes. Alyea andMeek15 have reported evidence for the exchange of phosphines on Ni[C3H5P(CH3)3]3(CN)3 in solution. Upon addition of free ligand, the methyl doublet resonance shifts upfield toward that of free ligand, indicating a time-averaging of added ligand and complexed ligand. In addition, Coskran, et al.13, reported the same result upon addition of P(OCH3)3 to a solution which contained Ni[P(OCH3)3]3(CN)2. Most of the five—coordinate NiL3x3 complexes that have been isolated have been reported to be reversibly unstable toward dissociation in solution, according to the following equilibriumu"20 (Equation 2): NiL3x3 > NiL3x3 + L (2) < (Ni(CN)5-3 also dissociates to Ni(CN)4-2).22o23 Electronic spectra of these complexes exhibit isosbestic points, which indicate the presence of an equilibrium in solution.12v15v1°v 18'30'23 Indeed, a number of the four-coordinate NiL2X2 complexes have been isolated.”"20 Thermodynamic studies have been carried out and equilibrium constants reported for some complexes involved in equilibrium 2.15:35 No 8 studies on the stabilities of the [NiL5]X2 complexes have been reported. The relative stability of five-coordinate complexes toward their corresponding four-coordinate complexes has been related to both electronic and steric effects.15.19 Gray and coworkers 19 have Shown that good o-donor ligands, as well as good n-acceptor ligands, should favor five-co- ordination for Ni(II) complexes with phosphine ligands. These authors have shown that the relative stability of five— coordinate complexes of C5H5P(0R)2 and P(OR)3 with nickel cyanide is determined by a balance of o and w electronic' effects. They have also suggested that P(C3H5)3 and (CGH5)2POR do not form five-coordinate complexes with nickel cyanide, but rather only four—coordinate complexes, because the bulkiness of these ligands precludes the attainment of pentacoordination about the Ni(II) atom. Simple oebonding considerations for these good o-donors would predict the formation of five—coordinate complexes.19 Alyea and Meek15 report results that suggest the inter-ligand interaction of (CBH5)3PC3H5 may prevent the formation of a five-coordinate Complex with nickel cyanide. These authors also found the tendency to stabilize a five-coordinate complex is less for a fluorophenylphosphine than for the corresponding, less bulky, unfluorinated phosphine. However, these latter re- sults could be due to electronic effects, since the fluoro- phenylphosphine would not be expected to be as good a o-donor as the corresponding unfluorinated phosphine. Coskran, 9 gt 213,18 have suggested that steric effects have been overo emphasized since they have isolated Ni[P(OC6H5)3]3(CN)2 which is stable with respect to dissociation. With these results, along with the already vast number of established four—coordinate Ni(II) complexes} it became of interest to consider more closely those effects, steric and electronic, which cause one coordination number over another. In this study an attempt was made to minimize the steric factors by utilizing "small"36 monodentate ligands with the intentions of relieving the ligand-ligand repul- sions and the restrictive geometry of polydentate ligands which may sharply influence the coordination number of the Ni(II) ion. It has been possible to relate the observed coordination number and geometry of the complex to electronic effects, i;gfi,the nature of the metal-ligand bond. To ac- complish this the series of ligands, (CH3)3P, (CH3)2POCH3. CH3P(0CH3)2, and P(OCH3)3 was chosen for study with Ni(ClO4)2'6H30, Ni(BF4)2°6H20, and Ni(CN)3. (Those com- plexes of P(OCH3)3 pertinent to this study have been pre- pared previously,13v21 and the complexes of (CH3)3P were reported while this research was in progress.37) The following five-coordinate complexes have been iso- lated and investigated in this study: NiL3(CN)2 (L = all of the ligands in the above series), and [NiL5](ClO4)2 (L = P(OCH3)21, CH3P(OCH3)2, (CH3)2POCH3). A number of four- coordinate complexes have also been isolated: gingiL2(CN)2, (L = (CH3)3PoCH3. CH3P(OCH3)2). trans-Ni[(CH3)3P]3(CN)3.37 and Ni[(CH3)3P]4(BF4)2 .37 10 Relative stabilities of the five—coordinate complexes toward their corresponding four—coordinate complexes were investigated by absorption spectral studies. The following types of equilibria were found to be present in solution (Equations 3 and 4): NiL3(CN)2 < > NiL3(CN)2 + L (3) . +2 . +2 NlL5 ;——> NiL4 + L (4) These equilibria are processes of intermolecular ligand ex- change, and proton nuclear magnetic resonance temperature studies were carried out on these systems to determine relative ligand exchange rates. In this latter study, similar equilibria (Equation 4) involving the complexes [CoL31Clo321 and [RhL5]B(C5H5)438 (L = P(0CH3)3) were also investigated in order to study the effect of the metal on ligand exchange rates. Both methoxy and methyl ligand resonances in the proton nmr are simple doublets and thus facilitated this study. Changes in the proton nmr spectra were induced by temperature changes and were related to these exchange processes (Equations 3 and 4). EXPERIMENTAL Materials.- Ni(CN)2-4H20 was prepared according to a method previously given39 and this blue solid was de- hydrated to buff colored Ni(CN)2 by heating at 1500 under vacuum for approximately 5 hr. The salts, Ni(ClO4)2-6H20 and Ni(BF4)2-6H30 were obtained from G. Fredrick Smith Chemical Co., Columbus, Ohio and Alpha Inorganics, Beverly, Massachusetts, respectively. The tertiary phosphorus compound, CH3PC12, was received from the Department of the Army, Edgewood Arsenal, Edgewood, Maryland as a gift and it was used without further purification. The dehydrating agent, 2,2-dimethoxypropane was obtained from Eastman Kodak Co., Rochester, N.Y. Trimethylphosphine was purchased from Strem Chemicals Im., Danvers, Massachusetts, and converted to the silver iodide complex‘0 which was stored at about -20° in a stOppered flask. Trimethylphosphite was purchased from Aldrich Chemical Company, Inc., Milwaukee, Wisconsin. The ligand, CH3P(OCH3)2, was prepared according to the method of Maierfl1 however, CH3PC12 was used as the starting mater- ial. The observed boiling point (bsoo 60-610) for CH3P(OCH3)2 agreed well with the literature.41 The ligand, (CH3)3POCH3, was prepared by heating a 1:1 molar mixture of (CH3)2PN(CH3)242 and CH3OH at 65° for one hr.and then 11 12 distilling the product at b730 58-590. The reactant, (CH3)2PN(CH3)3, was prepared by two methods: a reaction sequence outlined by Burg and Slota,“2 which gave very poor yields, and by heating (CH3)2PC143 with two moles of (CH3)3NH in Na-dried diethyl ether at -78° under a nitro- gen atmosphere. The (CH3)2NH was slowly added through an addition funnel with a pressure equalizing side arm. The reaction flask was also equipped with a very efficient stirrer and a dry ice condenser. After complete addition of (CH3)3NH, the reaction mixture was allowed to slowly come to room temperature while magnetically stirred. The amine hydrochloride was filtered off under suction, and the product was distilled at 95-1000/760 mm. Reaction of (CH3)2PC1 or CH3PC12 with CH3OH directly in the presence of base gave very poor yields of the phosphinite and phos- phonite, respectively. The 1H nmr spectrum of CH3P(OCH3)2 consists of two doubkfis at 1.13 ppm (CH3, JP = 8.5 Hz) CH and at 3.52 ppm (OCH3, JPOCH = 11.1 Hz), and Similarly, the Spectrum of (CH3)3POCH3 consists of two doublets at 1.19 ppm (CH3, = 6.0 Hz) and at 3.32 ppm (OCH3, J JPCH POCH = 13.5 Hz). The complexes [NiL5](ClO4)221, [C0L51Clo421, and [RhL5]B[C5H5]438 (L = P(OCH3)3) were prepared according to previously described methods. NilCHaP(OCHs )2]3(CN)20- T0 2.2 g (20 [1111101) Of Ni(CN)2 suSpended in 50 ml of acetone and cooled in ice water was 13 added 5.4 g (50 mmol) of CH3P(OCH3)2 all at once. The solu- tion was stirred for approximately 10 hr, filtered under nitrogen, and concentrated under vacuum to give a dark red solution. Red-orange crystals formed when the solution was cooled to -78°. These crystals were filtered, rapidly washed with -780 acetone, and dried in a stream of dry nitro- gen. Aggl, Calcd for C11H27N205P3Ni: C, 30.35: H, 6.22; P, 21.39; N, 6.44. Found: C, 30.21; H, 6.11; P, 21.48; N, 6.63. Ni[CH§P(OCH§)Z]2(CN)2 .- This yellow complex would typically appear as an impurity in the above five-coordinate complex. Treatment of Ni(CH3P(OCH3)2]3(CN)3 with anhydrous ethyl ether in a manner analogous to that reported earlier18 produces a yellow powder which is insoluble in acetone and diethyl ether. Also, if the above mentioned five-coordinate complex is maintained under a dynamic vacuum for an hour or more, conversion to Ni[CH3P(OCH3)2]2(CN)2 is apparent by formation of the yellow solid. Aggl, Calcd for C3H13N304P2Ni: C, 29.35; H, 5.52; P, 18.95; N, 8.57. Found: 0, 28.95; H, 5.17; P, 18.95; N. 8.50. Ni((CH3)3PoCH3]3(CN)3.- To 1.1 g (10 mmol) of Ni(CN)3 suspended in 50 ml of acetone and cooled in ice water was added 2.8 g (30 mmol) of (CH3)2POCH3 all at once under a nitrogen atmosphere and the mixture was stirred magnetically. 14 The solution was allowed to stir for several hr, and then it was filtered, concentrated under vacuum, and cooled to give red-orange crystals. Attempts at recrystallization in the presence of excess ligand yielded mostly the four-coor- dinate complex and repeated attempts at obtaining a satis— factory elemental analysis were unsuccessful. This complex could, however, be identified by its infrared spectrum in the solid state and further characterized by its absorption spectrum, also in the solid state. Ni[(CH3)3POCH3]3(CN)2.- As mentioned above, this yellow complex consistently appeared as an impurity in the five-coordinate complex and could be isolated from such solu- tions. Also, by allowing Ni(CN); to react with (CH3)2POCH3 in a molar ratio of 1:2 or less and in a manner similar to the above, this complex could be prepared easily. A32}; Calcd for C8H18N202P2Ni: C, 32.61; H, 6.13; P, 21.08; N, 9.53. Found: C, 33.04; H, 6.11; P, 19.78; N, 9.69. Ni[(CH3)3P]3(CN)2.- A 50 ml acetone suspension of 0.44 g (4.0 mmol) of Ni(CN)3 was evacuated on a high vacuum line and cooled with liquid nitrogen. The ligand, P(CH3)3. 0.91 g (12 mmol), liberated from the A91 complex by heating, was condensed onto this frozen suspension and the mixture was then allowed to warm to room temperature. Stirring (magnetic) was initiated as soon as possible and a dark red solution resulted. After the solution had stirred for about 15 1/4 hr, it was removed from the vacuum line, filtered, con- centrated, and cooled to -780 to give red crystals. Alter- natively, the acetone solution may be evaporated to dryness and the resultant red-orange powder recrystallized from g: hexane or diethyl ether. A321, Calcd for C11H37N3P3Ni: C, 38.95; H, 7.97; P, 27.41: N, 8.26. Found: C, 38.53: H, 8.02: P, 26.59; N, 8.12. Ni[(CH§)3P]3(CN)3.- Prolonged evacuation of Ni[(CH3)3P]3(CN)3 would produce this yellow solid which could be recrystallized from an acetone-diethyl ether mix- ture (10:1). 5&2}, Calcd for C8H13N2P3Ni: C, 36.50; H, 6.85; P, 23.60; N, 10.65. Found: C, 36.39: H, 6.88; P, 23.41; N, 10.43. [Ni[CH3P(OCH3)a]§)(ClO!)2.- In an acetone-2,2-di- methoxypropane“ solution (10 ml of each) was dissolved 0.73 g (2.0 mmol) of Ni(ClO4)3‘6H20. This solution was stirred at room temperature fcr1/2 hr and then 1.25 ml (> 10 mmol) of CH3P(OCH3)3 was added all at once to give an orange precipitate and a similarly colored solution. The solution was stirred for 1/4 hr, filtered, and the solid was washed well with diethyl ether and dried under vacuum. Recrystallization can be accomplished from a very concen- trated acetone solution. 16 5331, Calcd for C15H45C12013P5Ni: C, 22.55: H, 5.64; P, 19.42; Cl, 8.90. Found: C, 22.42: H, 5.63; P, 19.06; Cl, 8.94. [Ni[(CH3)3POCH3]5](C104)2.- This compound was pre- pared in a manner identical to that listed above for {Ni [CH3P (OCH3 )3] 3 I (ClO4I3 . Anal. Calcd for C15H45C13013P3Ni: C, 25.08: H, 6.27: P, 21.55: Cl, 9.88. Found: C, 25.27: H, 6.19; P, 21.47; Cl, 10.08. {Ni[(CH3)3P]4}(BF4)3.- The air over a 25 ml solution of absolute ethanol and 0.34 g (1.0 mmol) of Ni(BF4)3-6H30 was evacuated on a high vacuum line and the solution cooled with liquid nitrogen. Trimethylphosphine (0.30 g, 4.0 mmol) was added to this solution in a manner as indicated above. A pale red solid formed as this mixture was allowed to warm to room temperature and stirred magnetically. The solution was removed from the vacuum line, filtered, and the solid washed well with diethyl ether and dried under vacuum. The complex dissolves in ethanol only very slowly to give a yellow solution, in acetonitrile to give a purple solution which eventually becomes colorless upon standing, and in acetone to give a red solution which rapidly decomposes to give a nearly colorless solution. 523;, Calcd for C13H3533P8P4Ni: C, 26.80; H, 6.70; P, 23.10. Found: C, 26.65; H, 6.56; P, 22.98. 17 Physical Measurements.- Elemental analyses were per- formed by Galbraith Laboratories, Inc., Knoxville, Tennessee. Infrared spectra were obtained in Nujol mulls by use of either a Beckman Model 12 grating Spectrophotometer or a Perkin-Elmer Model 225 grating Spectrophotometer. Reflec- tance spectra were obtained by use of a Bausch and Lomb Spectronic 600 with BaSO, as the diluent and reference. Spectra in the solid state were also obtained by use of the Cary Model 14 spectrophotometer and a Nujol mull of the com- pound which was “painted” onto filter paper.‘:5 A reference cell with Nujol only was needed. These "filter paper cells“ were simply taped to the wall of the cell compartment nearest the detector; however, little difference in the spectra were noted when the Opposite wall was used. Spectra were also recorded by use of Nujol mulls mounted between quartz windows. In the visible region, almost identical Spectra were obtained by the two methods, however, in the ultraviolet, the resolution was very poor when quartz win— dows were used. Solution ultraviolet-visible spectra were recorded in CH3C12 by means of Unicam Model SP.800B and Cary Model 14 SpectrOphotometers. The CH2C13 was previously dried by refluxing over CaHg. Samples were prepared under nitrogen in ground glass stOppered cells in order to mini— mize concentration changes while the Spectra were recorded at various temperatures. A thermostated methanoldwater bath, reproducible to i 0.10°C, was attached to the Spectropho- tometer; thus, the samples could be recorded at different 18 temperatures. Proton nmr spectra were recorded by use of a Varian Model A56/60D analytical Spectrometer, equipped with a Varian Model V-6040 variable temperature controller. Low temperatures were determined by measuring the chemical Shift difference for methanol, while ethylene glycol was used for high temperatures. All of the Spectra were recorded in CH3C12 solutions except the high temperature Spectrum of {Rh[P(OCH3)3LdB(C3H5)4 which was recorded in CH3CN: both solvents were used after being dried over CaH2. The high temperature Spectra were taken in thickdwalled nmr tubes which were sealed under an atmosphere of N3. Chemical Shifts_ were measured by using the solvents as references (CH3C12, 6 = 5.30 ppm: CH3CN, 6 I 2.00).46 Magnetic susceptibilities ‘were determined by Evans' nmr method47 in CH3C13 solutions. RESULTS AND DISCUSSION Five-coordinate complexes of the type NiL3(CN)2(L = (CH3)3P,37 (CH3)3PoCH3, CH3P(OCH3)3, P(oCH3)3)18 are produced when Ni(CN)3 is allowed to react with an excess of the above ligands in acetone solution. The corresponding NiL3(CN)3 complexes are prepared by subjecting the five- coordinate complexes to high vacuum at room temperature, or by allowing Ni(CN)3 to react with these ligands in a 1:2 molar ratio. Complexes of the type [NiL5](C104)3 (L = (CH3)3POCH3, CH3P(OCH3)2, P(OCH3)3)21 are isolated from acetone-2,2—dimethoxypropane solutions of Ni(ClO4)2-6H20 and excess ligand. When L = (CH3)3P no five-coordinate complex could be isolated. However, the complex, Ni[(CH3)3P]4(BF4)2. was prepared when Ni(BF4)2'6H30 was allowed to react with (CH3)3P in ethanol solutions, similar in manner to the prep- aration Of Nil (CH3 )3P] 4 (C104 )2 .37 Infrared Spectra Complexes of Ni(CN)3 which were prepared in this study permit easy assignment of the cyanide stretching frequency. The mull infrared Spectrum of each NiL3(CN)2 (L = (CH3)3P, (CH3)3P0CH3, CH3P(oCH3)3, P(0CH3)3) complex consists of one Sharp absorption (Table II), which is consistent with a trigonal bipyramidal structure with the cyanide groups in 19 20 Table II. Infrared data, cyanide stretching frequencies. L vCN(cm_1) Structural Assignment NiL3(CN)3 (CH3)3P 2097 L CN (CH3)3P0CH3 2100 h“ CH3P(oCH3 )2 2104 CN P(OCH3)3 2125 N1L.(CN)2 L {N (CH3 )3P 2106 b N j NC 3 (CH3)3POCH3 2129, 2113 L JN CH3P(0CH3)2 2122, 2103 Ni L CN a P(OCH3)3 2149, 2144, (Figure 4) 2127, 2122 aSee reference 18. 21 E£2n§_apical positions. Infrared Spectral data in the cyanide region for the corresponding NiL3(CN)3 (L = all ligands in the above series) complexes is also presented in Table 11. When L =(CH3)2POCH3 or CH3P(OCH3)3, two cyanide stretching frequencies are observed, and when L = (CH3)3P a Single cyanide stretching frequency is observed. These data are interpreted in terms of a Sigrsquareoplanar geometry for the complexes of (CH3)3POCH3 and CH3P(OCH3)3, and a Egaggf Square—planar arrangement for the complex of (CH3)3P. When L = P(OCH3 )3’ four cyanide stretching frequencies are observed. Coskran, g£_§£.13 have postulated that a dimer is present in the solid state, which contains both bridging and terminal cyanides (Figure 4). (CN) NC CN (C N) Figure 4. Proposed dimeric structure of Nil.3(CN)3 (L = P(0CH3)3).18 Absorption Spectra Ni(CN); Complexes.- The band maxima in the absorption Spectra for both the four- and five-coordinate Ni(CN)2 com- ;pleXes are listed in Table III. These Spectra are recorded 22 .wooasonm u Anmv Asmvnms.omm Asmvmmc.omm Armvcmc.msm Asmvoms.nsm Asmvmfls.mmm Armvmec.mmm Armvomc.mnm Aruvcue.mmm Anmvoem.flsmyocm.oom Asmvnmc.cmm.firmvccm.oeu Acmvcsc.Armvmmm.mmm Acnvcms.msm.sem Armvmes.flsnvmmm.ceu Asmvmes.mmm.sem Asmvcue.xsmvnsm.ccu Acmvcms.mmm.eem msmtmem Accounvmmm Asmvcsvxfirmvmmmtncm Acuvose.flsmvomm Asmvcms.xsmvmmm.ncm Acmvcms.xrmvmmm.mmm Asmvcmc.firnvmsm.ocu Accounvmcm.ocm «Azove_mmAcmovaaz «AzovaOmAmmuv_Hz «Azova_cmoom«Ammovahz «AzovmmemoomaflmmOVHAz «Azova_aflmmoovmcmo_az «Azovm_ahmmoovmsmo_az «Azovamafimmoovm_az «Azovmflmxmmoovm_az dwcmmfla mmooxm .uHOnmUV Adev mCOADQHomnfl Issuance Aifiv nCOADQHOmA< Acumum owaomv Aiev mcowumuomnm ocsomaoo .moxoamaoo «Azuvwz Mom sumo Hmnuoomm cowumuomné .HHH magma 23 in both solid and CH2C13 solution media for each complex, along with the effect of adding excess ligand. For each five-coordinate complex in solution, there is a very intense absorption in the region 290-300 mu. When excess ligand is added to these solutions (and also to the reference cell), the absorption is not present (Table III). This Spectroscopic feature is demonstrated in Figure 5 for CH3C13 solutions of Ni[P(OCH3)3]3(CN)3, which are typical of the five-coordinate complexes. The presence of this ab- sorption (290-300 mu) in the solution spectra of the four- coordinate complexes, and the absence of this absorption in the solid Spectra and solution Spectra with excess ligand of the five-coordinate complexes, suggests that the band is due to the four-coordinate complexes (Figure 6). The rather constant position of this band in all the Spectra of the four-coordinate complexes suggests that it is a Ni —> n*CN charge-transfer band.19 These experimental results suggest that an equilibrium of the following type is present in solutions of the five-coordinate complexes: > NiL2(CN)2 + L (3) NiL3(CN)2 <__—. This equilibrium is also consistent with the experimental Observation that four-coordinate complexes can be easily isolated from solutions of the five-coordinate complexes. The intense absorption in the visible region, with its low energy shoulder (325-455 mu), that appear in both the solid and solution spectra of the five—coordinate complexes, is assigned to the (e“)4(e')4 -—> (e")4(e')3(a1)1 24 .ocmmfla mmmuxm £Dfl3 Amv CH coausaom Anv .aHORmo an «Azovmmmfimmoovaaez Amv «mo Esuuommm manwmfl>lu0H0H>mHuHD .m onsmam 25 onv obs .m musmflm 38:35.20; 0mm nmm com nun od 3 8 .33 cauoqaosqv E! 26 wumum oflaom 0V .«HONTO Gd «Azovu .mHONmo as «Azovmmmm mmoVHRZ .«mzovaOnmmmUvgaz anlsmev Hz Ant Amv «m0 Esuuoomm cofiumuomnm .o ouomfim III}: 1 27 .m musmfim 355923.03 0mm aauoqlosqv 28 transition in complexes of D3h symmetry3 (Figure 2). Daw— son, g£_gl,,33 have discussed the splitting of this band (p. 5), and their argument implies that this band Should become more symmetrical at lower temperatures (~a77°K). A dipole forbidden transition, (e")4(e')4 -+ (e")3(e')4(a1)1, v2, of low intensity is expected at shorter wavelength than v1 (Figure 2), but overlapping of intense ultraviolet ab- sorptions and the above allowed absorptions prohibits accur- ate assignment of its position. These assignments are in agreement with Similar assignments which have been made for other five-coordinate Ni(II) complexes of tertiary phosphorus, ligands.3 The band assignments that are proposed above assume that these five-coordinate complexes have trigonal bipyra- midal structures (D in the solid state and in solution. 3h) However, it may be argued that these Spectra are those of a square pyramidal complex (C4v)' Since three spin-allowed transitions are predicted for these complexes34 (Figure 3). The proof of this argument would rest in the low temperature absorption Spectra, as these three bands in C4v symmetry are independent of temperature, and the two low energy bands (325-455 mu) would persist at low temperature (~a77°K), contrary to the predictions for these bands in D3h sym— metry. At this time, the geometry of these five-coordinate complexes cannot be unequivocally assigned, but the band assignments have been made on the basis of the trigonal bi- pyramidal structure, because it is predominant in Observed 29 complexes of Ni(II) with monodentate ligands“’"‘29 and it is predicted by ligand-ligand repulsion theories.25 When the absorption spectrum of each five-coordinate complex in solution is recorded over a narrow temperature range (0-30°C), isosbestic points are observed. This sup- ports the proposed equilibrium (Equation 3), since these isosbestic points indicate an equilibrium.between two ab- sorbing Species. Figure 7 illustrates the isosbestic point obtained for Ni[P(0CH3)3]3(CN)2 and is typical of these five-coordinate complexes.l It can also be noted (Figure 7) that the low energy bands due to the five-coordinate com— plexes (325-455 mu) increase in intensity as the temperature is allowed to decrease, while the bands assigned to four- coordinate complex (290-300 mu) decrease in intensity. This latter observation indicates the proposed equilibrium (Equation 3) shifts toward five-coordinate complex upon a decrease in temperature. Estimates of the equilibrium constants, or instability ), for the system Kinst NiL3(CN)2 ———> NiL3(CN)3 + L (3) < constants (Kinst have been obtained for these five-coordinate complexes. {rhese calculations were facilitated by the spread in the loand maxima positions for the four- and five-coordinate «complexes. The extinction coefficient (8) for the intense .absorption in the visible region (the high energy band of ‘VI was used ) for each five-coordinate complex was deter- rnined from the apparent absorption maxima observed when 30 aaOamo ca «AZOVSHSAmmoovm_az mo Ednuommm oaflawa>uu0a0fl>muuas may no Omsmno musumnomaou mo Doommm .b oudmflm 31 com ems oov .5 madman 3813:3303 cmn can con man . . Qo a -vo rw.° UV! m Q. I m D . m m .m .u.— s I 00 ea _ .0; .oo . Q N 32 extremely large excesses of ligand were added to solutions of the five-coordinate complexes of known concentration. These values of s were then used to calculate the concen- tration of five-coordinate complex present at equilibrium with no added ligand and the difference between this concen— tration and the initial concentration was taken as the four- coordinate complex concentration and the dissociated ligand concentration, also at equilibrium. The calculated equi— librium constants (or instability constants) are recorded in Table IV along with the calculated e's. The uncertainty in these values is such that (CH3)3POCH3 and CH3P(OCH3)2 may be interchanged but it is definite that these constants follow the order (CH3)3P > P(oCH3)3 > CH3P(OCH3)3 ~ (CH3)3P0CH3. Surprisingly, there is no regular order in K. inst paral- lelling the systematic change in the nature of these phos- phorus derivatives. This particular order, however, is observed elsewhere. The band maxima (Table III) for the allowed d-d transitions (see later) are in approximately the same order indicating that the P(CH3)3 complex has the 'weakest ligand field interaction. The large extinction coefficients that are obtained for these five-coordinate complexes are similar to previous observations,3o15 and dem— onstrates the strong mixing of ligand and metal orbitals in these complexes. 33 Table IV. Kinst for NiL3(CN)2 complexes at 25°. 4 L Kinst X 10 -5 -1 (M) (mole cm ) (CH3)3P 2.5 2390 (CH3)3P0CH3 0.4 2480 CH3P(ocH3)2 0.5 2506 P(0CH3)3 1.7 2756 Ni(ClO!)2 and Ni(BF$)£_ComplexeS.- The band maxima in the solid state and solution spectra of the five-coordinate [NiL3](Clo,)3 (L = (CH3)3P0CH3, CH3P(0CH3)3, P(0CH3)321) complexes and the fOur-coordinate Ni((CH3)3P]4(BF4)2 com- plex are presented in Table V. The ultraviolet-visible Spec- tra of the five-coordinate complexes in the solid state and in solution to which ligand has been added (see later) are simi- lar to those reported for {Ni[P(oc::H)3(CH3)3]3}(Clo,)3,21:48 *which is trigonal bipyramidal, according to a three-dimen- sional X-ray crystal structure analysis.29 Assignment of the bands in the five-coordinate complexes is accomplished xvith the assumption of a similar structure, and are similar -to the assignments in the NiL3(CN)3 complexes. The low (energy absorption bands in the spectra of these complexes (400-560 mu), assigned to the (e")‘(e')4 -9 (e")4(e')3(a1)1, ‘v1 transition (D3h symmetry) (Figure 2), are Split in the solid state and solution with added ligand33 (p. 5). The Imigh energy band (300—320 mu) is assigned to the dipole 34 .uooasorm u Anne Isucomc.cemiasuvccm Ismccom.ccc.xsnccam Arncmmm.nse.xsmcocm Armcocv.mac.xraeocm oos.AsmCoov.omu cos.ocm AccounCmNe.Aancmm freeman.osv.1smcomm Apaches.c~s.xsnvofim Aomomncmme.AsuCoum Igneous.sov.Araooom Igneous.eoe.xrmcmcm al.macm.naaiasoc_azc «AeoaovmsnamooaaxcmocLazw «AcoHoCHanaxamoocmmmo_azw «AeoHoCMSHSAamooCSLAzL Aocmowa mmmoxm .«HOamUv Ajay meowumHOmnE Auavumuv Aoumum paaomv Aiev msoaumnomnd Anew mcoaumnomnd ocsomaoo .moxoamsoo «Avmmvaz osm «AvoHovaz How sumo Hmuuoomm documuomnd .> wanna 35 forbidden, (e")4(e')4 -—> (e")3(e')4(a1)1, v2, transition. Isosbestic points are obtained in the solution spectra of {Ni[P(OCH3)3]5}(ClO4)2, when the spectra are recorded over a small temperature range (Figure 8). Similar isos- bestic points are obtained at 362 mu and 424 mu for {NilCH3P(OCH3)2]5](C104)2. (The complex {Ni[(CH3)3POCH3]5}(C104)3 decomposes rapidly in CH2C12 and did not allow for the observance of isosbestic points.) Therefore, an equilibrium Similar to that found in the NiL3(CN)2 complexes is proposed: ___> mun” + L (4) < [NiL5] +2 Further evidence for this equilibrium is given by the ob- servation that the solution Spectra of the five-coordinate complexes are altered upon addition of ligand. The posi- tions of the low energy bands shift to lower energies when ligand is added, approximate the pOSitions of the absorptions in the solid state Spectra (Table V), and indicate that the four- and five-coordinate bands overlap to a large extent in the solution spectra. Indeed, the solution spectrum of [Ni[(CH3)2POCH3]5}(ClO4)3 more closely resembles the spec- trum of the four-coordinate (Ni[(CH3)3P]4](BF4)3 complex and the five-coordinate Species is probably largely dissoci- ated in solution. Unlike the corresponding NiL3(CN)3 complexes, the [JN'in]+2 complexes have not been isolated. Their existence, 'though, is tenable in view of the fact that (CH3)3P does 36 .mHOnmo ca «AvoHovmonmAnmoovasz now omcfimuno mucflom oaumonmomH .w ousmflm 37 .w musmam as; face—20>) 00m. onv ooh om” mum com tron 0:339: '21:? --—'-—_.— , ”M"? 0 r0. 8.... 6A .__I. 0. o 0! O O aauoqlosqv 38 form an isolable NiL4+2 complex, and this complex will as- sociate with added (CH3)3P in solution to form a five-co- ordinate complex. The solution and solid state spectra of this four-coordinate complex are similar (Table V), and similar to other square-planar, Ni(II) complexes.3‘ How- ever, when excess (CH3)3P is added to a solution of (Ni[(CH3)3P]4](ClO4)2, a dark blue color results, and the Spectrum of this solution resembles those of the five-co- ( ordinate complexes in this work (Table V), and also that I of the blue complex,[Ni(TAP)Br]ClO4 (TAP = P(CH3CH2AS(CH3)3)3).49 It is worthwhile to comment further on the associative nature of Ni[(CH3)3P]42+ in solution. When the tetrafluoro-i borate salt of this complex is dissolved in CH3CN, a dark purple solution results. Addition of a large amount of diethyl ether to this solution causes trace amounts of a blue solid to be precipitated. This solid exhibits a VCN at ~v2300 cm_1, indicative of a coordinated CH3CN group,50 which is not surprising, in view of the associative character of this complex. The proposed five-coordinate complex present in solution when excess (CH3)3P is added to [Ni[(CH3)3P]4]2+ has not been isolated, however. The dark blue solution that forms when excess ligand is added rapidly fades to pale yellow and addition of diethyl ether to this solution causes precipitation of a pale yellow solid. More- over, when diethyl ether is added to the initial blue solu- tion, the same yellow solid is obtained. Attempts at recrystallization inevitably produced light green solutions 39 indicative of solvated Ni(II). Elemental analyses of the compound were inconsistent and offered no information about it. Nevertheless it is tempting to speculate that it is a Ni(O) complex of (CH3)3P. The complex, [Ni((CH3)3P]4}, has been prepared and its color is light yellow}6 The mecha- nism for its formation in this work might be considered to be similar to that used for preparation of Ni(O) phosphite complexes!“52 in which excess ligand causes reduction of Ni(II) to Ni(0) and the excess ligand is oxidized. Reference to Table V, shows that in the five-coordinate Ni(ClO4)3 and Ni(BF4)3 complexes, the order of increasing 3 ligand field strength is (CH3)3P < (CH3)3P0CH3 < CH3P(0CH3)3 < P(0CH3), This order is Obtained from those spectra to which excess ligand has been added. This order of ligand and field splitting closely parallels the stability of the five-coordinate complexes. Thus, the five-coordinate com- plex with (CH3)3P is unstable and cannot be isolated from solution, while the isolable P(0CH3)3 complex appears to be stable in solution and its absOrption spectrum is affected very little upon the addition of ligand. Therefore, it is proposed that the above order of ligands reflects the stabil- ity of the five-coordinate [NiL3]+2 complexes toward the corresponding [NiL4]+2 complexes. Electronic Effects in Five-Coordinate Complexes.- In five-coordinate, d3, complexes with trigonal bipyramidal 40 structures four of the filled metal d orbitals can be used in w-bonding, if 7- acceptor ligands are available.19 In square-planar, d3, complexes only three d orbitals can be used in wébonding, for reasons of symmetry.19 Thus, a good w— acceptor ligand would favor the formation of a five-coordin- ate complex over a four-coordinate complex. In the ligands considered in this study, the w—acceptor ability would be expected to decrease in the order P(0CH3)3> CH3P(OCH3)2 > (CH3)3POCH3 > (CH3)3P.53 The more electrOnegative OCH3 groups would have the effect of lowering the energy of the empty phosphorus 3d orbitals, thus enhancing their #- acceptor character.5‘ At the same time the increased elec— tronegativity of OCH3 groups should withdraw electron density from the phosphorus atom, and produce a contracted and stable o-donor orbital, which would be less suited for otbonding to the relatively more expanded Ni 43 and 4p orbitals.1’ Since ligands which bond to these S and p orbitals in preference to d orbitals favor the five-co- ordinate structure, the tendency toward five—coordination Should decrease in the order (CH3)3P > (CH3)2POCH3 > CH3P(OCH3)3 > P(0CH3)3, according to oébonding arguments?“19 .The order of ligand V-acceptor ability follows the order of proposed stability of the [NiL3]+2 complexes rela- tive to [NiL3]+2 complexes. An opposite trend is predicted by the above oébonding arguments. This suggests that #- bonding effects are of primary importance in determining the relative stability of these five-coordinate complexes. 41 Support for this hypothesis is given by the observation that attempts to isolate the complex, {Ni[(CH3)3P]5}(BF4)3 (p. 39), result in the apparent reduction of Ni(II) to Ni(O). This suggests that the increased metal electron density could not be sufficiently relieved by these poorer v-acceptor ligands.5‘ The stability of the NiL3(CN)2 complexes toward the correSponding NiL2(CN)3 complexes decreases in the order (CH3)3P0CH3 ~ CH3P(0CH3)3 > P(0CH3)3 > (CH3)3P (Table IV). This does not reflect the Order predicted by either n-ac- ceptor or o-donor arguments, but suggests that the stability of five-coordinate complexes is determined by a more equal combination of these two effects. This is consistent with observations that suggest the CN- anion is a good n-ac- ceptor.20 Thus, the cyanide groups reduce the electron density on the metal and the ligand n-acceptor ability becomes less important in these NiL3(CN)2 complexes than in the [NiL5]+2 complexes. Proton Nmr Spectra Three classes of complexes were studied, namely, the four-coordinate complex, Ni[(CH3)3P]2(CN)2, the five-coord- inate complexes, NiL3 (CN)2 (L = P(0CH3)3, CH3P(OCH3)2. (CH3)2POCH3, (CH3)3P), and the five-coordinate complexes ML5n+ (Mn+ = Ni“, L = P(0CH3)3 and CH3P(0CH3)3; M1M = 21' Co+ and Rh+ 38, L P(0CH3)3). The complexes, NiL2(CN)2 42 (L = (CH3)3P0CH3, CH3P(0CH3)3. P(0CH3 )3), {Ni[(CH3)3PoCH3]3)- (€104)3 and {Ni[(CH3)3P]4}(BF4)3 were not studied here be- Cause of insolubility and instability in solution. The Co(I) and Rh(I) complexes were included in this study be- cause they permit comparisons of the effect of the metal on ligand exchange rates. The temperature dependent proton nmr spectra obtained for all of these complexes are similar over their temperature ranges. For each complex there are three distinct Spectral features that occur at Specific temperatures. These spectral features are: doublet (high temperature Spectrum), sharp singlet (intermediate tempera- ture Spectrum), and multiplet (low temperature Spectrum). These classes of complexes will be discussed in turn. Ni[(CH3)3P]3(CN)3.- The temperature dependent proton nmr spectrum of this complex is Shown in Figure 9. At probe temperature (~'38°) the Spectrum is a sharp Singlet at 6 = 1.52 ppm. If the solution is cooled, this singlet broadens and collapses and finally at —65° a three-line pattern is observed. Alternatively, if the solution is heated from the point of the Singlet, this resonance is observed to broaden and finally a sharp doublet forms at 97° with a coupling constant of about 10 Hz and the chemical shift is slightly downfield from the singlet. The general features of these Spectra are identical to those obtained for some Eggggfsquare-planar Pd(II)55 and Rh(I)56 methylphosphine complexes. In these systems the room temperature proton nmr spectra of the pure complexes 43 Figure 9. 1H nmr spectrum of Ni[(CH3)3P]2(CN)2 in CH2C12 at various temperatures.. 44 a.» 44° -56° 1 F If 1| PPM 1.70 1.50 1.30 Figure 9. 45 were virtually coupled triplets and upon successive addi- tions of ligand they collapsed forming singlets, and finally, doublets formed at high ratios of added ligand to complex. An intermolecular process of ligand exchange between added ligand and complex was prOposed to account for these Spec- tral changes. The similarities between these spectra and those observed in this work for Ni[(CH3)3P]2(CN)3 are obvious and we Similarly propose that a process of intermolecular ligand exchange is occurring. However, there is one important difference in the system studied in this work and that of the Pd and Rh complexes and that is that no excess ligand has been added to solutions of the nickel complex. Therefore, this four~coordinate nickel complex must be dissociatively unstable with respect to free ligand and a lower coordinated Species. The absorption spectrum of a methylene dichloride solution of Ni[(CH3)3P]3(CN)3 pro- duces an isosbestic point at 345 mu.when the Spectrum is recorded over a relatively small temperature range. (Fig- ure 10.) Thus, there must be more than one absorbing Species in solution and this is in agreement with a dis- sociative process occurring in solution for which the fol- lowing equilibrium is proposed: NiL3(CN)3 <___:_ 1~1iL(cn)2 + L, L = (CH3)3P (5) This equilibrium similarly accounts for the intermolecular ligand exchange process. The three-coordinate nickel com- plex in the above equilibrium can be classed as a fourteen 46 . .u « HO mu CH «Azovmmmmmmmovafiz mo Esnuoomm wanwmfl>luoaofl>muuas mo coauuom .OH oudmflh 47 .oH ousowm AiEv faced 962, can nun . 0.0 J O... oauoqlosqv 48 electron complex (d8 metal ion and three two-electron donor ligands). Such systems are not at all common, but there is no other Obvious recourse in the interpretation of the data. One might speculate, however, that the three-coordinate complex dimerizes through cyanide bridging forming [L(CN)Ni(CN)3Ni(CN)L] . In this complex there would be a square planar arrangement about each four-coordinate Ni. Bridging cyanide groups are, of course, well known in some Ni(II) systems.57 Other equilibria that may be considered to be present in this solution do not account for the observed 1H nmr and absorption Spectra. For example, the complex Ni[P(OCH3)3]3- (CN)2, has been prOposed to dimerize through cyanide bridg- ing.18 A Similar monomer-dimer equilibrium for Ni[(CH3)3P]2(CN)2 would not involve ligand exchange and thus, could not explain the 1H nmr Spectra. Also, it is known that the four-coordinate complex, Ni[(CH3)3P]3(CN)2, will associate with ligand to form a five-coOrdinate complex (see later), however, this equilibrium is not present in solutions of the pure four-coordinate complex because a characteristic isosbestic point is observed at 335 mu for the four-coordinate - five-coordinate equilibrium (p. 29) and only after additional ligand is added toIsolutions of the four-coordinate complex is this isosbestic point observed. Therefore, the three Spectral features which are ob- served, doublet, singlet, and multiplet (triplet) are at— tributed to an intermolecular ligand exchange process and 49 they represent the Spectra for processes under fast, inter— mediate, and slow rates of exchange, respectively. It will be convenient to delay discussing the doublet and singlet Spectra until after the results of the five-coordinate com- plexes are presented; but the triplet,or Slow exchange Spec- trum.will be discussed now. Numerous examples of Egaggfsquare-planar d8 metal com- plexes58 of organOphosphorus ligands have exhibited proton nmr Spectra Similar to the triplet spectrum observed for Ni[P(CH3)3]2(CN)2 at -65°. This is an example of an XnAA'Xn' type spectrum where X and A are the H and P of one ligand respectively, X' and A' are the H and P of the second ligand, and n = 9. The separation of two outer lines of the triplet equals |J ,| and the AX+JAX central line, which will vary in intensity from complex to complex, can be qualitatively used as a measure of JAA.:59'5° the more intense the central line, the larger qAA" There- fore, the triplet spectrum for this trans—square-planar nickel complex with an intense central resonance means that the two ligands are strongly coupled to each other through the metal. For a complex exhibiting ligand exchange this coupling could only occur under the limiting conditions of very slow ligand exchange. Therefore, the triplet spectrum is representative of the ligand-ligand interaction under conditions of Slow exchange for this trans-squareoplanar arrangement of ligands about the metal atom. 50 NiL3(CN)z Complexes.— The five-coordinate Ni(CN)2 complexes of (CH3)XP(OCH3)3_i (x = 0, l, 2, or 3), which are trigonal bipyramidal with p£3p§_apical cyanides,18 exhibit very Similar temperature dependent 1H nmr Spectra to the four-coordinate complex, Ni[(CH3)3P]3(CN)3 (see Figure 11). For these complexes, hOwever, the low tempera- ture multiplet is a poorly resolved quartet contrasted to the triplet observed for the four-coordinate complex. For the complexes of CH3P(OCH3)2 and (CH3)2POCH3, two resonances are observed for the two types of ligand protons in each complex (Figure 12). Therefore, two doublets and two sing- lets are observed for each complex as the temperature of the solution is lowered: however, only for the OCH3 protons is the quartet Splitting observed while for the CH3 protons only a broad Singlet is recorded. The failure to resolve a quartet is surprising because the OCH3 proton Splitting is observed in the same ligand and furthermore, the CH3 proton splitting is readily observed in the five—coordinate (CH3)3P complex (see Figure 11). To account for these temperature dependent Spectra we again propose a process of intermolecular ligand exchange which is consistent with the postulate that all of these complexes are dissociatively unstable according to the pro- posed equilibrium: NiL3(CN)2 :3 NiL3(CN)2 + L (3) 51 Figure 11. 1H nmr Spectrum of Ni[(CH3)3P]3(CN)2 in CH2C12 at various temperatures.. 52 42° —-24° A _390 N0 L7 1 l 1 l 1.60 1.40 1.20 ppm Figure 11. 53 ANA «moovmmmo .mmoomfimmov ..monsumummEmu msoflnm> um uHUnmo CH u AV «Azuvmqflz mo Eduuoomm Had we .Na musmflm 54 «IU omv- «Amzuornxu H. «.80 .NH mHsmHm omd ‘-“ 55 The poorly resolved quartet observed at low temperatures represents a spectrum of the type, AA'AFXhXh‘Xh", where A and x have similar notations as before. This assignment is based primarily on the similarities in the overall Spec— tral changes for these five-coordinate complexes and the previously mentioned four-coordinate complex. In the latter case, the triplet Spectrum due to ligand-ligand interaction through the metal is well understood"9 and:fl:is believed that under similar limiting conditions of Slow exchange, coupling through the metal between the three ligands in the equatorial plane does take place for these trigonal bipyramidal com- plexes and gives rise to the quartet spectrum. A simple experiment to demonstrate the intermolecular nature of this exchange process is summarized in Figure 13. If P(0CH3)3 is added to the five-coordinate complex, Ni[P(OCH3)3]3(CN)2, at 54° only one doublet is observed which is Slightly upfield from the original doublet for the complex only. If this solution is cooled, the doublet col- lapses, a sharp singlet forms, and eventually a quartet forms similar to a solution of pure complex, but in addition, a doublet due to the added P(0CH3)3 appears at this low tem- perature. Therefore, under conditions of fast and inter- mediate exchange rates the observed Spectra are representa- tive of a weighted average of all Species in solution and the added P(0CH3)3 has been averaged into the above equilib— rium. However, under conditions of slow exchange, the doublet Spectrum for the added P(0CH3)3 is Observed along 56 Figure 13. 1H nmr Spectrum of: (a) Ni[P(OCH3)3]3(CN)3 in CH2C12; (b) solution in (a with . added P(0CH3)3; (c) solution in (b) when cooled.- . _ 57 54' b -24' C -75' C k I. : 4. ; a 4 2. 4mm 4.00 3.80 3.60 3.40 3.20 Figure 13. 58 with the quartet for the complex. Furthermore, this process is reversible as one doublet appears if the above solution is warmed to 54°. Further support for the proposal that the doublet Spec— trum in these complexes represents an average of all species in solution comes from the experiment which is summarized in Figure 14. AS successive amounts of (CH3)3P are added to the complex, Ni[(CH3)3P]3(CN)3, at 420 (the temperature at which the 1H nmr Spectrum of this complex is a doublet), the separation of the doublet is observed to decrease, pass through zero, and form a new doublet with physical constants very much like that of pure (CH3)3P. Therefore, under these conditions of fast ligand exchange, coupling between ligands through the metal has been removed and a doublet results which is simply J but whose coupling constant and PCH' chemical shift is strongly influenced by the metal ion com- plex. The addition of free ligand heavily weights the solution in favor of uncomplexed ligand as is indicated by the changes in chemical shift and J . An interesting PCH point to note is that at a mole ratio of added ligand to complex of approximately 7.6, the coupling passes through zero which means that free and complexed ligands have cou- pling constants of opposite Signs. Similar observations have been observed for other methylphosphine complexes“:82 and it is prdbable that J for the complexed ligand is PCH negative since a positive Sign has been proposed63 for free (CH3)3P. 59 Figure 14. 1H nmr spectrum of Ni((CHa) P]3(CN)3 upon successive addition of (CH3)3P.- 60 [L] I’PCHI TNil3(CN)2] 9.0 0.0 4.4 H 2.7 2.1 0.0 7.6 J - 1.7 12.8 J .2 ; e 1 4r— ¢ 499'" 1.50 1.30 1.10 0.90 Figure 14. 61 This same type of experiment was also performed by suc- cessive addition of small amounts of the corresponding four- coordinate complex, Ni[(CH3)3P]2(CN)2. Addition of this complex heavily weights the equilibrium in favor of the four-coordinate complex and the 1H nmr spectrum indicates this fact by a Slow change in chemical shift and J to PCH' the values of the four-coordinate complex at room tempera- ture (0 = 1.52 ppm and J = 0.0 Hz). PCH If CH3P(OCH3)3 is added to a solution of Ni[CH3P[OCH3)2]3(CN)2, the doublet for the ocn3 protons Shifts upfield but there is essentially no change in the coupling constant. The CH3 protons similarly Shift upfield: however, the coupling constant passes through zero as be- fore (Figure 15). Therefore. J (coupling to the OCH3 POCH protons) does not change Sign on going from free to com- plexed forms but JPCH does. The same result is observed for complexes of P(0CH3)3 and (CH3)3POCH3. Since JPOCH has been predicted to be positive in the free ligands,a3'64 it is also positive in these complexes. However, JPCH appears to be negative in complexed ligands, since it is positive in the free ligands.63 The sharp singlet which is observed at intermediate temperatures and which is representative of an intermediate rate of ligand exchange, indicates that the phosphorus and hydrogen nuclei have been decoupled. The temperatures at which the singlets for each complex reach a maximum intensity are listed in Table VI. The mechanism for this decoupling 62 .nfiemoovmemo mo coeufloow com: «avamu ca «Azovmmuflnmooamnmuaflz mo Eduuuomm use me .oH «Homes 63 .mn musmflm om _ o: o: end one... owv _ _ _ _ J Be. a _ _ H J o. \\. pp— m «.2 o& X N;— oi 8 ‘ . Qd mp, vim N; Q;— . mg — _n /<\/<\ 90 d _ 2 IU m: MIUO .IU 44 64 Table VI. Temperatures of singlet maxima and magnetic susceptibility data T(°C) Xh Xh corr Complex CH3' CH3O (x103cgsu)a (x103cgsu)b Ni[(CH3)3P]3(CN)3 28 22 202 Ni[(CH3)3P]3(CN)3 -24 ~'0 251 Ni[(cna)2pocna]3(cu)2 25 16 24 289 Ni[CH3P(OCH3)2]3(CN)2 - 5 1 13 291 Ni[P(OCH3)3]3(CN)3 -17 3.4 296 {Ni[P(OCH3)3]5}(C104)3 >77 32 531 [Ni[CH3P(0CH3)2]5](C104)2 >72 >72 24 500 {Rh[P(ocns)3]5101o, 57 170 645 {Co[P(OCH3)3]5]C104C ~10d 1549 2016 aMagnetic susceptibility at the corresponding singlet maxi- mum temperature. For complexes with both methyl and methoxy groups, a temperature intermediate between the two singlet temperatures was used. bMagnetic susceptibility corrected for diamagnetism. See reference 70. cTrace paramagnetic Co(II) impurities are present in this complex. dOnly a very broad singlet is formed for this complex. 65 cannot be due to an averaging of cOupling constants of op- posite signs because at the singlet temperatures the con- centration of free ligand is very small and moreover, JPOCH has the same sign in both free and complexed forms. Para- magnetic relaxation effectsa5'57 similarly cannot be re- sponsible for the decoupling since the magnetic suscepti- bilities of all of these complexes in solution at the tem— peratures where the singlets are at a maximum intensity have been measured and are essentially diamagnetic. The magnetic susceptibilities were measured by Evans' nmr method47 and the corrected susceptibilities for the four- and five—coord- inate Ni(CN)2 complexes are recorded in Table VI. Some of these corrected susceptibilities are large, but they are within the range of diamagnetic behavior.68 Therefore, by analogy with the Pd(II) and Rh(I) com- plexes, the decoupling mechanism must be similar to that proposed by Fackler, 22.21.55r59 In an elegant fashion, theSe authors showed that the phosphorus-hydrogen decoupling could arise from Spin exchange of the strongly coupled phosphorus atoms under conditions of chemical exchange. The similarities between the Pd(II) and Rh(I) systems and the complexes studied in this work are Obvious, for indeed, strong 31P-31P coupling across the metal ion is apparent from the virtually coupled Spectra and,in addition,chemical exchange occurrs. Some more qualitative comments may be made, however, concerning the decoupling mechanism. In the slow exchange 66 limit the exchange time, Te, is very long and Te >> 1/JPP,. Hence, the virtually coupled spectra are observed for these complexes whose features are dependent upon J and PP" JPH' JPH" As the temperature is increased and the exchange time decreases, Te approaches 1/J and the large J PP' constant becomes a decoupling mechanism for J PP' . The PH rapid intermolecular chemical exchange of one ligand has the effect of averaging the spin states, at the site of phosphorus atom exchange, that the remaining phosphorus atoms (momentarily bonded to the metal) "see". This spin flipping causes the spin lattice relaxation rate of the re- maining phosphorus atoms to increase which in turn, causes the PéH coupling to be averaged out. It is apparent then that a large J value is necessary for this mechanism. PP' PP' were very small, Te would exceed l/JPP' at rela- tively slow exchange rates and the induced relaxation rates If J sufficient for decoupling would never be realized. If the rate of exchange is increased beyond the point where the singlet appears, eventually Te > 1/JPP, and the "resonant" ligand does not “see" the exchanging ligand. Hence the de- coupling mechanism is lost, the spectrum appears as if there were no P-P coupling or at least very weak coupling, and the fast exchange spectrum is a Simple doublet of separation JPH which is a weighted average of complexed and uncomplexed ligand in solution (yigg_§gp£§). It is interesting to note that in the complexes of (CH3)2P0CH3 and CH3P(OCH3)2 the singlet maxima for the methyl 67 and the methoxy protons occur at different temperatures for each ligand (see Table‘vfl. This difference is understand- able . . f . . . in View 0 the above discuSSion Since JPCH and JPOCH for each ligand have different values. Thus, since the frequency of P-H coupling is different, the two types of protons in each ligand will reach their maximum relaxation rate at different temperatures. .ML5n+ Complexes.- The proton nmr spectra of these complexes are also temperature dependent and the general features of the Spectra are Similar to the previous two classes of complexes discussed. Figure 16 shows the Spectrum for the rhodium complex, (Rh[P(OCH3)3]5}+ at the temperatures indicated. Again, a high temperature doublet, an intermedi- ate temperature singlet, and a low temperature multiplet are observed; however, in this case the multiplet is a well resolved six line resonance. The corresponding Ni(II) and Co(I) complexes of P(0CH3)3 show very similar Spectra: how- ever, the nickel complex decomposes at approximately 90°, when the spectrum is a broad singlet: therefore, no high temperature doublet is observed. In addition, the Co(I) complex exhibited only a very broad singlet intermediate between the doublet and sextet. This may be due to the trace paramagnetic Co(II) impurities which were identified by esr and which may be present from unreacted starting material. Also, like the Ni(CN)2 complex of CH3P(OCH3)3, the complex, {NilCH3P(OCH3)3]5}3+, exhibits two resonances for the two types of protons, and again only the OCH3 protons exhibit 68 Figure 16. 1H nmr SpeCtrum Of [Rh[P(OCH3)3]5}B(C6H5)4 in CH2C12 at various temperatures. l 57° A2 2° ”:6. IL I T lr fi‘ P Pm 3.80 3.60 3.40 Figure 16. 70 the low temperature splitting pattern while the CH3 reso- nance is a broad singlet. The doublets and singlets in these Spectra have the same interpretation as before--they correspond to fast and intermediate exchange rates, respectively. Absorption spectra studies reported herein on the Ni(II) complexes demonstrated their instability in solution with reSpect to free ligand and a lower coordinated species. Similar ex- periments on the rhodium and cdbalt complexes in this work and elsewhere38 have shown their dissociative instability and we have observed isosbestic points in the absorption spectra as the temperature is varied. For the rhodium com- plex the isosbestic points occur at 280 mu and 336 mu and for the cdbalt complex at 341 mu and 402 mu. Furthermore, the correSponding four-coordinate complexes of rhodium, [RhL4]+ can be isolated for L = P(0CH3)3 and other similar ligands.38 In the case of Ni(II) and Co(I), these lower coordinated Species have not been isolated. Therefore, an equilibrium of the form, 11+ n+ ( 6) M115 < > ML4 + L is present in solutions of all of these complexes. In the case where M = Rh, this equilibrium has already been pro- posed in order to account for the ligand equivalence indi- cated by the singlet in the 1H nmr spectrum at 38° for {Rh[P(OCH3)3]5]B(C5H5)4.3° This observation is consistent with the work reported here since a singlet was observed in 71 proton nmr spectrum at approximately 38° (see Figure 16). Therefore, under the conditions of the above equilibrium, an intermolecular process of ligand exchange is occurring in solution where the singlet recorded at intermediate tem- peratures represents an intermediate rate of ligand exchange and phosphorus-hydrogen coupling is averaged out due to Spin exchange similar to the Ni(CN); complexes. The symmetrical six-line multiplet which is nearly a 1:5:10:10:5:1 Spectrum represents the strong ligand-ligand coupling through the metal in these trigonal bipyramidal complexes. It may be considered as an AA'A"A"'A“"Xan'Xn" Xh"'xn"" type of Spectrum in the limiting case where A and X again have notations similar to before. The separation of the two outer lines in the sextet observed for {Ni[P(OCH3)3]5}2+ is about 11.8 Hz, a value very similar to the doublet separation at high temperature, and it repre- sents the sum J + J + J + J A-X Ax + JA'X A"X A'“ x A""x‘ coupling through the metal (five bonds) should be very small or almost negligible, thus the separation of the two outer lines of the sextet Should be approximately JAX. It is commonly observed54.58 that J for P(0CH3)3 (10.5 Hz) POCH . changes very little upon coordination, hence the value of 11.8 Hz for the separation of the two outer lines is con- sistent with this type of spin system. Also, the fractional intensity of the two outer lines is predicted to be (1/2)r-1 (r = number of A nuclei) that of the entire resonance."o For these complexes r = 5, hence, the intensity of the two 72 outer lines is calculated to be 1/16 that of the sextet-— close to what is actually calculated from the spectrum, and what would be expected from the above binomial expansion. The symmetrical nature of the six-line proton nmr pat- tern for the ML5n+ complexes is interesting since there are two chemically different ligands in the trigonal bipyramidal geometry; the two axial ligands and the three equatorial ligands. (Other five-coordinate geometries would also re- sult in at least two chemically different ligands.) The 1H resonance is symmetrical (even at 100 MHz) and indicative of only one type of ligand and may suggest that the ligands are rapidly exchanging sites intramolecularly, whereby all lig- ands would be equivalent. This process could be similar to the Berry71 mechanism postulated for pentavalent phOSphorus compounds and M(PF3)5 complexes (M = Fe, Rh, Os). The multiplet pattern is broad at -90°: however, this is prob- ably due to solvent effects. At the present time, low tem- perature spectra in other solvents have not been investigated. The 31F spectra were recorded for some of these complexes: however, the resonances are very broad and poorly resolved and hence, no information can be obtained from them. Rates of gigand Exchange.- If it is assumed that there is present a similar ligand exchange rate where the NiL3(CN)3 complexes all exhibit the singlet maximum, then the temperature at this maximum singlet (Table VI) should provide a measure of the relative exchange rates. For ex- ample, Ni[(CH3)3P]3(CN)3 exhibits a singlet maximum,or a 73 similar rate of exchange.at a lower temperature than Ni[(CH3)3POCH3]3(CN)2, therefore, (CH3)3P should exchange more rapidly at a given temperature. A comparison of all these “singlet" temperatures gives the following relative exchange rates: (CH3)3P > P(0CH3)3 > CH3P(OCH3)3 > (CH3)3POCH3. This ordering of these four ligands has been observed before in the instability constants (K ) for inst the same NiL3(CN)2 complexes (p. 33) where the (CH3)3P complex has the largest K. The correlation of ligand inst' exchange rate and Kin suggests that the equilibrium rate st is primarily determined by only one of the reactions, either the forward or reverse (Equation 3). Also, the order of ligand field strength as noted by the position of the d-d transitions, is approximately the same as the above with again the (CH3)3P complex having the lowest ligand field strength (p. 33). In a similar manner a qualitative exchange rate order can be obtained for the P(0CH3)3 complexes of Co+, Rh+, and Ni2+ (see Table\n3. The order of exchange rate is: Co+ > Rh+ > Niz+. This order might be expected since Ni2+ with the highest nuclear charge should be the least labile. In turn, Rh+ should exhibit greater overlap of its 4d or- bitals with the ligand orbitals than does cobalt's 3d orbitals. PART II REACTIONS OF SOME CYANOPHOSPHINES WITH DIBORANE AND BORON TRIFDUORIDE 74 INTRODUCTION Recently, a rapid and simple method of preparation of several cyanophosphine compounds, R3PCN (R = CH3, OCH3, OC2H5. C3H5, N(CH3)2), was reported.72 The compounds [(CH3)2N]2PCN73 and (C2H50)3PCN74 had been reported pre- viously, but their preparation involved rather extensive reaction times. Cyanophosphines have been found to exhibit unusual coordinating properties. Reactions of these RaPCN compounds with M(C0)4C7H8 (M = Cr or Mo and C7H3 = norborna- diene) produce dimeric complexes of the type [M(CO)4L]2. These latter complexes have been assigned the following bridged structure (Figure 17) on the basis of infrared and proton nmr data. 12 (c0) ’\ I“ :N‘M(CO) “~°FECF1V' 4 Figure 17. Structure of LM(CO)4L]3 complexes (M = Cr or Mo; R = CH3, oczns. C6H5. N(CH3)2).75 Nixon and Swain prOposed the same structure for the complex [Mo(CO)4(CF3)2PCN]2.76 Thus, these bridged complexes have demonstrated that two potential donor sites are present on these cyanophosphines. 75 76 In order to further investigate this dual basicity, the reactions of these cyanophosphines (R = CH3. CH30, C3H50, C3H5. N(CH3)2) with the Lewis acids, borane and boron trifluoride were investigated. Similar reactions of RP(CN)2 (R = CH3. C6H5) were also investigated. Borane and boron trifluoride were of particular interest since they have shown markedly different affinities for phosphorus and nitrogen base sites in adduct formation. Toward BF3. the donor strength is in the order N > P, while toward 8H3, the order is P > N.77'79 For example, in 1:1 adducts of (CH3)3NPF3 with BH3 and BF3. the borane group is bound to the phosphorus atom while BF3 is bound to the nitrogen atom.8° Many authors""""81 have explained the increased af- finity of borane for phosphorus to be the result of overlap between a BH3 molecular orbital of 0 symmetry with a vacant 3dw (or 4prr)“"83 orbital on the phosphorus atom. Thus, there is a dv-pw (or pw—pn) component enhancing the strength of the PéB oébond. In addition, this drift of electron density lowers the energy of the molecule, since it partially neutralizes a charge separation of the type X3P+ - 8H3.79 For the BF3 group, the high electronegativity of the fluorine atoms does not allow a similar drift of electron density.79 This "hyperconjugation" argument has been used to explain the existence of the compounds OCBH384 and F3PBH3,35 and the nonexistence of OCBF3 and F3PBF3.79:°° "Back bonding" of this type is similar in kind to that which 77 occurs between a phosphorus ligand and a transition metal atom.1 Indeed, for the [M(CO)4L]3 complexes (Figure 17), Jones and Coskran75 reported the.M-N bonds are broken in the reaction [M.(C0),L]2 + 2L ——> 2M(co),L3 (7) rather than the M-P bonds. However, the hyperconjugative model of P-B bonds has not been consistent with other ex- perimental results. For example, the model would predict that 3H3 would form a more stable adduct with PF3 than with PFaH. because of enhanced v-acceptor character of PF3 over PFzH,87 but HFzPBH3 is more stable toward dissociation than F3PBH3.°° Rudolph and Parry have proposed another P—B bond models.8 The Lewis acidébase interaction energy (E) can be approxi— mated by E = F[p+(Fa/2)1 (8) where F = Lewis acid field strength, [)2 dipole moment of the lone pair, and a = polarizability of the lone pair. Borane bond strength reversal (P > N) has been shown to be a function of the Lewis acid strength,89 and is consistent with this model. The greater stability of HFzPBH3 over F3PBH3 toward dissociation has been attributed to the larger polarizability, a, of the lone pair on PF2H than that on PF3.87 78 It was of great interest to determine whether the ob- served basicity toward BF3 (N > P) and BH3 (P > N) would also be observed in the cyanophosphines of this study. Any spectroscopic data obtained for such adducts might also help to discriminate between the two P-B bond hypotheses. Furthermore, since adducts of the type, RsPBH3 (R = CH3.9° 01130.91 C2H50,91 03115.92 N(CH3)3°3), have been pre- pared previously, as well as the R3PBF3 (R = CH3.9‘ C3H595) adducts, a comparison of the experimental results and spec- troscopic data for the corresponding [R3PCNlBH3 and [RaPCN]BF3 adducts that are formed in this study helped to determine what effect the cyanide group has on the o-donor Character of the phosphorus atom to which it is bound. EXPERIMENTAL All work was carried out by standard vacuum line techniques or under inert atmosphere. Materials.- Boron trifluoride was obtained from Matheson Scientific Inc., Chicago, Illinois and purified by passing it through a -145°C trap. Diborane was prepared by the action of H3804 on NaBH4.96 The cyanophosphines RgPCN (R = CH3. c330, C3H5O. cana. N(CH3)2) and RP(CN)3 (R = CH3, C3H5) were prepared according to a method pre- viously described.72 Methylene chloride was dried over CaH, before use. Preparation of [RzPCN]BH3 Adducts.— An excess of BgHe was condensed from a calibrated trap into a flask which contained the R2PCN ligand (R = CH3, CH3O, C2H50. C5H5. N(CH3)2). The mixture was stirred and allowed to warm from -196° to room temperature. When the uptake of B3H6 was complete («'1 hr), as measured by an attached manometer, the excess B3H5 was condensed into the calibrated trap. .A molar ratio of approximately 1:1 was obtained for BH3:R3PCNO(Tab1e\HIL (For R = C5H5 and CH3, the value of BHazRgPCN was determined when these RaPCN ligands were dis- solved in methylene chloride solutions (see BF3 reactions). 79 80 J Jl-..‘ .umwc can conuomomn . «acnmo ca can coauomo¢w mefinmmv_«fizoooomooa no.” 1-- oa.n -1- mo.m so.“ 1-- 6H.n -1- oH.n muxnnmv_axzovoemu_ oo.o mo.o H¢.n oe.m on.n mxmmmvnmm_zomnxnmoov_ oo.H Ho.o me.n Ho.“ en.n mxnmnvemm_zomaxonmov_ no.o mo.o eo.u on.H nm.u mAemmvnmmHzoouxemUVH oo.o nun oo.v In: ne.v nemmxzoouxnmeoVH no.” 1-- ov.n nu- am.n mnnm_zomuxonmov_ oo.o -u- vo.n -1- oH.e nemmflzooexnmov_ nu- oo.fi In- on.m nn.v nnmmmzom«~z«xemov_w -u- oo.o -1- on.n oo.o nemn_zom«xameov_ 1-- mo.o -1- um.H no.m nemm_zomuxoemnov_ -11 Ho.o -1- Hm.H sw.n nemmzzomaxoemoVH 1-- no.o -11 en.n be.» memmflzooexnmovn. ocwpmnonm oaoe ocasmmonm maoe ms cmxmu mo coxmu ocflsmmonm .lnwm oaoa .lmmm oaoe ohm mo Hoes omam mo Hose mo HOEE ocoomaoo L .coquEHom uoooom «mm ocm nmm Mom memo Hmucoefluomxm .HE/manme 81 Since these adducts were solids, efficient stirring and complete reaction were not possible when the reactions were run neat.) White solids were formed when R = CBHa or CH3, and these solids were recrystallized from a CH2C12-hexane mixture. These complexes decomposed in a matter of hours at room temperature. When R = CH30, N(CH3)2, and C2H5O. light yellow liquids were obtained which contained small amounts of decomposition products, and also decomposed in a matter of hours at room temperature. Preparation of [RzPCNlBFa Adducts.- Boron trifluoride ( was condensed into a -196° flask which contained the RaPCN ligand (R - CH3, c330, cans) dissolved in cnzclz. This mixture was allowed to warm to -78°, until the BFs uptake was complete (~45 min; when R' CH3O. the reaction was stop- ped when the uptake slowed markedly). Excess BF3 was con- densed through a -145° trap into the -1960 trap, in order to remove CH3C12. A mole ratio of approximately 1:1 was Obtained for BF3:R2PCN (Table VII). The adducts formed clear solutions in CH2C12 and were unstable above -20° toward loss of BF3. Preparation of [RzPCN]BF3(BH3) Adducts.- The RzPCN(BH3) adducts (R s CH3, CH30, cans) were formed, in. situ, and allowed to react with BF3 in the manner described above. This resulted in a molar ratio of approximately 1:1:1 for BH3:BF3:R2PCN (Table VII). These complexes formed clear solutions in CH2Clé and were unstable with respect to loss of BF3 above -20°. 82 Reaction of RP(CN)2 with BF3.- The RP(CN)2 ligands (R = CH3, C6H5) were allowed to react with BF3 in the manner described above. A molar ratio of approximately 2:1 was obtained for BF3:RKCNL (Table VII). These compounds formed insoluble white solids and were also unstable with respect to loss of BF3 above -20°. Nmr Spectra.- Room temperature 118 spectra were re— corded on a modified NMR Specialties Model MP-1000 Spectrom- eter at 60 MHz. Low temperature 113 spectra were recorded on a Varian Associates Model HA-100 spectrometer at 32.1 MHz. Phosphorus-31 spectra were recorded on a Varian Associates Model DA-60 spectrometer at 24.3 MHz. Proton and 19F spectra were recorded on a Varian Associates Model A56/60 spectrom- eter at 60.0 MHz and 56.0 MHz, respectively. A Fabri-Tek Instrument Co. Model 1082 computer was used to time-average the proton spectra. The 118, 31P, 19F, and 1H chemical shifts are relative to external B(OCH3)3, 85% H3P04. CC13F. and TMS. respectively. Spectra were obtained for neat samples of [RzPCNlBH3 (R = CH30, cznso, N(CH3)2), and methylene chloride solutions of all other adducts. Infrared Spectra.- All spectra were recorded on a Perkin-Elmer Model 225 grating Spectrophotometer in methylene chloride solutions. For boron trifluoride adducts, a low temperature solution cell cooled to -78° was used. RESUDTS The reaction of RaPCN (R = 03,, 0330. canao. cans. N(CH3)3) with Bane resulted in 1:1 borane adducts. First order nuclear magnetic resonance spectra (113, 31F, 1H) were obtained for these adducts; the spectra (Figure 18) of [(CH3)3PCN]BH3 are typical. Coupling constants and chemical shifts Obtained from the Spectra of these compounds are recorded in Table VIII, along with the cyanide stretch- in the 113 ing frequencies (v The observation of J CN)° PB spectra for all of these complexes is indicative of adduct formation on the phosphorus base site, rather than on the nitrogen. A second borane molecule was not taken up by these ligands, even when the reactions involved lower tem- peratures (-78°) and longer reaction times (24 hrs). The reaction of RaPCN (R = CBS. onso, can,) with BF3 also resulted in 1:1 adducts. Nmr (118. 19F. 31P, 1H) and ir data (VCN) obtained from the spectra of these compounds are reported in Table IX. Typically, the 113. 31F, and 19F spectra are broad singlets, and therefore, only chemical shifts are reported. However, in the 1H Spectra splitting patterns are observed for the R protons (R = CH3, cnao) and J and J values are reported along with the chemi- PCH POCH cal shift data. All nmr spectra were obtained at 83 84 Figure 18. Nmr Spectra of [(CH3)2PCN]BH3. (a) Time - averaged 1H spectrum at 60.0 MHz. (b) 1H spectrum at 60.0 MHz. (c) 11B spectrum at 60.0 MHz. (d) 31P spectrum at 24.3 MHz. 85 1 1 A I 1 l "'1 __I_ “JR ) 61H(3H3) I‘m '15:, E ] (SI '3 I400Hz '1 d - a» L- |, 5319 Figure 18. 86 .mcououm mamamnuoa mo mcaamsoo .mwmmnucwumm ca co>fim mum mooam> ocmmfia combo .mcouonm Hanuoe mo umazm Hmowfimnoo o .mcououm ocoamnuoa mo amaze Hmoweonon .mem unwound mo umwnm H60H80£Um Amoaavmmsm as so we n.oH ¢.on on . no.o vs.m «Anmovz . onn.H Aoofiuvsofim mo Hoe -1- oo o m on no u on o aHm.¢ oamao Acmemvnofiu co Hoe we «.mn s.mn see- no.o nm.n oamo Anaemvsoaa on Hos he u.oH n.nn o oe.o Hs.H emu oxnsflavooau no: so 1-- -1- o.mn 6H 1 -1- one.» ammo .Asmv 1 5V :5 :5 Es M wages as: sag Es ace 7 mm mm and m . n n m 20 o o n AEnoch mass mans A mmvmao Amvmao > .muuooom mmmnzumnma How mumo He too HEZ . HHH> THQMB 87 .momosucoumm ca mum muosoom Anmmvzomam mo mosam>n .mmmmnucmumm ca mum mosam> osmofia momma Sam «.3 no Tao 1mm 82: 37 013.886 AnmmVammEonaAosmo: scan a. 3 an o. so H. mm a. o3 2.. AA 2.. Sum. H “£3 amazon: #8: some ...i .i .i- n. on o. «3 & onto? 1% on. s v on. .c. ”an Eon: :16: 3am see .i .1- mom on: «.32 van? ~35“;st aeflzomaxoamo: moms n. v .... u..- oém he: mxmoionl. 6:158. H «mm Sum: «so: fusov AEeaooae Ammv Assoc Ede Assay 38v Assoc ounces 20> Amvmomo b Anmmvmuno Anmmvmuuo hose mane Amvmuo .mnosoom AnmmvnmmHZOmsmH osm nmm_zomsm_ now some He can nsz .xH manna 88 low temperatures («v-40°), since these adducts are not stable above -20°. Inasmuch as JPB is not observed in the 118 spectra, it is proposed that the BF: molecule is bound to the nitrogen atom of the cyanide group. A second BF3 molecule was not taken up when R 3 CH3 or C3H5, even with longer reaction times (~t24 hrs). Reaction of RgPCN (R = C3H50, N(CH3)2), or further reaction of (CH30)3PCN, with BF3 resulted in the decomposition of the ligand unit. Boron trifluoride was taken up in greater than E 2:1 amounts (BF3:RaPCN), which is prObably due to the reac- 1,- tion of BF3 with oxygen and nitrogen atoms of the substituents. At room temperature, BF3 was given off, and the proton nmr spectra indicated the R2PCN ligands were not present. Table Ix also contains similar nmr and ir data for the [R3PCN]BF3(BI-13) adducts (R = CH3, cnao), which are formed in the reaction of BF3 with [RgPCNlBH3. When R - C3H5, the adduct that is formed is insoluble and no Spectroscopic data were obtained. The 31F and 19F spectra consist of broad singlets, while in the 1H Spectra only methyl and methoxy resonances are Observed. Boron-11 spectra exhibit resonances due to the BF3 and 3H3 groups (Figure 19). It is apparent that the 8H3 group is bound to the phOSphorus atom, since JPB is Observed in the borane resonance. The nmr spectra for these latter two complexes were also Obtained at ~v-25°, since these adducts are not stable above -20°, and “thermal decoupling"97 induced by quadrupolar boron coupling was observed below this temperature, as evidenced by loss of 89 Figure 19. 113 nmr spectrum of [(CH30)2PCN]BF§ 8H3) at 32.1 MHz. ((a) Spectrum at -25°. b) Spec- trum at -50- . - 91 JPB in the 118 spectra (Figure 19).98 When R = C3350 and N(CH3)2, decomposition Occurred which was Similar to that observed in the corresponding 1:1 BF3 adducts. No reaction occurred when RP(CN)2 (R = C3H5. CH3) was allowed to mix with 3236 at low temperatures (-78°) and long times (~o24 hrs). The 2:1 adducts formed from the re- action of BF3 with RP(CN)3 (R = c311,, 033) were insoluble and unstable above -20°. No Spectroscopic data were Obtained for these compounds. DISCUSSION For convenience, the 1:1 borane adducts will be dis- cussed separately from the other adducts of this study. [RaPCNlBHgyAdducts 31P - 113 Coupling Constant, JPB.- In a discussion of the relative merits of the two P-B bonding models described previously, J is of particular importance, PB since it is a property of the bond in question. The sign of JP has been indicated to be positive33t99 in borane B adducts. Therefore, the adducts in TableVIII are listed in order of increasing JPB’ Rudolph and Shultz99 have reported an empirical cor— relation between the magnitude of JPB and dative bond strength for a series of smoothly varying phOSphine ligands such as F3 —anP —> 8H3. Me3_anP -—> 8H3, (MezN)3_nFnP —> BH3, and F3XP -9 8H3 (x = F, Cl, Br). These authors have also reported that J values of similar magnitude in borane PB adducts of phosphines which differ markedly may not reflect dative bond strength, but quite different JPB values reflect correct relative bond strengths. Therefore, the JPB values in this study (Table VIII) indicate the following decreasing order of P-B bond strength in [RaPCNlBH3: R = N(CH3)2 ~' C2H50 ~ CHaO > CH3 > C3115. 2 9 93 The borane hyperconjugation model is not consistent with this order of P-B bond strength. If the P-B bond is stabilized by "back bonding“, bond strength should depend on the ligand's w-acceptor ability.83 In studies of metal carbonyl complexes and their carbonyl stretching force constants, Jones and Coskran5‘ found the following order of decreasing w-acceptor ability: P(0CH3) >P(CH3)3 > P[N(CH3)3]3. If this same relative order is asSumed to be present in the corresponding cyanophosphine ligands, it is not reflected in the above P-B bond strength order, since {[(CH3)3N]PCN}BH3 has a larger J than [(CH3)2PCN]BH3. PB Cowley and Damasco83 have found an empirical correla- tion between JPB (P-B bond strength) and the basicity of the phOSphine. This is consistent with the model of Rudolph and Parry (Equation 8), since the lone pair polarizability. o. is related to the phosphorus basicity.87 The JPB values in this study reflect the same order of basicity as proposed by Cowley and Damasco83 for R3P (R a N(CH3)3, CH30, CH3) and RPH, (R = CH3, C3H5) ligands, and thus, they are also con- sistent with equation 8. It should be noted that these authors83 report the following order of decreasing JPB (ligand basicity): P(0CH3)3 ~'[(CH3)2N]3P > (CH3)3P. How- ever, on the basis of base displacement reactions, Reetz91 has reported the following basicities: (CH3)3P ~' [(CH3)3N13P. Foester and Cohn1°° have reported that base strengths toward borane decrease in the order CH3PF2 > (CH3)2NPF2 > CH30PF3, based on displacement reactions, 94 although J decreases in the order (CH3)2NPF2 > CH30PF3 > PB CHaPFa. The measurement of large basicities for these methylphosphines toward borane may be due to entropy effects“1 in the base displacement reactions. For example, dissociation constants of (CH3)3NB(CH3)3 and (CH3)3PB(CH3)3 suggest the phosphine is the stronger base, while enthalpies of formation of these adducts indicate the N-B bond is stronger than the P-B bond.101 Therefore, the order of basicities as proposed by Cowley and Damasco,83 and followed in the corresponding cyanophosphines of this study, may in- deed be correct, and thus be reflected in J . PB The relatively high base strength (and high JP in the B borane adducts) of (CH30)3PCN, (C3H50)2PCN, and [(CH3)2N]2PCN is probably due to (srd) wébonding between the filled 2p orbitals of oxygen and nitrogen and the empty 3d orbitals of phosphorus. The vébonding increases the electron density on the phosphorus atom, and thus, increases the basicity.“2 This type of (p-d) wébonding has been shown to be signifi- cant in structural studies of HzNPF31°3 and CH30PF2.1°‘ The nitrogen atom of HzNPFz is in a planar environment,”3 and the [(POC) of 123° in CH30PF2 suggests sp2 hybridiza- tion in oxygen orbitals.1°‘ 31 _1 - . P H (Substituent) Coupling Constants, JPng JPOCH' _ l - 63,63 63.64 JPNCH' The coupling constants JPCH' JPOCH’ and JPNCH64 in the free cyanophosphines are assumed to be posi- tive. There is an increase in the absolute value of all these couplings in their borane adducts (Table X). 95 Table x. Values of J , JP , and JPNCH 1n R2PCN and [RaPCN1B33 . JPCH' JPOCH or IJPCH I ' IJPOCHI J Sign: Of R .m .. W .. semi .. 130:: ”SE [RaPCN13H3 ' PNCH [RzPCN]BH3 CH3 4.5 10.2 _10 .2 E1 CH30 10.3 12 .2 +12 .2 ' C2350 7-5 9.0 + 9.0 1..- N(CH3 )2 10.0 10_5 ”0.5 These results can be interpreted in terms of the in- creased R-P-R and NC-PeR bond angles in the borane adducts as compared to the free ligands, which results in an in- crease in the s character of the phosphorus bonding orbi- tals ("isovalent hybridization hypothesis").1°5 Muller and Pritchard 105 have observed an increase in JCH as the amount of S character in the CH bond increases. Cowley and Damasco83 have proposed a similar relationship between JBH and the amount of 2s character in the B-H bond. Manatt. et al.,°3 have prOposed that J becomes more negative as PCH the amount of 3 character in the phosphorus bonding orbitals increases in a variety of trivalent phosphorus com- pounds. For example, JPCH in (CH3)3P is +2.66 Hz, where 1 ' . the Z.(CPC) of 98.6° 07 suggests a large amount of p orbi- tal contribution to the bonding orbitals, while in the 96 nearly pure Sp3 hybridized (CH3)4P+ and (CH3)3P=O. JPCH is -14.4 Hz and —13.4 Hz, respectively.63 Therefore, for [(CH3)3PCN]BH§ J must be negative (-10.2 Hz) in order to PCH become more negative upon adduct formation. This change in the Sign of JPCH for methylphosphines upon complexation has been observed previously55:61 (p. 58). However, for JPOCH and JPNCH similar Sign changes are not observed54 (p. 61), and only small changes in the magnitude of JPOCH and JPNCH are observed upon complexation.54'°‘ Therefore, these JPOCH and JPNCH values for the borane adducts are assigned posi- tive values. Thus, upon adduct formation, JPCH decreases while JPOCH and JPNCH increase. 118 - 1H Coupling Constant, JBH.- The sign of JB has H been assumed positive.108 Valuesgf.JBH Obtained in this study are not consistent with a simple isovalent hybridiza- tion hypothesis.“5 Electronegativities of the substituent groups decrease in the order CH30 o'C3H50 > N(CH3)3 > CBH5 > CH3,1°9 and the 5 character in the B-H bond should decrease in the same order.“5 From Table VIII, it can be seen that JBH of [(CH3)3PCNJBH3 is not consistent with this order. However, since the above hypothesis depends on an increase in the borane bonding angles ([.(HBH)),1°5 the bulkiness of the phosphine may have an effect on the order of J Tol- BH' man36 has measured the “cone angles" (Figure 20) of a num- ber of tertiary phOSphorus compounds, and the method used would predict the following order of increasing cone angle 97 for the RaPCN ligands: R = CH30 ~'C2H5O «'cas > cana. (The value of2282 is used as the average value of a Ni-P bond. However, the relative order of ligand cone angles should not change, if an average P-B bond distance is chosen instead.) Figure 20. The ligand cone angle for tertiary phosphorus compounds.36 Due to the two methyl groups in the N(CH3)2 group, the cone angle in the corresponding cyanophosphine would be expected to be larger than the CH30. C3H50, and CH3 ligand angles, but smaller than the ligand angle with the bulky C5H5 group. Therefore, the ability of the borane group to increase H-B-H bond angles, or increase 5 character in the B-H bond1°5 (JBH), should decrease in the following ligand order due to ligand steric requirements: R 8 CH3 c: C3H5O ~'CH3O > N(CH3)2 > CoH5. This order is consistent with the J values obtained for these borane adducts (Table BH VIII). The above argument considers that an increase in H-B-H bond angles is caused by o-bond electronegativity effects, although the magnitude of this increase is due to ligand steric requirements. However, the borane hyperconjugation model also allows for increased H-B-H bond angles, with 98 increased ligand v-acceptor ability.83 Although the order of JBH (Table VIII) does not strictly mirror the ligand 7- acceptor ability, since (C5H5)2PCN is a better w-acceptor than (CH3)2PCN,1°°o11° hyperconjugation can still be con- sistent with the above model. It assumes a negligible effect on the sterically determined, relative J order due BH to electron density drift in the B-H bonds. 31F - 1H Coupling Constant, J The observed PBH' ' values of JPBH (Table VIII) follow-the order of substituent electronegativity, 1:9,, OCH3 > N(CH3)2 > CH3."9 The iso- valent hybridization hypothesi31°5 would predict an increase in the 8 character of both the P-B and B-H bonds as substituent electronegativity increases. If the PéB-H bonding situation is analogous to P-C-H, then J H should PB decrease with increasing 8 character in the phosphorus o o as bonding orbitals, as does JPCH' In order for JPBH of these cyanophosphine adducts to decrease with increasing electro- negativity, J must be assigned a negative Sign. However, PBH this result is ambiguous, since it neglects the contribution of the ligand and borane adduct bond angles33'105 to the amount of 3 character in the P-B-H bonds. 31P Chemical Shift, 031P.- Table XI shows the change in 031P of these cyanophOSphines upon adduct formation. If only the molecular ground state electron distribution is considered, a downfield shift of 031P might be expected because of the o-donor character of these trivalent 99 3+ so? an? zoo «A Onmov 8+ co 1 ST zomaxoamaov 3+ on u no .. 2232265: on: 3 .. on + zomaxomoov «on o no + zomaxnmov .Aemmv . . Amcfinmmonnvmwno I Auoooom flammv . Aemmv . TQMHOQ “H00 N mfiflflq AUUHpHvUM QCMHOvafiQO .HATCfl—hmuogmvmflflQ ”Cwfignmmogm .nmmmzumnma ocm zomnm cw mumo umwnm HmowEOso monogamocm .HN magma 100 phOSphorus compounds.111 A small downfield shift may be attributed to the w-acceptor character of such ligands.111 There are two other contributions, however, to the chemical shift; a paramagnetic term, derived from electronic asym- metry around the nucleus,87 and the contribution from the other atoms in the molecule.112 This latter contribution is usually small and does not appreciably affect the chemi- cal shift.112 Phosphines with substituents that contain lone pairs next to the phosphorus atom have a large para- magnetic contribution to the 31P chemical shift.33o113 The paramagentic contribution is reflected by large negative values of 031P for these ligands, and an upfield shift upon ' adduct formation.33:113 The 31P chemical Shifts Observed in this study are consistent with these theories (Table IX), and are similar to the chemical shifts obtained in fluorOphosphine-borane83 adducts and phosphoryl compounds.113 In Chemical Shifts, 01H(BH3) and 01H(R).- Table XII shows the change in 01H of the substituents upon adduct formation. In general, 01H(R) shifts downfield, and indi- cates a primary dependence of 01H(R) on the molecular ground state electron distribution. The protons of the substi- tuents are slightly deshielded by a drift of electron density toward the borane unit. Borane proton chemical shifts generally move upfield as J increases (Table VIII). This is consistent with phos- PB phine basicity arguments, since the more basic phOSphine 101 will shield the borane protons to a greater extent and move 61H(BH3) upfield. Table XII. 01H(R) in RzPCN and [RZPCN]BH3. R 61H(R2PCN)1 51H([R2PCN]BH3) 4.08a 4.21a °2H5° 1.30b 1.33b can5 7.20C 7.63° N(CH3)2 2.75 2.74 aMethylene protons. bMethyl protons. CChemical shift of largest peak. Cyanide Stretching Frequency, vCN°— The cyanide stretch- ing frequency of these cyanophosphines increases upon adduct formation (Table VIII). An increase in this frequency has been observed in compounds of the type RCN, when boron com- poundsn-fi'118 or transition metals5°o113I117 coordinate to the nitrogen atom. It has been suggested,115o117 that the CN bond in the free nitriles is intermediate between a double and triple bond, due to a mesomerism of the following type: R-CEN(III) <—-9 R-C:N-(IV). Once the nitrogen atom becomes coordinated, such a phenomenon is not possible be— cause of the different geometries and hybrid orbitals involved (Figure 21). An X-ray structural study has shown 102 structure V is present in CH3CN-BF3.118 Therefore, the CN bond is a triple bond in the coordinated complexes and is indicated by an increase in v . CN + O. R-CEN—8F3 R-C=N\ - 5P sp2 BF3 Y EL Figure 21. Possible structures of [RCN-lBF3.51 If in the cyanophOSphines of this study, a mesomerism + _ of the form RaPCEN(VII) <-> RaPc=N (VIII) can be drawn, then adduct formation on the phosphorus atom should cause a drift of electron density away from the cyanide group, which makes structure VIII less favorable, and an increase in VCN should be observed. [RzPCN]BF3 and [RzPCNlBF3(BH3) Adducts 31P - 11B Coupling Constant, JPB'- The 11B nuclear magnetic resonance spectra of [R2CN]BF3(BH3) (R = CH3, CH30) indicate that the borane unit is bound to the phosphorus atom, since JPB is observed in the borane resonance (Figure 19). Thus, the BF3 molecule is bound to the other base site, the nitrogen atom of the cyanide group. In these complexes. JPB is slightly smaller in magnitude than that observed for the corresponding [RzPCNlBH3 adducts (Table VIII). The smaller values are consistent with the correlation of JPB 103 with phosphine basicity, since the electronegative BF3 group should reduce the basicity of the phOSphorus atom and de— crease JPB' The borane hyperconjugation model is not con- sistent with this observation, since the BF3 group would be expected to increase the w-acceptor ability of the phos- phorus atom, and lead to an increase in JPB' 31 _1 - - _ P H (Substituent) Coupling Constant, JEQE and JPOCH' The 31P - 1H coupling constants in [RgPCN]BF3(BH3) (R = CH3. CHSO) adducts (Table IX) do not change from those observed in the RzPCN(BH3) adducts (Table VIII). A BF3 molecule on the nitrogen atom of the cyanide group should not cause a large change in the S character of the phosphorus bonding orbitals and J and JP should not change substantially. PCH OCH Likewise, in the [RgPCNlBFs (R = CH3. CH30) adducts JPCH and JPOCH do not change from those observed in the free ligands (Table X), which again indicates the BF3 molecule is bound to the nitrogen atom. 31F Chemical Shift, 03:3.- The phOSphorus chemical shift in the [RZPCNlBFa (R = CH3, CH30, cans) adducts moves slightly downfield in all cases. If the BF3 molecule was bound to the phosphorus atom, large changes in 031P might be eXpected, similar to those obtained in the [RzPCN]BH3 adducts. The consistently small downfield shift is indicative of a small deshielding of the phOSphorus nucleus111 caused by the electronegative BF3 molecule on the nitrogen atom. Further evidence for a N-BF3 bond is given by the similar 104 effect of BF3 on 031P of the free ligand and borane adduct (Table XIII). Since the BFg molecule is bound to nitrogen in theIRzPCNiBFg(BH§) adducts it is probably bound in a similar manner in [RgPCNlBFa adducts. 1H(Substituent) Chemical Shift, 01H.- The proton chemical shifts of the substituent groups in the [RaPCN]BF3 (R = CH3. CH30, CoHa) adducts (Table IX) are moved down- field from the corresponding free ligand values (Table XII). These protons are deshielded by the drift of electron density toward the electronegative BFa group. In the [RaPCNlBF3(BH3) (R - CH3, CH30) adducts, the borane unit enhances the elec- tron density drift away from the substituents and 01H(R) is moved further downfield (Table IX). The effect of BF, ad— duct formation on 01H of the substituents in both free ligands and borane adducts is similar (Table IX), which again indicates the coordination of BF3 to the nitrogen atom. 19F Chemical Shift, 619F.- The fluorine chemical Shifts of the [RaPCN]BF3(BH3) (R = CH3, CH30) adducts are at lower field than the corresponding [RzPCN]BF3 adducts (Table IX), since the shift of electron density toward the borane unit deshields the fluorine nuclei. 11B Chemical Shifg, 0¥1§,- The boron chemical shifts of the SP3 units in [RaPCN]BF3(BH3) (R = CH3, cnao) are at higher field than in the corresponding [RzPCNlBF3 adducts. The shifts are in accord with the fluorine chemical shifts 105 o . ossu Ros: nmmmzomaflonmoV. oH mos- «as- zooaxosmov ms mu 1 o nmm_zomsxnmov_ as on + so + zoasxsmov Auosoom «savanna n uosoom ocmuog n Auoooom common um maesmmosmvmsno Aposoom mmvmsno so onecmnosmvmsno ossomsoo n mnno< .coquEHom poooom «mm com: nmmnzomnma can zumam ca memo cw omcmsu .HHHN magma 106 that indicate a drift of electron density away from the fluorine atoms and onto the boron atom, which increases the shielding of the boron nuclei. Boron chemical shifts in the borane units follow the same trend, indicating a drift of electron density from the borane hydrogen atoms, and increased shielding of the boron nuclei in [RZPCNlBF3(BH3) as compared to the corresponding [RaPCNlBH3 adducts. Cyanide Stretching Frequency, VCN.- The cyanide stretching frequencies of the [RzPCNiBFa adducts (Table IX) agree with the assignment of an N-B bond. As shown pre- viously (p. 102), the CN bond in the free cyanophosphines is intermediate between a triple and double bond. Upon adduct formation on the nitrogen atom, the CN bond has more triple bond character (structure VII). Thus, the CN stretch- ing frequency should increase, and to a greater extent than in the borane adducts, where the mesomerism is only partially shifted. The cyanide stretching frequency in the [RzPCNlBF3(BH3) (R = CH3. c330) adducts is similar to that observed in the corresponding [RzPCN]BF3 adducts, since the maximum bond order is attained in both cases. CONCLUSION Spectroscopic results have indicated that the 8H3 molecule is bound to the phosphorus atom and the BF3 mole— cule is bound to the nitrogen atom of the cyanide group in all the adducts of this study. The formation of 2:1 adducts (BF3:RP(CN)2) with RP(CN)3 (R = CH3, c635) ligands supports this asSignment. ' In general, the borane hyperconjugation model has been shown to be inconsistent with the nmr data obtained for the borane adducts in this study. A correlation between JPB (P-B bond strength) and phosphine basicity has been observed. This is not to say that borane hyperconjugation does not occur in these adducts, only that this effect does not determine the observed values in the nmr data. The JPB values of the [RaPCN]BH3 adducts (Table VIII) are consistently lower than those found in the borane adducts of the corresponding PR383I113 compounds, and indicate that replacement of R by a CN group decreases the basicity of the phosphorus atom. Experimental results also support this hypothesis, since borane adducts could not be formed with RP(CN)2 (R = CH3, cans) although [RaPCNlBH3 adducts are readily prepared. Furthermore, boron trifluoride does not coordinate to the phosphorus atom of R2PCN (R = CH3. CH3O). 107 108 although BF: does coordinate to the corresponding PR3 com- pounds.94I°5 BIBLIOGRAPHY 10. 11. 12. 13. 14. 15. 16. 17. BIBLIOGRAPHY G. Booth, Advan. Inorg. Chem. gadiochem., 6, 1 (1964). M, O. Workman G. Dyer, and Devon‘W. Meek, Inorg. Chem, 6, 1543 (1967). M. J Norgett, J. H. M. Thornle and L. M. Venanzi, Coord. Chem. ReV., 2, 99 (1967) . Luigi Sacconi, Pure Appl. Chem" 11, 95 (1968). C. Furlani, Coord. Chem. ReV., 3, 141 (1968). L M. Venanzi, Angew. Chem., Internat. Edn. (in English), 3, 453 (1964). L. P. Haugen and R. Eisenberg, Inorg, Chem., 8,1072 (1969). B. Bosnich, R. S. Nyholm, P. J. Pauling, and.M. L. Tobe, J. Amer. Chem. Soc., 90, 4741 (1968). G. A. .Mair, H. M. Powell, and L. M. Venanzi, Proc. Chem. Soc., 170 (1961). L. Sacconi, P. Nannelli, N. Nardi, and V. Campigli, Inorg. Chem., 4, 943 (1965). J. Chatt and B. L. Shaw, J. Chem. Soc., 1718 (1960). R. G. Hayter, Inorg. Chem., 2, 932 (1963). J. W. Collier, F. G. Mann, D. G. Watson, and H. R. Watson, J. Chem. Soc., 1803 (1964). D. A. Allen, F. G. Mann, and J. T. Millar, Chem. and Ind., 2096 (1966). E. C. Alyea and D. W..Meek, J. Amer. Chem. Soc., 91, 5761 (1969). Otto Dahl, Acta Chem. Scand.,~23, 2342 (1969). G. N. Schrauzer and P. Glockner, Chem. Ber., 91, 2451 (1964). 109 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 110 K. J. Coskran, J. M, Jenkins, and J. G. Verkade, J. Amer. Chem. Soc., 99, 5437 (1968). B. B. Chastain, E. A. Rick,-R. L. Pruett, and H. B. Gray, J. Amer. Chem. Soc., 99, 3994 (1968). R. Ri o, C. Pecile, and A. Turco, gnorg. Chem., 9, 1636 (1967). K. J. Coskran, T. J. Huttemann, and J. G. Verkade, Advances in Chemistry Series, No. 62, American Chemical Society, Washington, D.C., 1967, p. 590. K. N. Raymond and F. Basolo, gnorg. Chem., 9, 949 (1966). J. S. Coleman, H. Petersen, Jr., and R. A. Penneman, gnorg, Chem., 2, 135 (1965). ' B. B. Chastain, D. W. Meek, E. Billig, J. E. Hix, Jr., and H. B. Gray, gnorg. Chem., 1, 2412 (1968). J. Zemann, Z. Anorg. Allg, Chem., 324, 241 (1963). J. K. Stalick and J. A. Ibers, gnorg. Chem., 9, 1084 (1969). J. K. Stalick and J. A. Ibers, Inorg. Chem., 9, 1090 (1969). J. A. Betrand and D. L. Plymale, Inorg. Chem., 9, 879 (1966). E. J. Riedel and R. A. Jacobson, Nucl. Sci. Abst., 99, 27936 (1967). K. N. Raymond, P. W. R. Cornfield, and J. A. Ibers, gnorg. Chem., Z, 1362 (1968). L. Sacconi, J. Chem. Soc., 9, 248 (1970). M. J. Norgett, J. H. M. Thornley, and L. M. Venanzi, J. Chem. Soc.,.g, 540 (1967). J. W; Dawson, H. B. Gray, J. E. Hix, Jr., J. R. Preer, and L. M. Venanzi, J. Amer. Chem._§9c., 99, 2979 (1972). J. R. Preer and H. B. Gray, J. Amer. Chem. Soc., 99, 7306 (1969). P. Rigo, G. Guastalla, and A. Turco, ;norg3 Chem., 9, 375 (1969). C. A. Tolman, J. Amer. Chem.§gc., 99, 2956 (1970). 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 111 K. A. Jensen and Otto Dahl, Acta Chem. Scand., 22, 1044 (1968) . 1...... ““ Linda.M. Haines, gnorg, Chem., 12, 1685 (1971). w; C. Fernelius and J. J. Barbage, gnorg. Syn., 2, 227 (1946). "’ K. A. Jensen, P. H. Nielsen, and C. T. Pedersen, Acta Chem. Scand., 11, 1115 (1963). L. Maier, Helv. Chim. Acta, 39, 2667 (1963). A. B. Burg and P. J. Slota, Jr., J. Amer. Chem. Soc., 99, 1107 (1958). G. W. Parshall, J. Inorg. Nucl. Chem., 12, 372 (1960). K. Starke, J. Inorg. Nucl. Chem., 11, 77 (1959). R. H. Lee, E. Griswold, and J. Kleinberg, gnorg, Chem., g, 1278 (1964). Varian Associates, “High Resolution NMR Spectra Catalog," National Press, 1963, p. 1,4. D. F. Evans, J. Chem. Soc., 2003 (1959). T. J. Huttemann, Jr., B. M. Foxman, C. R. Sperati, and J. G. Verkade, Inorg. Chem., 4, 950 (1965). G. S. Benner, W. E. Hatfield, and D. W..Meek, Inorg. Chem., Q, 1544 (1964). B. L. Ross, J. G. Grasselli, W. M. Ritchey, and H. D. Kaesz, inorg. Chem., 2, 1023 (1963). R. S. vinal and L. T. Reynolds, Tnorg. Chem., §, 1062 (1964). A. A. Orio, B. B. Chastain, and H. B. Gray, £gprg: Chim. Acta, Q, 8 (1969). J. Chatt, G. A. Gamlen, and L. E. Orgeo, J. Chem. Soc., 1047 (1959). C. E. Jones and K. J. Coskran, gnorg, Chem., 12, 55 (1971). J. B. Fackler, Jr., J. A. Fetchin, J. Mayhew, W. C. Seidel, T. J. Swift, and M, weeks, J. Amer. Chem. Soc., 2.1.. 1941 (1969 ) . 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 112 A. J. Deeming and B. L. Shaw, J. Chem. Soc., 597 (1969). D. Dows, A. Haim, and W; K. Wilmarth, J. Inorg. Nucl. Chem., 21, 33 (1961). P. B. Ogilvie, J. M,. Jenkins, and J. G. Verkade, J. Amer. Chem. Soc., 92,1916 (1969), and references- therein. R. K. Harris, Can. J. Chem., 22, 2275 (1964). R. K. Harris, gnorg. Chem., 5, 701 (1966). A. R. Cullingworth, A. Pidcock, and J. D. Smith, Chem. Comm., 89 (1966). W} McFarlane, Chem. Comm., 58 (1967). S. L..Manatt, G. L. Juvinall, R. I.Wagner, and D. D. Elleman, J. Amer. Chem. Soc., 88, 2689 (1966). R. D. Bertrand, F. B. Ogilvie, and J. G. Verkade, J. Amer. Chem. Soc., 92,1908 (1970). R. Engel, Chem. Comm., 133 (1970). L. Frankel, J. Mol. Spect., 22, 273 (1969). R. Engel and A. Jung, J. Chem. Soc., 2, 1761 (1971). F. A. Cotton and G. Wilkinson, “Advanced Inorcanic Chemistry,“ second edition, New York: Interscience Publishers, 1966, p. 634. J. p. Fackler, Jr., :norga Chem., 9, 2625 (1970). I. M. Kolthoff and P. J. Elving, ”Treatise on Analytical Chemistry," Part I, Vol. 4, New York: Interscience Publishers, 1963, p. 1778. R. S. Berry, J. Chem. Phys., 32, 933 (1960). C. E. Jones and K. J. Coskran, Inorg. Chem., 10, 1536 (1971). H. Noth and H. J. Vetter, Chem. Ber., 96, 1109 (1963). K. A. Petrov, L. G. Gatsenko, and H. A. Neimysheva, Zh. Obshch. Khim., 29,1827 (1959). C. E. Jones and K. J. Coskran, gnorg. Chem., 12, 1664 (1971). 113 76. J. F. Nixon and J. R. Swain, J. Orgonometal. Chem., 77. F. G. A. Stone, Chem. Revs., 66, 101 (1958). 78. T. D. Coyle and F. G. A. Stone, grggress in Boron Chemistry, 1, 135 (1964). 79. W. A. G. Graham and F. G. A. Stone, J. Inorg. Nucl. Chem., 6, 164 (1956). 80. M. A. Fleming, R. Wyma. and R. C. Taylor, Abstracts, Symposium on Molecular Structure and Spectroscopy, Columbus, Ohio, June 1962. 81. T. D. Coyle, H. D. Kaesz, and F. G. A. Stone, J. Amer. Chem. Soc., 61, 2989 (1959). 82. L. C. Cusachs and J. R. Linn, J. Chem. Phys., 66, 2919 (1967). 83. A. H. Cowley and M. C. Damasco, J, Amer. Chem. Soc., 66, 6815 (1971). 84. A. B. Burg and H. J. Schlesinger, g. Amer. Chem. Soc., §g, 780 (1937). 85. R. W. Par and T. C. Bissot, J. Amer. Chem. Soc., Z6, 1524 (195r). 86. A. B. Burg, Rec. Chem. Progr., 16, 159 (1954). 87. R. W. Rudolph, Ph.D. Dissertation, University of.Michi- gan, Ann Arbor, Michigan (1966). 88. R. W. Rudolph and R. W. Parry, J. Amer. Chem. Soc., 66, 1621 (1967). 89. D. E. Young, G. E. McAchran, and S. G. Shore, J. Amer. Chem. Soc., 66, 4390 (1966). 90. F. Hewitt and A. K. Holliday, J. Chem. Soc., 530 (1953). 91. T. Reetz, J. Amer. Chem. Soc., 66, 5039 (1960). 92. R. A. Baldwin and R. M. Washburn, J. Org. Chem., 26, ' 3549 (1961). 93. T. Reetz and B. Katlaksky, J. Amer. Chem. Soc., 66, 5036 (1960). 94. ,M. A. A. B69 and H. C. Clark, Can. J. Chem., 66, 119 (1960). 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 112. 114. 114 M. A. A. Beg and H. C. Clark, Can. J. Chem., 66, 283 (1962). ‘W. L. Jolly, "Synthetic Inorganic Chemistry," Englewood Cliffs, New Jersey: Prentice Hall, Inc., 1960, p. 158- 160. C. H. Bushweller, H. Beall, M. Grace, W. J. Dewkett, and H. S. Bilofsky, J. Amer. Chem. Soc., 66, 2145 (1971). H. L. Hodges and R. W. Rudolph, private communication. R. W. Rudolph and C. W. Schultz, J. Amer. Chem. Soc., 66, 6821 (1971). R. Foester and K. Cohn, Inorg. Chem., in press. H. c. Brown, J. Chem. Soc., 1248 (1956). W. A. Hart and H. H. Sisler, Inorg. Chem., 3, 617 (1964). ,3 A. H. Brittain, J. E. Smith, P. L. Lee, K. Cohn, and ' R. H. Schwendeman, J. Amer. Chem. Soc., 66, 6772 (1971). E. G. Codding, C. E. Jones, and R. H. Schwendeman, private communication. H. A. Bent, Chem. Revs.,66, 275 (1961). N. Muller and D. E. Pritchard, J. Chem. Phys., 66, 768 (1959). L. S. Bartell and L. O. Brockway, J. Chem. Phys., 66, 512 (1960). E. B. Whipple, T. H. Brown, T. C. Farrar, and T. D. Coyle, J. Chem. Phys., 66, 1841 (1965). J. H. Letcher and J. R. Van Wazer, J. Chem. Phys., 66, 815 (1966). C. A. Tolman, J. Amer. Chem. Soc., 66, 2953 (1970). L. S. Meriwether and J. R. Amer. Chem. Soc., §§, 3192 (1961). Leto, J. A. Saeka and C. P. Slichter, J. Chem. Phys., 66, 26 (1954). K. J. Packer, J. Chem. Soc., 960 (1963). D. R. Armstrong and P. G. Perkins, Chem. Comm., 377 (1965). 115. 116. 117. 118. 115 W. Gerrard, M. F. Lappert, H. Pyszora, and J. W. Wallis, J. Chem. Soc., 2182 (1960). H. J. Coerver and C. Curran, J. Amer. Chem. Soc., 66, 3522 (1958). I. W. Stoltz, G. R. DObson, and R. K. Sheline, Inorg. Chem., 6, 323 (1963). G. Jugie and J. P. Laurent, Bull. Soc. Chim. Fr., 838 (1970). 302 59 lfi‘L’sIMW 0314 W