LIBRA: 7w ' THESIS Mifligan State University This is to certify that the thesis entitled Nuclear Magnetic Resonance Studies of Some Organometallic Molecules presented by William David Vernon has been accepted towards fulfillment of the requirements for Ph.D . degree in Chemistry WM Major professor " ‘4;‘5l8 Michigan State University i This is to certify that the thesis entitled Nuclear Magnetic Resonance Studies of Some Organometallic Molecules presented by William David Vernon has been accepted towards fulfillment of the requirements for Ph. D. degree in Chemistry WM Major professor {Q} 26‘ Date W94» 1 0-7639 *- *MQ‘I—l ' ‘ ‘ .‘VHOOvM1 -s--o .31 ... .1“ o.— ‘0.....‘- . - . “ ~ 3» u..- V: 6 ."~ "“' b. "' ._._ ‘~ meae C3 0 I-‘A*...A unruog-‘ 'c C ~ P N: S‘ED y- ' . M. - L. A" In. "4:5. , ‘. .I-v’h ctr. .‘Cfic 1“ ABSTRACT NUCLEAR MAGNETIC RESONANCE STUDIES OF SOME ORGANOMETALLIC MOLECULES By William David Vernon Temperature-dependent carbon-13 and proton magnetic resonance studies of solutions of ten dinuclear nS-dienyl ruthenium, iron, and nickel carbonyls have revealed that the rate of carbonyl interconversion and isomerization in these complexes is a function of both the size and the inductive character of the nS-dienyl ligand. The rate determining factor in the isomerization and carbonyl inter- conversion of the iron complexes was found to be rota- tion about the iron-iron bond, whereas bridging carbonyl opening to form nonbridged rotamers is the rate determin- ing step for interconversions seen in the ruthenium com- plexes. The difference in rates of interconversion be- tween homologous iron and ruthenium complexes may be attributed to multiple metal-metal bond orders in the ruthenium complexes. Temperature dependent carbon-l3 nmr studies of some dinuclear cobalt carbonyl complexes have revealed that these complexes are stereochemically nonrigid, and that the carbonyl interconversions observed do not proceed oa- - ...- . Q. ..a "‘ 4—, , _, -.,_v- , \ ‘:. -.4O-- ‘ a .- .- -;: ,4. ‘_ ”av 5" . .~’- " ‘- ”‘w‘ _‘_‘...~-—. '- - .-..O—1 > u .- ...uy-b-fi ' .o-.-‘- us‘ .~.uv-~. .. \_‘ s- '- ¢«.- «c . ,‘ ~~¢v-.- 5. 9-.-.“. w ~4......- _, "' sv- .- -.-.-..‘*- ‘-_ n. . _, VA A "-'--‘ ‘u‘ r- ».-._‘ - ‘ \‘ b» — ‘UI ..‘ 7' h "v- ab‘.‘ a .. V m 7‘ u .- t A ‘ ..‘.I‘c-‘—v- Vuo v. .1- ‘u‘ald . Ayn 4.‘_‘Iu .0 '5‘ ‘ ‘4“ I V'r t ~o.‘ ". .l -. 2,. n) O 'I F) William David Vernon via the well-known Cotton-Roberts carbonyl interchange mechanism. By modeling several conceivable mechanisms, it has been determined that carbonyl rearrangement in complexes with one non-carbonyl bridging ligand is effected through a concerted bridge-opening,bridge-closing process. Similar temperature dependent carbon-l3 nmr studies of some five-coordinate organotricarbonyl-iron and -ruthenium complexes indicate that these complexes are fluxional, and that it is necessary to observe the nmr spectra of some of these complexes at very low temperatures in order to fully describe the rearrangement processes. Studies of the carbon-l3 T1 relaxation times of benzene and cyclohexane with added E£i§(acetylacetonato) chromium(III) and -iron(III) demonstrate that the iron relaxation reagent is more efficient than the chromium relaxation reagent. It was also noted that benzene car- bon Tl relaxation times are decreased more rapidly than those of cyclohexane. It has been demonstrated that selenium-77 magnetic resonance measurements are easily obtained, and that the use of E£i§(acetylacetonato)chromium(III) as a shiftless relaxation reagent enhances observation of selenium-77 nmr signals of compounds with long selenium-77 T1 re- laxation times. Measurement of selenium-77 Tl relaxa- tion times as a function of temperature and measurement William David Vernon of nuclear Overhauser enhancements have established that the dipole-dipole relaxation mechanism is of little importance in selenium-77 nmr. It was ascertained that the spin-rotation mechanism dominates Tl relaxation in dimethylselenide. Studies of similar tellurium com- plexes by using tellurium nmr techniques were not as successful as the selenium studies. NUCLEAR MAGNETIC RESONANCE STUDIES OF SOME ORGANOMETALLIC MOLECULES By William David Vernon A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1975 - n"“‘\ “- . 'h-oca.: .‘v J ‘ Q 'Vou 4 . ‘A‘ A ."." ‘.:v o w. A . V-‘\‘ ..' "V‘ m... r... "V-A F .A O.» 9., . ‘. .-"“V'r_,~. o ‘4d.~»‘ - " "“Un. '4 . ‘w... u ’ ~ ., \ Vfl?,' 14"! ACKNOWLEDGEMENTS There are many people to whom I owe a debt of gratitude for their help in my educational experience. I wish to express my thanks to Dr. Otto A. Gansow for his guidance of my research activities for the last five years. He took me in as a senior at Rice to do under- graduate research, he strongly encouraged me to make the move to Michigan State, and he has been neither over- bearing nor lax in his direction of my research. I would also like to thank the powers-that-be at Rice Uni- versity and at Michigan State University for making my transition in midstream an easy one. For my synthetic background I would like to express my appreciation to Dr. Donald A. Schexnayder and Dr. Aaron R. Burke. The sound inert atmosphere techniques which they imparted to me have made my projects much less frustrating. I gratefully acknowledge the influence of my high school chemistry teacher, Mr. Alvin F. Moudy, who continually encouraged me to the point of providing supplementary materials to expand my knowledge and who has maintained an interest in my progress since those days. I must also mention my appreciation to Mr. Jim Grannan at Rice and to Mr. Wayne Burkhardt and Mr. Frank ii .4 ' ’ ’.'doc 9" '. o ..g “0 U u a. a a o-.-..':' ‘ 4 ‘. \- .vfivofi a ‘ ,.. u...» s , . o -. -u < ’— ...- In.“ 5: . - c . .A _ V“" P. I « uh. oo- v.4 . . . u \‘:‘G eve—q ~0-a... .__, v n:VI b. "“ «V ,r- I "“-w.-,: :‘V-v-..u O p ‘D.' I a‘.‘ W s "\ uoctb_-..‘ Bennis at Michigan State for their efforts in keeping the Bruker in operating condition. Thanks to Dr. Thomas A. Hardy, formerly of this chemistry department, now at Stauffer Chemical Co., for his help and concern during our trips to New York. Thanks also to Ms. Diane Mitchell, who did the artwork for this dissertation, and to Mrs. Lee Burkhardt for her meticulous typing. I wish to extend special appreciation to my wife, Susan, whose love and understanding have always been there, even though at times neither of us realized it. May the years get better! And finally, I express my most profound thanks to my parents. Since my earliest re- collections they have encouraged my curiosity and my attempts to learn. They have backed me economically when necessary, but more importantly, they have backed me with intangibles that I can never repay. It is to my parents that I dedicate this work. iii V- n-OI‘A-C div-IO- 0 -. ... b“ .p _ 2; ..... C .5 T. if. L» “V... J‘ «J ... Cay nu. P.J In“ cnPJCC FiUU mhL WU; — a TABLE OF CONTENTS Page INTRODUCTION . . . . . . . . . . . . . . . . . . . . 1 Chapter 1 CONCEPTS IN NMR . . . . . . . . . . . . . . 5 The NMR Experiment . . . . . . . . . . 5 The Nuclear Overhauser Effect . . . . . . . 23 T1 Relaxation Mechanisms . . . . . . . . . . 30 Chemical Shift Theory . . . . . . . . . . . 42 2 A CARBON-l3 AND PROTON MAGNETIC RESONANCE EXAMINATION OF SOLUTE STRUCTURES, EQUILIBRIA, AND STRUCTURAL INTERCONVERSIONS IN SOME DINUCLEAR nS-DIENYL RUTHENIUM, IRON, AND NICKEL CARBONYLS . . . . . . . . . . 52 Background . . . . . . . . . . . . . . . . . 52 Experimental . . . . 66 Acquisition and Treatment of Spectral Data . 74 Results . . . . . . . . . . . . . . . . . . 81 Discussion . . . . . . . 123 Suggestions for Subsequent Research . . . . 142 3 A CARBON-l3 MAGNETIC RESONANCE EXAMINATION OF THE FLUXIONAL BEHAVIOR OF SOME FIVE- COORDINATE ORGANO-IRON AND -RUTHENIUM CARBONYL COMPLEXES . . . . . . . . . . . . . 145 Background . . . . . . . . . . . . . . . . . 145 Experimental . . . . . . . . . . . . . . . . 150 Results . . . . . . . . . . . . . . . . . . 151 Discussion . . . . . . . . . . . . . . . . . 159 4 SELENIUM-77 NMR,MEASUREMENT OF T1 RELAXATION TIMES AND ELUCIDATION OF T1 RELAXATION MECHANISMS OF THE SELENIUM-77 NUCLEUS IN SOME ORGANOSELENIUM COMPOUNDS. THE UTILITY OF TRIS(ACETYLACETONATO)CHROMIUM(III) As A RE- EAXATION REAGENT IN SELENIUM- 77 NMR . . . . 163 iv _.4 2. v O c —— 4 Aw. - “I. Chapter Page 4 Background . . . . . . . . . . . . . . . . . 163 Experimental . . . . . . . . . . . . . . 168 Results and Discussion . . . . . . . . . . . 172 5 THE EFFECT OF TRIS(ACETYLACETONATO)CHROMIUM (III) ON CARBON:I§ T1 RELAXATION TIMES . . . 193 Background . . . . . . . . . . . . . . . . . 193 Experimental . . . . . . . . . . . . . . . . 196 Results and Discussion . . . . . . . . . . . 197 6 DINUCLEAR COBALT CARBONYLS IN SOLUTION. A CARBON-l3 MAGNETIC RESONANCE EXAMINATION OF THEIR STRUCTURES, EQUILIBRIA, AND STRUCTURAL INTERCONVERSIONS . . . . . . . . . . . 205 Background . . . . . . . . . . . . . . . . . 205 Experimental . . . . . . . . . . . . . . . . 217 Results . . . . . . . . . . . . . . . . . . 219 Discussion . . . . . . . 256 Suggestions for Subsequent Research . . . . 277 REFERENCES . . . . . . . . . . . . . . . . . . . . . 280 .- .‘C‘ o (1‘ \J r-\ (’3 3" L, Table 10. ll. 12. 13. LIST OF TABLES Page Physical Data for the Previously Unreported 5 Complex, [(n -C9H11)Ru(CO)2]2 (VII) . . . . 70 Chemical Shift and Rate Constant Data for [(nS-C5H5)Fe(CO)2]2 (I) . . . . . . . . . . 83 Chemical Shift and Rate Constant Data for [(nS-CH3C5H4)Fe(CO)2]2 (II) . . . . . . . . 85 Chemical Shift and Rate Constant Data for [(nS-C9H11)Fe(CO)2]2 (III) . . . . . . . . . 87 Chemical Shift and Rate Constant Data for [(nS-(C9H7)Fe(CO)2]2 (IC) . . . . . . . . . 89 Chemical Shift, Rate Constant Data, and Isomer Distribution for [(nS-C H )Ru(CO) ] S 5 2 2 (V) . . . . . . . . . . 91 Chemical Shift and Rate Constant Data for [(nS-CH3CSH4)Ru(CO)2]2 (VI) . . . . . . . . 94 Chemical Shift and Rate Constant Data for [(n5-09H11)Ru(00)212 (VII) . . . . . . . . . 96 Chemical Shift and Rate Constant Data for [(n5-09H7)Ru(CO)2]2 (VIII) . . . . . . . . . 98 Chemical Shift and Rate Constant Data for (nS-C5H5)2FeNi(CO)3 (IX) . . . . . . . . . .100 Chemical Shift Data for [(nS-C H )Ni(CO)] 5 5 2 (X) . . . . . . . . . . . . . . .102 Thermodynamic Parameters for Cis-Trans Equilibria . . . . . . . . . . . . . . . . -116 Activation Parameters for Bridge-Terminal Carbonyl Interchange Processes . . . . . . ~ll8 vi 9 _.— V-"‘ . n u r . x ' on" Do- a. \‘V' 5.,“ - \ a 1 c-o -; I ‘ . I 0.. ' A 'v |v0 -- u- ‘- I -- .- I 0v. I P" “ I“. l ' Vt.‘ DP..." 0 \ L . A‘ -" Lon ‘v‘- p9 "" \a.. u— s. ' r “b. bon. 5‘ f. u A» s o .4. "‘ he '0 I Is Table 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. Mass Spectral Data for the Ruthenium Complexes (V)-(VIII) . . . . Structural and Activation Energy Data for Some Valence Orbital Isoelectronic nS-dienyl Metal Carbonyls 77Se T1 Data for (C6HSSe)2 with Added Cr(acac)3 77 Se Tl Data for C6H58eH . 77 Data for (CH3)ZSe Se T1 Chemical Shift and Linewidth Data for Co (CO) (XIII) . . . . . . 2 8 Chemical Shift and Linewidth Data for C02(CO)7(CAH202) . . . . Chemical Shift and Linewidth Data for C02(CO)7 Ge(C6H5)2 (XV) . Chemical Shift and Linewidth Data for C02(CO)6 DPPM (XVI) . . . . Chemical Shift Data for [C02(CO)7]2 DPPA (XVII) Page - 120 - 134 . 174 . 180 . 188 - 222 225 228 231 234 Chemical Shift Data for (nS-C5H5)Fe(CO)2Co(CO)4 (XVIII) in Various Solvents vii 237 v . v. v. . a” .C I. .e a“ “A : I. .. . 1C .31 p: w. P.» .J.. C . a . a.. an. .i.\ nu .IV A» nu.» a . r 5 “J V. p“ EVA up. 5 . ac F... . aid . .~.J phi a i A‘Hd A, AI.- LIST OF FIGURES Figure Page 1 Energy versus field strength diagram for a spin-% nucleus . . . . . . . . . . . . . . 8 2 Transitions and Boltzmann population dis- tributions for a spin I = % nucleus . . . . 10 3 Precession of a magnetic moment, i, about a magnetic field HO . . . . . . . . . . . . 14 4 (A) A pulse which rotates the magnetization vector an angle 8 onto the XY plane; (B) loss of coherency by the magnetization vector; the component vectors "fan out" in the XY plane 18 5 Energy level diagram for the interaction of two spin-k nuclei. The W's represent transi- tion probabilities . . . . . . . . . . . . . 24 6 Diagram of two nuclei I and S showing their relationship to each other and to the magnetic field, which is along the X axis . . . . . . 33 7 Representation of a T1 experiment. (A) A 180° pulse is applied; (B) the system is allowed (to relax for a period I; (C) a 90°pulse is applied; (D) the magnetization vector returns to equilibrium . . . . . . . . . . . . . . . 4o 8 Description of a terminal M-CO bond,(I) illustrates the overlap of a filled carbon 0 orbital with an empty metal d-orbital; (II) represents the overlap of a filled metal d— orbital with empty n“ orbitals of co . . . . 46 9 The crystal structure of trans: [(n5-c5H5)Fe(c0)2]2 (I) . . . . . . . . . . 53 10 Postualted structures in the Manning three structure hypothesis for solution equilibria of [(n5-05H5)Fe(C0)2]2 (I) . . . . . . . . . 58 viii ‘0 .Nb V. . V. I. .s A ll. .6 .> . n! my . ”I «H. . .~.d a O V O 'dlbl,.0v '1’- N‘ ..N....- A... Figure Page 11 Schematic view of the carbonyl interchange predicted by Cotton for complexes with bridging and terminal carbonyls . . . . . . . 61 12 A mechanism for cis-trans isomerization and carbonyl interchange in dimetallic (n5-dienyl) metal carbonyls and related com- plexes . . . . . . . . . . . . . . . . . . . 64 13 The [(nS-dieny1)M(CO)2]2 complexes utilized in the study of solution dynamics outlined in Chapter 2 . . . . . . . . . . . . . . . . . 80 14 Variable temperature carbon-l3 nmr spectrum of the carbonyl region of [(nS—CSHSFe(CO)2]2 (I) . . . . . . . . . . . 80 15 Variable temperature carbon-13 nmr spectrum of the carbonyl region of [(nS-CH3C5H4)Fe(CO)2]2 (II) . . . . . . . . 84 16 Variable temperature carbon-l3 nmr spectrum of the carbonyl region of 5 [(n -C9H11)Fe(CO)2]2 (III) . . . . . . . . . 86 17 Variable temperature carbon—l3 nmr spectrum of the carbonyl region of 5 [(n -C9H7)Fe(CO)2]2 (IV) . . . . . . . . . . 88 18 Variable temperature carbon-13 nmr spectrum of the carbonyl region of [(nS-C5H5)Ru(CO)2]2 (V); the highest field peak is CS2 . . . . . . . . . . . . . . . . . . . 90 19 Variable temperature carbon-l3 nmr spectrum of the carbonyl region of [(nS-CH3C5H4)Ru(CO)2]2 (VI); the highest field peak is CS2 . . . . . . . . . . . . . . . . 93 20 Variable temperature carbon—13 nmr spectrum of the carbonyl region of [(ns-C9H11)Ru(CO)2]2 (VII); the highest field peak is C82 . . . . . . . . . . . . . . . . 95 ix V .‘J .0. .5 '4 L l P\~ v; ,V/ ’Jl § ‘9 Figure 21 22 23 24 25 26 27 28 29 30 Page Variable temperature carbon-13 nmr spectrum of the carbonyl region of [(n5-09H7)Ru(CO)2]2 (VIII); the highest field peak is CS2 . . . . . . . . . . . . . . . . 97 Variable temperature carbon-13 nmr spectrum of the carbonyl region of (nS-C5H5)2FeNi(CO)3 (IX); the higher field peak is CS2 . . . . . . . . . . . . . . . . 99 Variable temperature carbon-13 nmr spectrum of the carbonyl region of [(n5-05H5)Ni(00)12 (X); the higher field peak is C82 . . . . . . . . . . . . . . . . 101 A more realistic depiction of the solution structure of (nS-dienyl) metal carbonyls and related complexes . . . . . . . . . . . . . 106 Plot of chemical shift vs. temperature for [(nS-C5H5)Ru(CO)2]2 (V) in various solvents 109 Temperature dependent proton nmr spectrum of [(nS-C9H7)Ru(CO)2]2 (VIII). Spectra (a)-(c), 50% CDZClz-CS2 solution; (d) CD2C12 solution 111 Description of a bridging M-CO bond. (a) illustratesthe overlap of a filled CO o-orbital with two empty metal o-orbitals; (b) illustrates the back donation ~rom two filled metal d-orbitals to empty n CO orbitals . . . . . . . . . . . . 127 Schematic diagram of potential energy vs. reaction coordinate for structural inter- conversion in [(nS-C5H5)Fe(CO)2]2 (I) . . . 129 Schematic diagram of potential energy vs. reaction coordinate for structural inter- conversion in [(nS-C5H5)Ru(CO)2]2 IV) . . . 132 Some mechanisms which might effect carbonyl interchange in singly-bridged complexes. (a) unbridged intermediate; (b) triply- bridged intermediate; (c) a concerted mechanism . . . . . . . . . . . . . . . . . 139 a. ... .O\ .n i i u . . n n v. a . c .1 it 1(- C 11 ac .r. 44 er. .2... nu. :55 .. CCec .» 3;”... ..CC n. C C ".0 n. Ciel. C. ‘ .sa rrd .r- .1. r, Q: d .v‘d . (I! .( d oild -.,.v 0-- Figure Page 31 Structure of the ligand 2- -vinylpheny1(di- pheny1)phosphine (SP) . . . . . 149 32 Geometrical isomers of the (SP)M(CO)3 complexes; (a)-(c) trigonal bipyramidal structures; (d)-(f) square pyramidal structures . . . . . . . . . . . . . . . . 149 33 Variable temperature carbon—13 nmr spectrum of the carbonyl region of cycloheptatrienetri- carbonyliron(0); the highest field peak is C82 . . . . . . . . . . . . . . . . . . . . 152 34 Variable temperature carbon-l3 nmr spectrum of the ring carbon region of cyclo- heptatrienetricarbonyliron(O); the highest field peak is TMS . . . . . . . . . . . . . 153 35 Variable temperature carbon—l3 nmr spectrum of the carbonyl region of norbornadienetri— carbonyliron(0); the high field peak is C32 . 154 36 Variable temperature carbon-13 nmr spectrum of the ring carbon region of norbornadienetri- carbonyliron(0); the highest field peak is TMS 155 37 Variable temperature carbon—13 nmr spectrum of the carbonyl region of (SP)Fe(CO)3 . . . . 156 38 Variable temperature carbon- 13 nmr spectrum of the carbonyl region of (SP)Ru(CO)3;. the highest field peak is CS2 . . . . . . 157 39 Figurative representation of the Berry mechanism . . . . . . . . . . . . . . . . . . 161 40 Plot of T1 vs. concentration of Cr(acac)3 for the selenium-77 nucleus in (CGHSSe)2 . . . . 173 41 Plot of (T1)'1 vs. concentration of Cr(acac)3 for the selenium-77 nucleus in (C6HSSe)2 . . 176 42 (a) Selenium-77 nmr spectrum of C6HSSeH with no proton decoupling; (b) spectrum with proton decoupling; (b) spectrum with proton decoupl- ing and in the presence of Cr(acac)3; spectrum with proton decoupling, no Cr(acac)3 present; note that all the integrals are approximately equal . . . . . . . . . . . . . . . . . . . . 179 xi all. .'" a ‘ WI ( C a)? 5 Ali , Q CL hlyV -\IJ F.» AIL ..J Figure 43 44 45 46 47 48 49 50 51 .52 53 Plot of T1 vs. temperature for the selenium-77 nucleus in 06HSSeH . (a) Free-induction decay (FID) of the selenium-77 nucleus in (CH3)ZSe; (b) the Fourier transform of this FID; (c) the selenium-77 spectrum of (CH ) Se obtained 3 2 by Lardon . . . . . . . . . . . . . . . (a) Selenium-77 nmr spectrum of (CH3)ZSe with no proton decoupling; (b) spectrum with proton decoupling and in the presence of Cr(acac)3; (c) spectrum with proton de- coupling, no Cr(acac)3 present; note that all the integrals are approximately equal Plot of T1 vs. temperature for the selenium-77 nucleus in (CH3)ZSe (a) Free-induction decay of the tellurium- 123 nucleus in (CH3)2Te; (b) the Fourier transform of this FID Plot of T1 vs. concentration of Cr(acac)3 for the carbon-13 nuclei in cyclohexane Plot of T1 vs. concentration of Cr(acac)3 for the carbon-13 nuclei in benzene Plot of (T1)"1 vs. concentration of Cr(acac)3 for the carbon-13 nuclei in cyclohexane Plot of (T1)-1 vs. concentration of Cr(acac) for the carbon-l3 nuclei in benzene . Plot of (T1)'1 vs. concentration of Fe(acac)3 for the carbon-13 nuclei in cyclohexane. Plot of (T1).1 vs. concentration of Fe(acac)3 for the carbon-l3 nuclei in benzene xii Page 181 184 187 189 191 199 200 201 202 203 204 ..» .- in up . c . . ea. 1 a .u. .e. In '14 eIL Ca ‘1. cc :0. pc 1’. r. a... Q. .2 .c .. u. C u. C fit at 3 x--. n. 3 (K C C u. . PC A... . .0. s .u. .1, I. t s I. Add V .) H1 .- e . d p. o .c a .A' 1': ’94 :Fi rDJ IF! '1'). But Figure 54 55 56 57 58 6O 61 62 63 64 65 66 Several structures postulated for C02(CO)8 (XIII) . . . . . . . . . . . . . . . . . . Crystal structure of C02(CO)7(C4H202) (XIV) Solution equilibrium postulated by Noack and Bor for C02(CO)8 (XIII) Crystal structure of C02(CO)8 (XIII) The dinuclear cobalt carbonyl complexes utilized in the study of solution dynamics outlined in Chapter 6 Variable temperature carbon-13 nmr spectrum of C02(CO)8 (XIII); the higher field peak is C82 . . . . . . . . . . . . . . . . . . . . Variable temperature carbon-13 nmr spectrum of the carbonyl region of C02(CO)7(C4H202) (XIV); the highest field peak is CS2 . Variable temperature carbon-l3 nmr spectrum of the carbonyl region of C02(CO)7Ge(C6H5)2 (XV); the highest field peak is CS2 Carbon-13 nmr spectrum of the carbonyl region of C02(CO)7Ge(C6H5)2 (XV) at -164°C Variable temperature carbon-13 nmr spectrum of the carbonyl region of C02(CO)6[(C6H5)2PCH2P(C6H5)2] (XVI); the highest field peak is CS2 Carbon-13 nmr spectrum of the carbonyl region of C02(C0)6[(C6H5)2PCH2P(C6H5)2] (XVI) at -l64°C . . . . . . . Variable temperature carbon-13 nmr spectrum of the carbonyl region of [C02(C0)7]2[(C6H5)2PCCP(C6H5)2] (XVII); the highest field peak is CS2 Variable temperature carbon-l3 nmr spectrum of the carbonyl region of (nS-CSHS)Fe(CO)ZCO(CO)4 (XVIII) xiii Page 206 209 211 213 216 221 224 226 227 229 230 233 236 fi .p' .q. no ‘o 2.. . . 2. .e 3 .AJ .u A.» ._ . v e . e :. «v. a.» ¢'. . my. ¢.i.\ fly a u a o n h - o . ee~ . v. _.. “HA. V1 in . w .474 . .... 2. a. »J ... ,» La Au 4.. L» we . a :1 r» 3 PC 2;. 5 ~. . any at .. f c... L... udJ ADJ Figure 67 68 69 70 71 72 73 74 75 76 77 ”Ambient temperature carbon-l3 nmr spectrum of the carbonyl region of (US-C5H5)C0(CO)2 (XIX); the high field peak is CS2 . . . . . . . . . . . . . . . . . . . Probable structure of C02(C0)7Ge(C6H5)2 (XV) Computer-generated carbon—l3 nmr spectra of C02(CO)7Ge(C6H5)2 (XV) at -102°C; (a) peaks in 4:3 ratio; (b) peaks in 5:2 ratio; (c) measured spectrum Possible isomers of C02(CO)6[(C6H5)2PCH2P(C6H5)2] (XVI) Possible isomers of [C02(C0)7]2[(C6H5)2PCCP(C6H5)2] (XVII) Plot of 6C0 vs. temperature for (HS-C5H5)Fe(CO)2Co(CO)4 (XVIII) in a variety of solvents Schematic depiction of Mechanism I. All seven carbonyls are scrambled in a polytopal rearrangement process through a trigonal bipyramidal intermediate . Schematic depiction of Mechanism II. Two pairs of terminal carbonyls and the bridging carbonyl are scrambled . Schematic depiction of Mechanism III. Concerted opneing of the bridging carbonyl and closing of one terminal carbonyl scrambles five of the carbonyls The measured carbon-13 nmr spectrum of C02(CO)7(C4H202) (XIV) together with computer—generated spectra based on Mechanism II (left) and Mechanism III (right). . . . . Alternate view of the rearrangement effected by Mechanism III xiv Page 238 241 244 246 246 251 263 267 269 272 274 Hi 4 ..v ' ‘ I- - . --‘ _.----‘-‘ o --»-—> -I — — , .......»—- o ‘ .-v<- ‘— . ‘ ‘ ~ .-.--- -—~ a ' ‘ -- 'r‘ . r-d. -.- d l c .. "4 ‘ - “*— —‘.‘ g I . »-... .1 . 4 I I ...-. ""' ‘ ‘ -~_—-~-- ' o .o. ' :"‘~A~ *v- , >- “‘.--»g.. 'v-o 3“ ._. P - ...,—".;:‘ -. e, _ 5.. I a- ”b I“ A I‘ * ‘w-t‘ INTRODUCTION This dissertation combines topics in two widely separated areas of chemistry, both of which have only recently come into prominence. In 1946 Bloch et a1.1 first observed the nuclear magnetic resonance phenomenon, and, in 1951, ferrocene was first synthesized,2 an event which marked the beginning of modern organometallic chemistry. By the late 1950's, nmr instrumentation was sufficiently sophisticated so that researchers could obtain routine proton nmr spectra, and a few organo- metallic researchers, most notably Wilkinson, utilized proton magnetic resonance in their studies. The advent of a routine fluorine-19 nmr spectrometer also created an interest in fluorocarbon-transition metal complexes during the late 1950's.3 Studies of organometallic complexes by using proton and fluorine nmr techniques became common- place during the next decade. Unfortunately, proton and fluorine nmr studies eventually became limited in scope because, in most complexes, these nuclei are usually of peripheral interest. For example, if one were studying the nature of a metal-carbon bond by using proton nmr, the spectrum obtained transmits information which has been attenuated through at least one extra bond. Only recently, 1 .— ‘11 U-" 4“ . - ‘- . 4.7. " ~x \ '.J I" - . .'A. ' _.. ' U a of H . _-, ..-- s I"‘ \- .... 4— . ' t .- ..- -- . . x I U , . - ....-.. . . - .o.--e. -J - 0"- -.. - ' v I ... -Ja- . o - "¢§-'- ‘ . . v‘.v... .4 '--¢— 1 ,u ‘ (1‘) .fl A F; 5.. “'fi'vu ‘- .-.. ..\ V”“‘v-.I "'e‘e‘utu j 0 u ._g o 9* “a: '5': v,‘ -; “-5. ‘u. ‘ ”as“ v 7’"-- u e. V— _. u. ‘A ‘f‘ ‘ “ u A 5’ n I. .‘ . u "c. has instrumentation been developed which measures nmr spectra of carbon and other more pertinent nuclei. The first carbon-13 nmr signal was observed in 1957 by Lauterbur.4 During the ensuing ten years, carbon-l3 nmr was only a laboratory curiosity available to but a few specialists who focused their attention primarily on carbon-l3 nmr spectra of organic compounds After 1968, when Fourier transform techniques made the acquisition of carbon-13 nmr spectra feasible on a routine basis, studies could be undertaken on organometallic complexes. Since 1968 the number of papers dealing with carbon-l3 nmr studies of organometallic complexes has increased dramatically.5 Meanwhile, other remarkable improvements of nmr instrumentation had been brought about. In the last year it has become possible to modify some existing nmr spectrometers in a manner such that, with minimal effort, the spectrometer may be tuned to any one of a wide range of radio frequencies and therefore may be used to detect nmr signals from any of a host of different elements. Equipped with such an instrument, the researcher can measure nmr spectra of several different nuclei within a complex. The experimental work which forms the backbone of this dissertation is based on applications of such a ”multinuclear" magnetic resonance spectrometer to prob— lems of interest in organometallic chemistry. a! :- ‘9 "‘ a .r ,o‘ .»..~ 0 .."" On ,. —00"". -v c I ‘ p ...-'.v a. .. -_.-,..-- o - .' -21:..'.'...C..- D ..—..'.—-'D a o u DI' -..—§... «I. . "‘ I‘v’w s.- ‘ “-O- ass-g-.. b .. ...."..‘~ o.‘ "Ov-v...:. _- v I. - .."". an 2'. ava.‘“ v.‘ 0.. .."UU.. ..‘ ""“uv- 1 ~ . "“h---G; 1: a... u -§-'5. _ - ”‘4. ‘§., ' - ‘ \ No. ..“' II“. A V. 5.1"“. “ .§ ““7 7 r '-. ._¢‘u I § u I‘ I. 0 '§;-‘ ‘- v-6 -- 'A..~' . . n“ a . “Ru-"5 “ Q . u~° I“. ‘An 5 - I ‘v- ‘0 g, The arrangement of the topics of this dissertation is as follows: a theoretical section is included initially in order to acquaint the reader with the basic principles of pulsed and carrier wave magnetic resonance, and of the theory of the nuclear Overhauser effect, Tl relaxation measurements and mechanisms, and chemical shifts. The experiments performed and data interpretation follows next. This portion has been divided into several chapters according to the several organometallic systems studied. The first chapter describes a variable temperature carbon-13 and proton nmr study of solute structures, equilibria, and structural interconversions of [(nS-C5H5)Fe(CO)2]2 and some of its derivatives. Following this part of the dissertation is a short chapter which oulines a carbon-13 nmr study of some fluxional five-coordinate organotri- carbonyliron(0) complexes. The next chapter describes a series of selenium-77 magnetic resonance studies of some organoselenium compounds and as well attempts to duplicate these studies on similar tellurium compounds by using tellurium magnetic resonance measurements. That chapter is essentially a physical chemical study of the selenium-77 nucleus in various environments. The next chapter reports a series of experiments designed to in- vestigate the mechanism by which "shiftless" relaxation reagents interact with organic substrates to effect a reduction in the T1 relaxation times of constituent car- bon atoms. The last chapter presents an examination of w- a. ..-.'- ~--»»ou on «o- .a . - w- - c .. ° N . - 0“ _-. ’V ‘ - ~ .-n v ‘ '... ,— ..- ‘fi ‘2“ .. .n-- .» '-.- a-v'“’ 4" D .4- "b o , ,-’-v n». . ".'I '3» r a - .-.-o— --~ .- '( .-.....-..--J A . .o b-» as D‘ \ .n. up..- how C02(CO)8 and some of its derivatives in solution similar to the study conducted in Chapter 2. The solu- tion behavior of the parent complexes in Chapters 2 and 6 has provided the fuel for a controversy for the past twenty years. These chapters will indicate how several techniques which we developed7 have allowed us to sort out the solution dynamics of these complexes as well as others now in the literature. .A a.“ \ D . _ _ ' — —‘ " .___——————‘ .— a 0v v . .,- ve~ a 7‘ V -0- I'D-iv " Q ’5‘ .. -- .-....\. .4 --—0 ,~> .— ‘."-§ ~ .~ ‘v.- a“. ‘ .H \ .."I He..- ‘Iov‘ A v i 7' 'rgra ,q_ . O I 8‘ ‘u. - p -.. "‘*-» -.. c ”3-; f. a ‘ ‘ ..». “~41: ..\ . 3 9 A *‘ I - I- \_ :»F‘ e. ’ "‘\ ~ 1 -. . L’VV ..A‘A‘c~‘ ~ - A~-.-.“ ‘~‘ [4 4 ~~»~“‘ ~- ..">,‘ . .l.‘ u'u '. H ‘5 'l '1 , -— 1 “‘o‘.' .' . h“ K.‘ .' ‘- s,_: -h \ h u u "v .1“ n CHAPTER 1 CONCEPTS IN NMR The NMR Experiment To explain the origin of the hyperfine structure observed in optical atomic spectra, Pauli8 hypothesized that some atomic nuclei possess an intrinsic magnetic moment, the magnitude of which is a function of the charge, spin, and angular momentum of the nucleus.~ If the spin quantum number of a nucleus is denoted by I, then the nucleus will have (21 + 1) states, all degenerate in the absence of an external magnetic field, the angular momentum component of which will have the values I, I - 1,..., -I along any direction. Therefore, as a consequence of spin and charge, a magnetic moment u may be defined for any atomic nucleus. The moment may be classically viewed as a vector parallel to the angular momentum vector and is defined as fi= YNh 1. (l) where YN is the magnetogyric ratio with dimensions -l.sec-l. If the nucleus is treated as a radians-gauss spinning spherical particle with uniform distribution of mass and charge on the surface, a magnetic moment arises 5 A... .‘_.w...l 0" -. - ..V I ‘— V; . n.--- ' a. . .v“ ._ ‘f ‘ . u~-* ‘. u , 4 ., __,-.- o x .--b-° 4 . ._--’-— - .— - ..1-v~.....- . c—v-.. ’— . ---'\.-4 '- _~-. ’-_. ....__-.-.... . - . ’ “ " '-~ - -h .,,, o F '7 p..‘ -.. r‘l ° " mun. -.. V'r- 7' .. ll - c . a u... .. . . '..C x r- “V L . ._ -v_‘_ . l' A A 1‘ "§‘ C- . \ru which imparts to h the value fiN = gN(efi/2MC)I = gNBNI’ (2) where gN is the nuclear g factor, e and M are the charge and mass of the proton, respectively, c is the velocity of light, and 8N is the nuclear magneton. When a nucleus with nonzero spin is placed in an external magnetic field, the field and the magnetic moment u interact; this interaction is described by the Hamiltonian H . (3) Or, if the direction of the magnetic field is defined as the z axis, H = -Yfi H12, (4) where H is a uniform external magnetic field. For a nucleus of spin I, there are (21 + 1) energy levels separated by a field-dependent energy. This splitting of energy levels in the presence of a magnetic field is termed the nuclear Zeeman effect. The basic nmr experiment involves transitions between nuclear Zeeman levels. Transitions are induced by irradiating a macroscopic ensemble of nuclei in a static external magnetic field with photons of a frequency v which satisfy the Bohr relation E = hv = uHO/I; or v = uBNHO/hl, (5) > .. _,.- ..— w‘ _.-~* , . . - .'.-~ ,. ‘ __ .--e..— . . -— — ‘ ‘v .. .- --~. ....-..- — ' \ - ofivL-I'd vi "'-v-.- 1 u. u—r -c-~.--C 4 . . '~-.- ‘> ’- ‘ fl .- ‘-"-v.b. . I o .I. _ ‘;0.—“ e.-. w .'I-. ‘0‘! . 7 h .. ; - .4... -. \ u ‘0 ye. . ' . u f‘ \ LI, 'u-, . _ \u—"‘I‘b-. '4». In ‘vd. ‘ .‘2‘l\‘1 u :'~, a -..,c Q . u..~"= -.' e“- 'c‘ \ 3“ "U V I\. I - "o . - ‘QP ~.": - \ . - 5 § A g V 7 where u is the scalar magnetic moment, u = gNI; this implies v = gNBNH/h . (6) In terms of the magnetogyric ratio, v = AE/h = yHo/Zn (Hz). (7) It is observed that in a given magnetic field Ho’ tran- sitions between atomic energy levels are possible only at discrete frequencies, with the magnetogyric ratio or the magnetic moment of the nucleus being the primary factor determining the resonance frequency. Figure 1 illustrates the dependence of the energy difference between Zeeman levels on both the magnitude of y and of Ho; I = % is this case. Since the classical nmr experiment involves observation of energy absorption by a collection of nuclei as transitions are induced between two energy levels, some concern must be directed towards the relative popula- tions of the two energy levels. As depicted in Figure l, the nuclei in the lower energy level are described by a spin a and those in the upper level by a spin 8. When a macroscopic ensemble of nuclear spins is at thermal equilibrium at a temperature T, there are slightly more spins in the a state than in the 8 state. The ratio of the two spin populations is given by the Boltzmann factor exp(gNBNHO/kT), or 8 >. hV==2flfiNrM 9’ ‘c’ u: a Heki --—3> Figure 1. Energy vs. field strength diagram for a spin I = a nucleus . .- . .l ‘~'V‘- .v . 4.. ~.. ‘- \ I q‘ q ~_ . u" \ . -_ . ._.- “‘5' «.3. .,. § . -. A.-- "C -. - s "‘L . s u», a '1 £‘>' l ‘\- ‘Q — r. V‘ . I _ _~>Q .-" - .J-. ’4‘ .-:Q ". " -5» --~-- . . i 3 u - \ x -- . , . ,‘fl -—" Ia ‘ _- -....- .— Q . . -—--- ‘7‘ .--..- -4 . u. - _. - . P .-......b -' A—I‘. ... ‘--‘— ‘v _ .4 '- o .".~‘~—. . .._ P "'---w . “ I -. . ‘ A . "~_._' ' ‘ ges.’ "fl. "‘-¢t, Na = %N(l + gNBNHO/ZkT) and N8 = %N(l - gNBNHO/2kT).(8) In practical terms, for carbon nuclei in an external magnetic field of ga. 20 kilogauss at room temperature, Na/NB is on the order of 1.00000345, or, we may rather say that the excess population is 9a. 3.5 ppm. The transitions which are induced by radiofre- quency irradiation of a collection of nuclei are illustrated in Figure 2. Let W018 and Wad denote the probabilities per unit time for a given nucleus to undergo an upward or a downward transition by some mechanism. These quantities are not equal, since at equilibrium the number of a + 8 transitions must equal the number of B + a transitions, or NowaB NBWBd ' (9) By combining Equations 8 and 9 we may write de/de = 1 + gNBNHO/kT . (10) If n denotes the excess number of a spins, then a transition 8 + 6 increases n by 2; thus we can write the differential equation dn/dt = ZNBWBG - ZNawaB . (11) By letting W = (waB + wBa)/2’ the equation becomes dn/dt = -2W(n - neq), (12) where neq = (Na + NB)gNBNHO/kT . (l3) 10 e-gNflNHO /2kT e+gNBNHO/2kT Figure 2. Transitions and Boltzmann population distribu- tion for a spin I = % nucleus . -u- . ,-- I . 'AA - I‘.. .. . ‘— __ ...- Iv»...n .. — . ., n- .O" - \ .A. u-o" -' ’ -_-_ 5---... .4 _ ... -a»-y hi . O “'P fig.“ 9" s - V "" -&-oa . I . . “V“~ ova ’ "‘--~L fl... .1 I ~ I ‘-‘ .-A F -~-- .-:- c -. up ’ . r ”"""‘ M» D ,- r-u a v-» A C. .4-» ‘p..‘"' o I ’“v-s 5‘. _ . “-s... I... " ‘ f“, ‘o. >4 I V ¢L‘_ 3 e: , v- >’_ ’r ., u- A .g' 51.1" J... . ,t l u- ‘._""'v' -‘p *V 9 v. . . ’V- 11 Equation 12 may then be rewritten dn/dt -(n - ne )/T1, (14) q in which Tl 1/2W. (15) The solution to Equation 12 is then seen to be (n - neq) = (n - neq)oexp(-t/Tl). (16) The factor T1’ with dimensions of time, is designated as the spin-lattice relaxation time. It represents the time during which the difference between the excess population and its equilibrium value is reduced by the factor l/e. The spin—lattice relaxation time is also a measure of the rate at which the spin system comes into thermal equilibrium with its surroundings, also called "the lattice"; or alternatively, it is a nonradiative transi— tion between two energy levels. There are several mechanisms by which spin-lattice relaxation can occur; the nature of these processes will be discussed in a succeeding section of this dissertation. An important practical aspect of nmr, implicit in the concept of the spin-lattice relaxation time,is satura- tion. Saturation occurs when the rate of absorption of energy exceeds the rate of the return to equilibrium; hence, the excess population of the a spins is reduced and therefore so is the probability of additional energy absorption. The consequences is that no nmr signal is observed when a spin system is saturated. . '_ u-s .-.-a ,.’_.C b ' . '-‘ V. a“ 1‘ “ 3.- dou~ . .-- .- "' ‘5 - I '.-l, - o . - -_.< ~_\ , .a‘ .c-uvhbd .uq-.. -. v a. in... VI" I ' “’------ U \ a -¢..»-.‘v~ '1 O ‘ "-*..a— “'NO¢-.. I pa, C \ ~ .. A- 'u: .4 ' U‘& s...“ '§ \ v».-. 9 p a c Q s '5 §Ar,a Y— .,.‘~ v sc :- \ '~ r p.‘ \y u t .1". ~ . .‘x '- u.\’: ‘¢ ‘5 V A r ‘. b—‘.'».. 5.}?— x. | H... ’- \ u- . o, u ' ' C A -L ‘ I '.. 1 Q‘ we” § I '. 0 ..V «v..- u t s , n A A‘l ‘§ .. u s 'a, l- e“: x, u" t. , , .7; ‘-: yank 12 The classical nmr experiment involves placing a sample containing the appropriate nucleus in a strong, fixed magnetic field (typically on the order 103 to 104 gauss) and irradiating it with radiofrequency photons by slowly sweeping the frequency. Observation of the fre- quencies at which nuclear spin transitions occur is the object of the experiment. In practice, however, the magnetic field is usually swept at a constant radio- frequency. This continuous wave (cw) technique is satisfactory for observation of nuclei such as the proton, lithium-7, fluorine-l9, or phosphorus-31. These nuclei possess large magnetogyric ratios and/or are relatively abundant. Unfortunately, most other nuclei with spin I 2 % exhibit neither attribute, and so special techniques must be employed in order to observe their nmr absorp- tions. Hahn9 experimentally observed the effect of a single radiofrequency (rf) pulse applied to a spin system as suggested by Bloch, gt 31.1 The pulse experiment is basically quite different from the cw experiment, since it depends upon the behavior of the spin system monitored just after the application of a single rf pulse, rather than its being observed during the continuous application of low-level rf energy. Since pulse techniques are used almost exclusively in sophisticated nmr instrumentation, the following discussion will focus on the theory of pulse techniques. 4 2‘ w— . ~; .. .'--b' ~.. _. _ - ‘ _ "--».- ...- ‘ -n; .- . ’ ..-,.. -~ . .... ’F‘ . -- ~-'~ . .7 .. ... 4 __ .--.v -', v .- .o u.- u.‘. -" ' o ._ ._. — ~ ..~» .- -"v\o. , - . .‘ -._ ~ “ ‘4 -. - . .- " v“ I Q... o. . ~u-a.‘ ’ H “'b "'h m. ..H V]. " an - . 4,. P» .7-.. :1” 3 u-v"“v ; .v'J‘ ‘ .- I. .. ' p Ll A . . 'c. . .w z - __ f"! if. :‘u. - . ~-_; I ‘> ._ *L’ l 'l to r - (n 13 To understand the processes that occur during a pulsed nmr experiment, it is necessary to begin with a classical mechanical treatment and then to proceed to the Bloch equations. From classical mechanics, it is known that the torque exerted on a magnetic moment by a magnetic field H0 at some angle 0 induces the moment to precess about the direction of the field with a frequency given by the Larmor equation: v0 = -YHO/2n (Hz), (17) Energy absorption occurs only at the resonant condition v0 = Vrf. This situation is pictured in Figure 3. The magnetic vector component of the applied rf field rotates in the xy plane perpendicular to H0 at the Larmor frequency. Upon absorption of energy from H1, the net magnetic moment vector is tipped to a different value of 0, but the value of v0 remains constant. During an nmr experiment the behavior of a large number of nuclear spins is observed. For this reason it is necessary in our analysis to treat the sample system as an ensemble of nuclear spins. As depicted in Figure 3, the z axis is defined as the direction parallel to the applied, fixed magnetic field, HO. Oriented along this axis is the net macroscopic magnetization vector M which is the resultant vector of the excess spins a aligned in the direction of Ho‘ The rate of change of the angular momemtum p of a spinning nucleus is proportional to the 14 Figure 3. Precession of a magnetic moment, u, about a magnetic field HQ . ‘ v' . - \ .- v"' . .- V ' . . 4‘-" C“ — ,O'. > - -' I s "‘ . u-~--" '- ~- 9.»... - ,., ...~ .~. -. ...>.. .. ‘ A ..—-A- - F I n " he. _ , .. -..-- ,‘c. . . ._v. . r- u-_ .. ., . ._~ 3 .- I " \ a - t n" (I) ( 1' (‘D 15 cross product D x H, or the torque, exerted on the nuclear magnetic moment by the field H, or dp/dt = D x H . (18) Since fi = yp, it is also true that dm/dt = Y(dP/dt) = Yfi x fi; (19) ~ or, since M = Zfi, ~ dM/dt = VM x h . (20) The term H in the preceeding equations includes both the static field HO and the magnetic vector of the applied field H1. The components of H are therefore HX = H1 cos wt Hy = -H1 cos wt (21) H = H , z o where m is the angular frequency of the vector. By combining these equations with Equation 20, the time de- ‘pendence of the components of M may be calculated: dMX/dt = y(MyHO + MZHl Sin wt) dMy/dt = y(MzHl cos wt - MxHo) (22) sz/dt = -y(MXH1 sin wt + MyHl cos wt) The relaxation times T1 and T2 are next introduced into the equations. The return of Mx and My to the equilibrium values of 0 and M2 to its equilibrium value MO is assumed to be a first-order process. T2 is the time constant describing the process in the x and y components of M and Iv! .., 'v on. 0.. (l‘ .n' a 4 ‘ .p- lt' a 16 T1 is the time constant describing this process in the 2 component. The Bloch equations in final form are then seen to be: dMX/dt y(MyHO + MZH1 sin wt) - Mx/TZ dMy/dt = y(MZHl cos mt - MxHo) - My/T2 (23) dMZ/dt = -y(Mle sin mt + MyH1 cos wt) - (Mz - Mo)/T1’ T1 is called the longitudinal relaxation time and is entirely equivalent to the spin-lattice relaxation time described previously. T2 is termed the transverse or spin-spin relaxation time. The implications of the Bloch equations are more easily understood if the system of equations is transformed into a rotating frame of reference. This reference frame is simply one that rotates about H0 in the same direction in which the nuclear moments precess. In a rotating reference frame, Equation 20 becomes ~ (dM/dt)rot = yM x H - a x M, (24) where a defines the angular frequency of rotation by its magnitude and the axis of rotation by its direction. Equation 24 then becomes “v (ail/dorot = Vfi x t + it x m/y = yM x (a + m/y>, (25) in which the term (H + a/y) represents an effective field Heff’ Therefore in the rotating frame of reference (dM/dt)rot = yM x Heff. (26) 17 It is thus seen that the net magnetization M precesses about Heff in the rotating frame of reference. In the rotating frame, the time dependence of the 2 component of M is given by sz/dt = -YH1My - (M - MO)/T1. (27) 2 It is now possible to describe fully the behavior of the magnetization vector during a pulse experiment. When an ensemble of nuclear spins comes under the influence of a laboratory magnetic field HO along the z axis and an applied, pulsed magnetic field Hl perpen- dicular to H0 (and therefore in the xy plane), both H1 and the rotating frame of reference rotate at an angular 1 velocity of m rad-sec- In the rotating frame the effective magnetic field is H = Ho + m/Y + H (28) eff 1 ’ which at resonance simplifies to Heff = H1. Hence, in the rotating frame of reference M, precesses about the z axis at a Larmor frequency VH1. (See Figure 4A.) The vector M precesses through an angle 0 during a time tp given by 0 = H t . 29 Ylp (7 When H1 is removed, the nuclear moments begin to lose energy to their surroundings and proceed to equilibrium with a time constant Tl' When the moment is tipped onto the xy plane, processes also occur which cause the nuclei Figure 4. A B (A) A pulse which rotates the magnetization vector an angle onto the XY plane; (B) loss of coherency by the magnetization vector; the component vectors "fan out" in the XY plane .— " \ a ‘ Hi 19 to exchange energy, and the vector moments thus begin to spread out, or lose coherency, as shown in Figure 43. Since H0 is not usually homogeneous, nuclei in different portions of the sample experience different values of HO and hence precess at slightly lower or higher frequencies than does the rotating frame of reference. In this case, ~" My proceeds to equilibrium with a decay constant TE. When processes other than field inhomogeneity cause My to return to equilibrium the time constant is denoted T2. It should be noted here that the linewidth of a resonance is dependent upon the magnitude of T2 the width at half- height of a Lorentzian-shaped resonance is given by V% = l/nTZ. (30) In practice it is usual to study a spin system which, at a fixed value of HO, contains chemically non- equivalent nuclei, each with a different precession fre- quency. In a pulsed nmr experiment, instrumentation is utilized which generates an envelope of rf power approximat- ing very closely the positive half of a square wave. The rapid switching necessary to generate such pulses generates Fourier components over the range v0 : 1/tp, where v0 is the monochromatic applied frequency of the instrument and tp is the length of the applied pulse. Every frequency component in the range v0 i l/tp is present, and any nuclear spin which precesses at a frequency included within this frequency range is excited by the appropriate . I..— o ‘-..b 0‘- ,..— .- ~ - ...— - ...v ‘ -o-‘ .v .v'-" A v .- ..o-v- "’ -._..- .. it ....v _ . 'vu... -. ‘ -.- -.‘ .- . -vv--v‘. . ---p-. _- . ..‘-_ n ‘ ", .o-..,,.‘. w. '~A ' .-."-"':‘ "rv..~‘ - v. n - \ " rev ' Q .. o \ xv- .— — ,4 ...-Io‘- . ‘n. rc- . v4 n_."‘ H ...' . .V‘u " u ., "Jh. in A 2. ~ . a , ,_"_ ,— M"§_ “‘\ ‘55. .- L vu.‘ ru~ l A H x I “v P' . “A i.: 'v - "b! .l -, - .VA'..I .. u. H) v 5“ I . 5V. - 20 frequency. A short rf pulse is therefore equivalent to the simultaneous application of a wide range of radio- frequencies to the spin system. With the aid of Figure 4A, we will consider the effect of a pulse on a spin system. After a pulse of duration tp’ the magnetization vector M is tipped to the angle described in Equation 29. After the pulse, the vector M lies somewhere in the xy plane, and a signal is induced in the detector coil in the xy plane which is proportional in strength to the magnitude of Mxy' The signal produced is a free induction signal, since thenuclei responsible for generating it precess without applied rf. As transverse relaxation occurs, the free induction signal decays as an exponential with time constant T3, ‘I where T; is the effective T2. Its value is determined from Equation 30, where T: replaces the T2 term and v% includes line broadening from field inhomogeneities. If the pulse is not precisely centered on the Larmor fre- quency, or if there are several chemically-different nuclei present in the sample, the free induction decay (FID) is not a pure exponential function but instead is a more complex waveform exhibiting interference effects identical to the "ringing" observed in conventional cw nmr. The free induction decay observed is actually an 10 nmr spectrum in the time domain. It has been shown that —. 4- .- -—-‘o.. '- u a“ o ‘c '-. 7 0- .“DE if. C _U : .-- .' .--"‘ ~ ‘w". '- ,_ . .v“ b.‘ .A .3: u» ‘ ','--vv- ‘\ ' . ‘0'... 5 . . ..-, y‘-.. (I) t 'A‘ t... . q \— a ‘e._' i 21 the time domain spectrum and the more useful frequency domain spectrum are Fourier transforms of one another, the Fourier transform pair being defined as F(m) = I f(t)exp(-iwt)dt -oo (31) f(t) = I (1/2n)F(m)eXp(imt)dm Modern computers, by using the modified Cooley-Tukey algorithm,11 can effect the Fourier transform in a matter of seconds. The resulting frequency domain spectrum may then be analyzed in the conventional manner. The power and utility of the application of pulse techniques to the measurement of nmr spectra can now be made clear. By using conventional cw techniques, a spectrum is obtained in a matter of minutes by slowly sweeping through the entire frequency range over which the nuclei contained in the sample may possibly precess. The pulse techniques (hereafter referred to as Fourier Trans- form NMR, or FTNMR), have the effect of examining the entire spectral width in a few hundred microseconds, and the FID is usually monitored for several hundred milli- seconds. liconventionaltime delay for use between pulses is perhaps one second, so in one minute a spectrum may be scanned sixty times, each scan fully equivalent to one cw scan. Each FID is coherently added to the preceeding FID's and the coherently—added FID's are stored in a computer. Sensitivity enhancement can be achieved by uguv ~.~ .ne’ "7 . a ...b -‘15 r _- 'n‘ | ‘ ._ ...t , l -.-a V‘ R u \ ..... .- 3 ..- - —-\. ‘.-~¥ .- . .. .-’.. ... .F‘ . h- b. 22 employing a large number of transients per spectrum, since the signal-to-noise ratio increases as /N, where N is the number of transients. Likewise, a savings in time is achieved; the use of FTNMR generally results in a time-saved factor of 10 to 100 as compared to the use of conventional techniques. .7 V...— s“ Linn-LI; SW run-v, ii... ”We; . . - I'- .-— .. ' '5 c. ,...4 . - - -:V‘ ‘ . coo—J. -u - ....V ‘ 4 ..._ A- .- ncuubb up --~v--.— — v o-vb-o-r-h 7“ W". A *7- -~.- ,, ‘O‘ as." F . "' -.. ~..._,_‘ “1- ." ts.-- . -.. ‘au _ .. -~“'“C ‘ . l-u § .. ‘ s _ . C u "h’“ h“ - u-.. ‘ .-~ _ . ~V~VI . t»‘. I. ‘ ._- - \..V~. “'5 L. ‘ \L .- .‘--\,.~ "“t,» H -‘D U) 23 The Nuclear Overhauser Effect The nuclear Overhauser effect is defined as the change in integrated intensity of the nmr absorption of a nuclear spin system as the result of the concurrent saturation of another resonance. This effect was pre— dicted by Overhauser,12 who originally applied it to electron spin resonance. His calculations indicated that saturation of an electronic resonance should increase the amount of nuclear polarization and, consequently, increase the magnitude of the associated nuclear resonance. The 13 then further predic- effect was subsequently observed, tions were made that a similar effect should be manifest in nmr. The nuclear Overhauser effect is commonly seen in molecules containing two different species,both with spin l/2. The most frequently encountered instance is in carbon-l3 nmr, in cases in which the saturation of the energy levels of protons coupled to carbon atoms is observed to enhance the intensity of the carbon reso- nanCes. The energy level diagram for a two-spin system (both spins = %) is shown in Figure 5, where it has been assumed that the two nuclear species are loosely J-coupled. If the spins are labelled I and S, we find that the energy levels are H 24 2 an W2 W3 \\\ W11 \WO\\ 3.130 [M 4. " 33 Figure 5. Energy level diagram for the interaction of two spin % nuclei. The W's represent transition probabilities. T‘“ aim-g3! - a hr“ -.'-'. . " . .l . O .1 ‘ D I r‘ :I .» .- . I I. . .v“"‘. ‘ ‘ a ,....-O. . I-I. "" vI-An v“ ‘ O ‘ I v-.-. -. ‘ ‘ I‘.._..‘ (I) ‘b 1" ",.,Y- 5‘: 25 level 1 spin I is a, spin S is a-aa level 2 spin I is a, spin S is B-oB (32) level 3 spin I is B, spin S is a-Ba level 4 spin I is 8, spin 8 is 8—88 The Wh terms in the figure represent the four transition probabilities which describe the spin lattice relaxa- tion times. Wi represents the single quantum transition probability that spin I will undergo the transition a + B or B + a while spin S remains unchanged; W? represents the analogous transition for spin S; W2 re— presents the two quanta transition probability for simultaneous relaxation of two spins in the same direction, do + BB or B8 + dd; and W0 represents the zero quantum transition probability for a mutual flipping of the spins, dB + Bo or Ba + dB. The equilibrium populations of the levels are denoted PO, where P? = K exp(-Ei/kT), (Ei is the energy of level I) and K is the normalization constant. The approach of Pi to equilibrium is given by the equation _ _ o _ _ 0 Now the longitudinal magnetization M2 is related to the average value of the nuclear spin operator IZ by the relation MZ - Nyfi , (34) where N is the number of spins per unit volume. Since the case in which the spins form a loosely coupled system is , r » . , ., - .- --'-. ‘ ... q 5 . . q l -fi._--_a ‘w... -‘ _ 'In- ‘ D "On. . a.;. 'O ~~ 'IA . DA. I 1‘ A. .‘~ ’.'.-..1 «‘: ‘1 _‘ ~ ‘, J 26 of interest, the energy levels may be viewed simply as eigenfunctions of I2 and consequently it is possible to discuss the magnetization of each single spin. The magnetization of a spin k is given by = i > i MkPi . (35) i where Mk is the quantum number Iz(k) appropriate to level 1; and the summation is carried out over all energy levels (there are four such energy levels in a two spin = % system). From this equation it is found for the two spin-% system that = %P1 + %P2 - gP3 - %P4 and (36) From Equation 33 we obtain the result dP l _ o 0 EE‘ ‘ ‘(le + w13 + W14)(P1 ‘ P1) + w12(P2 ‘ P2) 0 O + W13(P3 - P3) + 1.114(1)4 - P4) , (37) where the other quantities P2, P3, and P4 yield similar results. By combining Equations 35 and 36 we observe that d z _ o _ _ 0 dt ‘ ‘(w13 + W14)(P1 ‘ P1) (”24 + w23032 P 2) (38) . o . O + (”13 + w23)“)3 ' P3) + (”14 + W24)(P4 ‘1 4)' .‘vo- .3, 1.. .Al‘ \..‘ 27 This equation may be simplified in the following manner: . . . . _ _ I in the loose coupling limit W13 — W24 — W1, W12 w23 = O, and W14 = W2. With these modifications, equilibrium value of IZ is = L 0 L O _ L 0 - L 0 IO 2P1 + 2P2 2P3 2P4 , which enables us to simplify Equation 38 to d z _ I _ - 0 dt — -2Wl[ - IO] - WOL(P2 - P 2) ’ (P3 O O . ' w2[(Pl ' Pl) ‘ (P4 ' P4)J S = w = W1, 34 the (39) O - p3)3 (40) By applying Equation 36 to Equation 38 we find that + (Sz> = P1 - P4 and - = P2 - P3; hence d dt = -OI[ - IoJ - OISE - So] and d dt = -pS[ - So] - OSIRIZ> - IoJ ’ where _ I = II pI 2W1 + W0 + W2 l/Tl , _ S . _ SS pS — 2W1 + NO + W2 - l/Tl , . _ _ IS _ SI and 018 - OSI — l/T1 — l/T1 (41) (42) (43) 28 The 0 terms are relaxation rates and the 0 terms are cross relaxation terms. ~'The cross relaxation terms will be nonzero if and only if a relaxation mechanism is present which couples spins I and S. The three spin— lattice relaxation mechanisms which couple spins I and S are the dipole-dipole relaxation mechanism and the two scalar relaxation mechanisms, mechanisms which will be discussed in depth in an ensuing section of this disserta- tion. The cross relaxation terms 018 and 081 are the terms that describe the nuclear Overhauser effect. If the spin S is saturated so that = 0, it is observed that the integrated intensity of spin I, , is enhanced by the fraction nI(S) = ( — 10mO <44) compared to the equilibrium intensity, IO. At the steady state = IO + OISSO/QI (45) and the fractional enhancement becomes nI(S) = GISSO/OIIO . (46) The fractional form nI(S) = ySoISS(S + 1)/yI I1(1 + 1) (47) is derived from the relations IOdI(I + l)YI and SOaS(S + l)yS. If S and I are heteronuclear with spin %, and if spin I relaxes solely through dipole—dipole coupling v‘ .v ..‘ . -- Q I \ A..-u 0-,. in»..- (It ( 29 to S, then OIS/pI = 1(1 + 1)/25(s + 1) , (48) which leads to the expression for the fractional enhance- ment of spin I nI(S) = yS/zyl . (49) This expression represents the maximum enhancement of a signal. The total maximum nuclear Overhauser enhance- ment is given by (l + n). Uh "- . -. \lrufl»li., Nb... unyrstlib E a.“ 30 T1 Relaxation Mechanisms The preceeding discussion of the nuclear Overhauser effect demonstrates the intimate relationship between the NOE and the nuclear spin-lattice relaxation time T1, as defined by the cross relaxation terms of Equation 43. The details of the several T1 relaxation mechanisms will now be considered in this section. When a nuclear magnetic dipole is in the presence of a local magnetic field H it couples to the field with L, an energy E = -u - HL . (50) The nucleus under consideration is usually a component of a molecule in motion, such as in a fluid state, so the local magnetic fields present will fluctuate in time such that HL is seen as an average value: = O . (51) However, molecular motions, such as translational and rotational motions (but not vibrational motions) randomly produce fluctuations in the value of H1. A "typical molecule" in a non-viscous liquid remains on the average about 10‘12 seconds in one state of motion, after which it undergoes a collision which changes its state of motion. The motion of the molecule will therefore be composed of 12 frequency components ranging from O to 10 Hz. The fre- quency component at the Larmor frequency of the nucleus 31 under observation will induce transitions between the nuclear energy levels and therefore cause relaxation. The result is that fluctuations in H1 induce a transfer of energy between the spins and the molecular motions (the lattice) and thermal equilibrium is attained. The equation 1/Tl = (2/3)y22Tc/(1 + ngi (52) L relates T1 to the mean—square average of the local mag- netic field and to a molecular correlation time TC which characterizes the motion which induces the relaxation process.14 All relaxation processes will be described by a similar equation; that is, an equation consisting of a correlation time characterizing the motion of the molecule and a multiplier consisting of universal constants and molecular parameters. The general form of the equation which describes any relaxation process is then seen to be = EET (53) _ 2 l/Tl — R C 1 Any mechanism whereby fluctuating magnetic fields are induced in the vicinity of a nucleus is a possible spin— lattice relaxation mechanism. The five processes which can link nuclei to the lattice are l) the magnetic dipole— dipole interaction; 2) the electric quadrupole inter- action; 3) the chemical shift anisotropy interaction; 4) the scalar-coupling interaction; and 5) the spin rotation interaction.15 _ a .h. 32 The dipole-dipole interaction is the most commonly encountered relaxation mechanism in carbon—l3 nmr. It arises in the following manner: if two nuclei I and S are located in a stationary molecule in a magnetic field, nucleus I will be influenced by a total magnetic field Ht which has components HO from the laboratory magnetic field and H10C from the magnetic moment “8 of nucleus S. Now two magnetic dipoles separated by a distance r couple with an energy E a u - u /r3‘ (54) dd 1 2’ hence the energy is dependent upon both the separation and the relative orientation of the nuclear dipoles. We then write _ 2 2 Hloc ’ i YS/rIS(3 C03 013 ‘1) _ 3 2- - YSISh/rIS(3cos o - l), (55) where O and rIS are as defined in Figure 6. If the molecule is in motion, 918 becomes time dependent and the average value of the local field goes to zero: 2n n 2 I [ (3cos O - l)sinOdOdo = O . (56) o o The motion of the molecule provides a means for nuclei I and S to move about in space and thus causes H to loc fluctuate in time. This fluctuation of Hloc induces the relaxation process. The expression which fully describes 33 c6 915 Figure 6. Diagram of two nuclei I and S showing their relationship to each other and to the magnetic field along the X axis 34 the dipole-dipole relaxation rate is seen to be16 _ DD _ 2 2 2 6 l/T1 — Rl — (4/3)YIYSS(S + l)h TC/r . (57) This interaction is therefore related to molecular geometry. It is difficult, however, to define a molecular correlation time for the dipole-dipole relaxation mechanism. The rotational correlation time for the intramolecular dipole—dipole interaction, identical to that of the quadrupolar interaction, is, for a small molecule in a nonviscous liquid, 10.10 sec > TC > 10'12 sec.16 One final note concerning this mechanism is that is is not absolutely necessary for nuclei I and S to be in the same molecule. There exists an intermolecular dipole-dipole interaction as well as an intramolecular dipole-dipole interaction. The correlation times of the two instances differ, but otherwise the treatment is essentially the same. Any nucleus with spin I > % has an electric quadrupole moment, denoted eQ. The interaction of the nuclear quadrupole moment with the electric field gradient, eq, produced by the surrounding electrons provides a very efficient mechanism for coupling the nucleus to the rotational motions of the molecule. The quadrupolar relaxation rate is given by17 1/Tl = RQ = (58) (3/40)[(21 + 3)12(21 - 1)](1 + n2/3)(e2Qq/h)rc. 35 where n is the asymmetry parameter of the field gradient and the term (eZQq/h) is the nuclear quadrupole coupling constant, the quantity which usually determines the magnitude of RQ. As stated above, the rotational correla- tion time describing this interaction is in the range 10-10 to 10'12 seconds. There can be no quadrupolar re- laxation in a nucleus with spin-%. A nucleus in a magnetic field experiences a field H10C given by the expression Hlo = Ho(l — o), where o c is termed the shielding factor. The interaction of the field and the shielding factor is represented by the Hamiltonian ”GS = -th - g - I (59) where g is the chemical shift tensor. The magnitude of the shielding at the nucleus is dependent upon the orientation of the molecule in the magnetic field. Through rapid molecular motion of the molecule in solution the fluctuations are averaged to give an average value of g: g = l/3(oXX + Oyy + 022) . (60) However, if 0 , o , and o are not equal, which xx yy 22 implies that the chemical shift tensor 9 is anisotropic, this anisotropy may create a relaxation mechanism. If m and Cl represent the shielding along and perpendicular to the symmetry axis of a molecule, the relaxation rate due to chemical shift anisotropy is seen to be 36 C3 = (2/15)y2H§(m - oi)2T . (61) A peculiar aspect of relaxation via the chemical shift anisotropy interaction is that the relaxation rate is field dependent. The carbonyl carbon in acetic acid is known to relax gig this mechanism,18 and recent work has shown that the chemical shift anisotropy spin-lattice relaxation mechanism contributes to the relaxation of the lead-207 nucleus.19'20 When two nuclei I and S are ”J-coupled” the energy is represented as ESC = JI . S, (62) which yields a Hamiltonian of the form HSC = hI - A - S. (63) In examining the relaxation of nucleus I due to fluctua- tions in the field Hloc’ it must be considered that HO can fluctuate in time in one of two ways; one in which S is time dependent or the other in which A is time de- pendent. The first instance is referred to as scalar relaxation of the second kind; this interaction occurs when I is relaxed by scalar coupling modulated by the rapid relaxation of S. This type of interaction is en- countered in carbon-l3 nmr when carbon nuclei are bonded to transition metal nuclei which possess large quadrupole moments, such as cobalt-59, manganese-55, or rhenium-187. The relaxation time of the directly-bonded carbon nucleus _. . o .A"" ,. ’_,_,.-. ' 1“ _.-...o- - _..u-¢I l p O . r (1 v... - ... -v ‘ s 4 4 u-v...‘ I ~- an. . \‘ ~7- ow- .-~& ‘y-A-~- "‘4 rut-v.» guy‘... ...h ~ ‘ I- v... w v ,1 ‘~—/. t I p,, . ,1 b“)" 1". - ‘ v. -u..( u]: ‘2‘- . ”N...“ - . Pl a, ‘.. a. ‘- u.,_ A . 5“}.3 . n. ‘ J "r ‘ 0- ..,~_ . . '- ." ' ‘va . .. cw» 1'Y\ (I) fr‘ 37 is modulated by the very rapid, quadrupolar-dominated relaxation rate of the transition metal nucleus. In the instance of scalar relaxation of the second kind, no multiplet splitting of S on I will be observed. When the operator A is time-dependent, the interaction is termed scalar relaxation of the first kind. This interaction is observed when the relaxation of I is modulated by a process such as chemical exchange. The relaxation rate due to scalar relaxation is 1/1:l = RS = (2A2/3)S(S + 1)(TC/1 + (m1 - w 212 , (64) S) where A represents the spin-spin coupling constant and TC represents the relaxation time Is of nucleus S in the case of scalar relaxation of the second kind and Te, the exchange time, in the instance of scalar relaxation of the first kind. It has recently been shown that scalar relaxation of the second kind governs the relaxation rate of the tin-119 nucleus through scalar coupling to the halogens in the compounds SnCla, SnBrA, and Sn14.21 The interaction of a nuclear magnetic moment with the motion of a molecular magnetic moment which arises from the fact that the electron density in a molecule does not rigidly follow the rapidly tumbling nuclear framework provides a mechanism by which nuclear spin energy can be transferred to the lattice. In a system in which a single electron revolves about a nucleus at a radius R, the 38 rotational frequency V is given by the eXpression v = hJ/2nI (65) for a molecule with moment of inertia I in rotational level J. The electronic motion generates current i = (e/c)V which induces a magnetic moment J. 22 (66) II? = i(nR2) = (eh/2ch)J “N “J The magnetic moment induced by the electronic motion produces a local magnetic field at the nucleus on the order of uNJ/R3. As the molecule undergoes collisions which change the direction and magnitude of its angular momentum, fluctuations in the local magnetic field at the nucleus are produced. The relaxation rate due to the spin-rotation interaction is SR 1/Tl = R = (2 IkT/h2)CZISR (67) where C is the average value of the spin rotation tensor and TSR is the spin-rotation correlation time. This relaxation mechanism is unique in that Tl relaxation times which result from its being the dominant relaxation mechanism exhibit an inverse dependence upon temperature. It is noteworthy also that freely-rotating groups may induce relaxation via the spin-rotation relaxation mechanism. Such is apparently the case in methyl iodide; in which the relaxation time of the carbon atom is dominated by the spin-rotation mechanism.23 39 The measurement of T1 relaxation times is best accomplished through utilization of pulse techniques. The most commonly employed technique involves pulsing the sample with a (180°, T, 90°, delay) sequence. The 180° pulse inverts the magnetization Mz along the z axis, as demonstrated in Figure 7A. As spin-lattice (or more appropriate to this discussion, longitudinal) relaxation occurs, MZ decays from the value -MO through the origin to its equilibrium value Mo' If, however, at a time I after the 180° pulse, a 90° pulse is applied along the x axis (Fig. 7C), the magnetization vector M2 is rotated into the y axis and a free induction signal is produced. Nuclear magnetic resonance spectrometers are designed such that the detector coil is aligned along the y axis, so that only signals oriented in the xy plane can be de- tected. The magnitude of the signal produced is propor- tional to the value of M2 at a time T; if a different value of T is then used in the sequence, the T1 re- laxation time can then be calculated. The Bloch equa- tions describe the decay of M2 back to equilibrium as dMZ/dt = -(M2 - MO)/Tl, (68) which yields upon integration MZ = Mo(l - 2exp(-t/Tl)). (69) In practice, the quantity ln(l - A/AO) is plotted against T, where A is the intensity of the resonance 40 A Figure 7. Representation of a T1 experiment. (A) A 1800 pulse is applied (B) The system is allowed to relax for a period T (C) A 900 pulse is applied (D) The magnetization vector returns to equilibrium H0 H0 Ho Ho \ ,’ I Y II \\ x .. ‘ z B C D 42 Chemical Shift Theory This section includes a discussion of chemical shift theory of nuclei other than protons, with particular attention being directed toward the carbon-l3 nucleus. A survey of the factors affecting the chemical shifts of carbonyl carbons bonded to metals and to organometallic moieties follows. i The first important theoretical interpretation of chemical shifts observed for nuclei in molecules was pre- sented in a series of four papers by Ramsey.24 Saika and Slichterzs simplified the expressions obtained by Ramsey by approximating the total shielding constant 0 for a particular nucleus as the sum of three terms: _ t o — 0d + Up + o (70) where Od’ the diamagnetic contribution, is the Lamb term26 for the nucleus of interest. The paramagnetic term op is a correction arising from the molecular environment and involves the mixing of ground and excited electronic states by the applied magnetic field. This term represents the contribution from magnetic fields induced by the orbital motion of the valence electrons. The contributions from all the other atoms in the molecule are designated by 0', a term which overall contributes very little to the total shielding because the inter- action falls off as l/r3. B.» r .- .A‘ i _..~" ,'-.--1 ‘ 4 .n- "' "Av-v r x._ ‘ u U .-‘ - . \ .ns. .1 )9—1 .. ‘9'...’ "L..- .__._‘ . ‘vx- .OV\.. ,A_- - .4 "-.s. . ll ~r- 0—. o ’u , u" o . '5 1 v. . '9 -'VP s.~ .. .5,“ . ,V. 1‘?- ‘ 43 The paramagnetic term was predicted by Saika and Slichter to dominate fluorine chemical shifts in all instances except in the completely ionic F- species. Griffith and Orgel27 derived an expression which predicted very large paramagnetic shifts for the cobalt—59 nucleus, and subsequently this was demonstrated to be true. Some- what later, Karplus and Pople28 concluded that the primary contribution to the carbon-l3 chemical shift originates from changes in the "local paramagnetic" term op. 29 Pople had earlier calculated the formula for the tensor 0AA, the local paramagnetic contribution to the shielding of a carbon atom A, to be 2 2 AA e h l -3 0 = - —-2-—2 —— Z Q (71) where = 4 36 P + P + P QAB / AB( xAxB yAyB ZAZB) - 2/3 (P P + P P + P P ) yAyB ZAZB ZAZB XAXB XAXB YAYB + 2/3 (P P + P + P ZX X PX) B AyB YA B yAzB szB A The summation Z is carried over all atoms other than A; B 2p is the mean inverse cube of the distance from the nucleus for the carbon 2p orbital, and the PW terms are elements of the change density and bond-order matrix in the unperturbed molecule: OCC. P = 2 z uv i CipCiv’ (72) 44 with the summation being taken over all occupied molecular orbitals, and the subscripts xA, XB’ yA, yB, ZA’ and 2B corresponding to the various 2p atomic orbitals of atoms A and B. The AE term is the excitation energy between the carbon ground state and the lowest lying paramagnetic -3 state of the molecule. The factor correlates 2p the electron density around the carbon atom to the size of the 2p orbital. As electron density is acquired by the atom, the 2p orbitals may expand and the 2p term decreases in magnitude and hence the magnitude of 0AA likewise decreases. Added electron density to the carbon 2p orbitals therefore results in a net increase in the total shielding and hence an upfield shift. The total variation of the screening constant due to changes in the 0d and 0' components were estimated 30 by Pople to be no greater than 20 ppm. These terms were also demonstrated to be of negligible importance in the 31 32 shielding tensor of cobalt-59 and manganese-55 , two nuclei which also exhibit very large chemical shift ranges. Stothers and Lauterbur33 surveyed over 150 organic carbonyl compounds in order to understand the factors influencing the carbon chemical shifts of carbonyls. These workers concluded that the chemical shift can be rational- ized almost entirely in terms of the shielding effect of the electron density around the carbon nuclei. Some of the afore-discussed quantitative concepts concerning carbon 45 chemical shifts were then applied by Maciel34 to the chemical shifts of organic carbonyls. Maciel concluded that carbonyl chemical shifts correlate well with the polarity of the carbonyl w-bond; he observed that varia- tions in the w-bond order were sufficient to explain the observed shielding. Gansow and co-workers first established that, for metal bonded carbonyls, the observed chemical shifts are dependent on the extent of carbonyl back-bonding. In investigations of a series of complexes of the type (n5-05H5)Fe(C0)2X,35 CN‘; and LM(CO)536, where M = Cr, Mo, or w, and L = where X = alkyl, aryl, halide, or phosphine, arsine, stibine, or amine, a linear correla- tion was observed between the carbonyl chemical shifts and the Cotton-Kraihanzel carbonyl stretching force con- stants of the carbonyl stretching frequencies. The data indicate that in the first series of complexes, when X is an electron-donating group, the carbonyl resonance is shifted further downfield than when X is an electron- ‘withdrawing group. The rationale for these observations 'may be understood by considering the nature of the metal- carbonyl bond. As illustrated in Figure 8, a forward coordinate o-bond is formed when the filled carbon o—orbital overlaps with an empty metal d-orbital and a dative metal-ligand n-bond is formed when a filled metal * d-orbital overlaps with the empty n antibonding molecular 46 In 9 LQMG + @Dcso: —» MGM c . .Qc-OQ Q $.60 II %' - .9. % Q Q M5 Q Figure 8. Description of a M-CO bond. (1) illustrates the overlap of a filled carbon 0 orbital with an empty metal d- orbital; (II) represents the overlap of a filled metal d-orbital with empty n* orbitals of co I I l‘> ' ‘.p __.-J‘ (I- 9- an cox-Al ..,u-—. g .- nil-‘4 . 1 .. b. s” u. t u .. u. s“ -. r ‘ 1 ., ‘ A b..- . ‘ x a. A c' on v. . a an.“ D-.~Ll n u.,“-‘ n -a-..‘ . ‘Ann. "cu“. ,, Al”? 5“ ‘ 47 orbital of the carbon monoxide. As electron density is released into the n* molecular orbitals of the CO by the metal, the n* orbital becomes more populated with electron density and the C-0 bond order (and hence the C-0 stretching force constant) decreases. From Equa- tion 71 we find that downfield shifts of carbonyl re- sonances (deshielding) with increasing electron density in the n* orbitals of the CO group can arise from changes in the QAB terms, the 2p term, or the AE term. A series of complexes (carbene)Cr(CO)5 and (carbene)W(CO)5 was studied by Todd 33 al.,37 who obtained an excellent correlation between the carbonyl chemical shifts and the k2 stretching force constant. These authors stated that as the metal-to-carbonyl n back- bonding decreases, the QAB term (for the M-C bond) should decrease; this would mean a smaller negative value for the QAB term and therefore a higher chemical shift for the carbonyl carbon. In a subsequent study by Bodner and Todd38 of a series of (n-arene)Cr(CO)3 complexes, a linear correlation was found between the carbonyl chemical shifts and the infrared stretching frequencies for the two carbonyl stretching modes. These authors concluded that the carbonyl chemical shift is dependent upon the extent of n back-bonding from the metal to the n* molecular orbitals of the carbonyl. Furthermore, they excluded the 2p term in Equation 71 as the dominant factor in the op tensor by demonstrating that the carbonyls of 48 (n-C H NH2)Cr(CO)3 and (n-C H COZCH3)Cr(CO)3 are the most 6 5 6 5 deshielded and most shielded, respectively; the opposite would be observed were 2p the dominant term. Bodner and Todd also noted, and quite correctly so, that 35 Gansow gt gt. had neglected the negative sign of the op screening tensor in their studies, but these authors incorrectly continue that Gansow gt gl.utilized this in- correct equation to predict a net shielding of the carbonyl resonance as the substituents become increasingly electron donating. The findings of Gansow gt gt., which antedate any contributions from Todd or Bodner to this particular area, are in perfect agreement with the findings of Todd and Bodner.37’38 Finally, Bodner and Todd discussed the contributions of the AE term and the QAB term with respect to the total shielding tensor. These authors note that changes in either term would result in the deshielding of the carbonyl carbon with increasing electron density at the transition metal. In studies of a similar scope Braterman, Randall 39 et al. obtained the carbon-13 nmr spectra of a number of (CO)6_nMLn complexes, where M = Mo, W; L = alkylphosphine, alkylphosphite, ttg(diphenylphosphino)ethane, ttg(diphenylphosphino)methane; n = 1,2. These authors observed a correlation between the carbonyl chemical shift and the lowest energy ultraviolet transition which in- volves a charge transfer from the metal to the carbonyl . 'A .. .-~ . V ' v. .. " o .— .vn-. -.v-. El" '- fl H'uAink' m r l. .. “my- u... 1...;- I _ . .="~ ..,.. 6.1 (l‘ (1‘ . a a .4 (1' (I) 49 L wk molecular orbital. They argue that the QAB term should be neglected because the symmetry of an MCO group is cylindrical (such is the case for the QAB term in acetylenes) and that therefore the AE term dominates the op screening tensor. It was also argued that for LM(CO)5 complexes the observed order for the carbonyl chemical shifts (6 increasing as L = C0 < P(OR)$ < PR3 < amine < carbene) correlates with the increasing charge donor ability (o-bonding ability) of the ligands. However, the ordering of the ligands in this manner also reflects the n-acceptor abilities of these ligands. Other communications of the general topic carbonyl chemical shifts are those which report experimental data but do not analyze the results to any significant degree. Carbon-l3 nmr studies of some gtg(RuClZ(CO)2L2) complexesao and of some ttggg square planar RhCl(CO)(PR3)241 complexes both include plots of the linear relationship between the carbonyl chemical shifts and the Cotton-Kraihanzel stretching force constants. The chemical shift tensors in CO, Ni(CO)4, and Fe(CO)5 have been rigorously analyzed in a recent paper by Mahnke, Sheline, and Spiess.“2 By using theoretical estimates, they calculated the diamagnetic screening term od for free CO and the transition metal carbonyls and from these calculations they maintain that the small dif- ference in screening between CO and Ni(CO)4 (~12.5 ppm43) can be completely accounted for by this term. These 50 authors then calculated the significance of the para- magnetic screening term op for Fe(CO)5 by using spin rotation constants which were calculated from the results 44’45 It was found that the in- of T1 relaxation studies. creased size of op, which results from increased M-CO n back-bonding, is solely responsible for the observed deshielded carbonyl chemical shift. Since the screening tensor Gav in Fe(CO)5 is considered to be axial, the relation Gav = l/3(oli + 2ol) arises. For this series of molecules the of term was found to vary to an in- significant degree, but the 0? term varied sizably for the axial and equatorial carbonyls of Fe(CO)5. That of is negligible is a consequence of the pseudolinear be- havior of the metal-bound CO. The only energy transition affecting the paramagnetic screening contribution is a transition from the ground state 2+ to a state with n-symmetry, and free CO has only one such state. The wk state of free CO is stabilized by addition of electron density from a transition metal, so the AE term in Equa- tion 71 will decrease as compared to free CO, and there— fore 02v will increase, thus inducing a downfield shift. The inescapable conclusion is that the chemical shifts of carbonyls reflect the nature of the n back-bonding from the metal; the more back-bonding, the greater the observed downfield chemical shift. Hence these authors believe the AE term is the major contributor to the observed chemical shifts. 51 It is our position that in terms of molecular orbital theory, a decrease in the amount of w back-bonding is synergic with an increase in the average excitation energy AE. That these two effects are synergic would seemingly imply that they are inseparable. From this discussion it is clear that no definitive model has as yet been put forth to account adequately for carbonyl chemical shifts in terms of only one of the components of the op tensor. It is most likely that further attempts to correlate chemical shifts with only one of the components will likewise be doomed; it would seem that within each series of complexes studied each of the terms should be weighted in a manner to accord accurate descriptions and predictions of the chemical shifts. Much work remains to be done in this particular area. CHAPTER 2 A CARBON-l3 AND PROTON MAGNETIC RESONANCE EXAMINATION OF SOLUTE STRUCTURES, EQUILIBRIA, AND STRUCTURAL INTERCONVERSIONS IN SOME DINUCLEAR nS-DIENYL RUTHENIUM, IRON, AND NICKEL CARBONYLS Background 46 the structure and Since its synthesis in 1955, behavior of [(nS-C5H5)Fe(CO)2]2 (I) and other similar complexes in solution have been the subject of over twenty scientific papers. Cotton, Wilkinson, and Liehr47 pro— posed a solution structure that contains both bridging and non-bridging carbonyl groups, the bridging carbonyls being functionally similar to ketone carbonyls. The in- frared study these authors undertook, convinced them that, in solution, (I) consists of a centrosymmetric structure with one bridging- and three terminal-type 48 obtained the carbonyl groups. Three years later Mills structure of I in the solid state by using X-ray crystallographic techniques. This work established that the molecules of I possess two bridging and two terminal carbonyl groups with all five carbon atoms of the cyclopentadienyl rings apparently being bonded to the iron atoms, since all Fe-C distances are identical. The iron atoms were found to be formally bonded through an electron 52 53 Figure 9. The crystal structure of trans x; L(n“-C5H5)Fe(CO)2]2 (I) 54 pair and the cyclopentadienyl rings are in a trans con- figuration. The two bridging carbonyl carbon atoms and the two iron atoms form a planar four-membered ring. This structure is shown in Figure 9. Bryan and Greene49 re- determined the structure in 1970. After Mills' announcement of the crystal structure of (I), Cotton gt gl.50 re-investigated the solution structure by using infrared and Raman spectroscopy and discovered that the data obtained were inconsistent with the existence of centrosymmetric structures in solution. These authors considered but rejected the possibility that more than one molecular species was pre- sent in solution. Noack51 subsequently investigated the temperature dependent infrared spectrum of (I) in a variety of sol- vents. He observed a solvent dependence of infrared absorptions in the carbonyl region, but observed no changes in the spectrum with temperature. He proposed that in solution (I) loses its center of symmetry and becomes so distorted that it exhibits CS symmetry. He rationalized the proposal by citing the observed dipole moment of I in benzene, which had recently been measured as 3.10 Debeye.52 Noack gt gt.53 , in an ensuing paper, detailed a thorough spectroscopic examination of (I) and its ruthenium homolog [(nS-C5H5)Ru(CO)2]2 (V) at temperatures 55 between ~100°C and +100°C. These results led these authors to propose that both complexes exist in solution in an equilibrium mixture of two structurally isomeric molecular configurations. At low temperatures a con- figuration with two bridging and two terminal carbonyl groups was found to be more stable whereas at higher temperatures an open form with four terminal carbonyls was proposed to exist. The bridged structure was found to be the dominant structure of (I) at room temperature, but (V) exhibited both isomers in approximately equal proportions at room temperature. Furthermore, these workers assigned a predominantly cis configuration to the isomers, which would explain the observed lack of a center of symmetry and the non-zero dipole moment. The tendency of the molecule to adOpt the cis configuration in solution was ascribed to stabilization of the metal-metal bond by o- and/or n-contributions. Complex (V) had been synthesized in 196254 and its dipole moment in solution 55 de- was measured as 2.75 i 0.12 Debeye. Mills and Nice termined the crystal structure of (V), and demonstrated that it is similar to (I), except that the metal-metal bond of (V) is 0.20 A longer. By relating infrared absorption intensities of solutions of (I) and (V) in the carbonyl region to the temperature at which the absorptions were recorded, 56 Noack obtained an estimate of the enthalpy and entropy 56 differences between the two postulated solution structures. For (I) he calculated: AH = 4 kcal/mole, AS = 3 cal/mole- degree. At +30°C the open isomer comprises 0.6% of the equilibrium mixture in nonane. For (V) he obtained the results AH = 1.56 kcal/mole, AS = 5.5 cal/mole-degree. In carbon disulfide solution at +30°C the open isomer was suggested to comprise 55% of the equilibrium mixture of isomers. The structures of (I) and (V) in non-polar sol- vents were then investigated by Cotton and Yagupsky,57 who utilized ir techniques. These workers concluded that solution (I) exists in a carbonyl—bridged dimeric config- uration in which the two cyclopentadienyl rings lie on one side of the plane formed by the two iron atoms and the two bridging carbonyl groups, and the terminal carbonyl groups lie on the other side of the plane. It was suggested that this structure and the centrosymmetric structure observed in the solid state are interconverted upon dissolution gig a non-bridged intermediate, a tautomer in which rotation about the metal-metal bond can occur. Moreover the conclusions of Noack regarding the behavior of (V) in solution were substantiated, and Cotton and Yagupsky proposed that the non-bridged isomer present to a large extent in solutions of (V) is the structure of the intermediate through which the cis and trans form of (I) interconvert. .1 57 Manning58 next investigated infrared absorptions of the carbonyl region of (I) in solution and found the absorptions to be markedly solvent-dependent. He pro- posed that, in solution, cis, trans, and non-bridged species are all present, and that their relative abun- dances are dependent upon solvent polarity. He concluded that in solutions of (I) the more polar cis species dominates the equilibrium and that there are only negligible amounts of the non-bridged isomer present. (See Figure 10.) The solvent-dependence of the dipole moment can also be explained by Manning's interpretation, to wit: the polar cis isomer is more stable in the more polar solvents, and hence the magnitude of the observed dipole moment increases as the solvent polarity in- creases. An infrared spectroscopic study of two similar iron complexes [(nS-CH3C5H4)Fe(CO)2]2 (II) and [(nS-C9H7)Fe(CO)2]2 (IV) was then performed by Manning and McArdle.59 This investigation revealed that these complexes also exist in solution as mixtures of cis and trans isomers, together with small amounts of the non- bridged species. Shortly thereafter these workers studied the infrared spectra of the homologous ruthenium complexes in solution60 and interpreted the results in terms of a complex equilibrium involving cis and trans, bridged and non-bridged tautomers. The bridged to non-bridged tautomer 58 n/ \ 0“\ 3% \no / Figure 10. Postulated structures in the Manning three structure hypothesis for solution equilibria of [(n5 - C 5H 5)Fe(CO)2] 2(1 1)) 59 ratio was observed to be dependent on ring substituent, solvent polarity, and temperature. The proportion of non- bridged species was observed to increase with increasing temperature and decrease with increasing solvent polarity. No evidence was found for a non-bridged species at any temperature or in any solvent for [(nS-C9H7)Ru(CO)2]2 (VIII). In 1969 Greene gt gi.6l’62 reported both the isolation and crystal structure of the cis isomer of (I). It may be formed by recrystallizing the material obtained from any synthetic route from a polar solvent such as dichloromethane at -78°C. These workers also observed the pmr spectrum of the cis isomer of (I); a single resonance at 4.78 ppm for the cyclopentadienyl protons was observed at room temperature. It began to be resolved at -49°C and appeared to be completely resolved into two resonances at -56°C. The two resolved resonances were observed to be of unequal intensity, suggesting that the cis-trans equilibrium postulated by Manning was sub- stantially correct. In 1969 Cotton gt gi.63 published "A definitive identification of the structures of dicyclopentadienyldi- iron tetracarbonyl in solution." These workers carried out a pmr study of (I) in various solvents and verified that, in solution, two isomeric species exist which rapidly interconvert above gg. -40°C. The relative intensities of the two resonances obtained below -40°C '- . . .- ' '_‘V I .... q . - w I 60 were seen to be solvent dependent. Cotton gt gi. postulated that bridged-nonbridged carbonyl interconver- sions of the type depicted in Figure 11 occur, and provide a mechanism by which the cyclopentadienyl rings of (I) may be interconverted between the cis and trans con- figurations. These workers undertook asimilar study of (V) as well,64 but the pmr spectrum at -100°C still showed a narrow singlet. A computer-simulated study of the pmr spectrum of (I) indicated that the activation enthalpy for the cis-trans interconversion process is gg. 12 kcal/mole. In our view the observations of Cotton gt gi.63’64 concerning the existence of the bridged-nonbridge carbonyl interconversion process fall short of actually proving that it occurs. Certainly there was a good case made for such an hypothesis, but it was not until techniques developed in our laboratory were applied to the problem that this structural interconversion process was demon- strated.7 The carbon-13 nmr spectrum of (I) dissolved in dichloromethane reveals at +55°C a single resonance 50.2 ppm downfield from the resonance of carbon disulfide, the internal reference preferred at the time for metal carbonyl spectra. At -90°C, two discrete resonances are detected, one at 18.1 ppm, the other at 82.3 ppm, a separation of nearly 1200 Hz. When the temperature is raised to -73°C a broad resonance appears approximately midway between the two low-temperature resonances. This new resonance sharpens to a maximum at -59°C, then it begins 61 (II) M——M :15 M_——M \c/ \CO ll (3 Figure 11. Schematic view of the carbonyl interchange predicted by Cotton63 for complexes with bridging and terminal carbonyls 62 to broaden concurrently with the other two resonances. All resonances are broadened into the baseline noise at -12°C. Then, a single resonance at 50.2 ppm appears and sharpens as the temperature is raised further. The car- bon-l3 nmr spectrum of the cyclopentadienyl rings, found 106.7 ppm upfield from carbon disulfide, exhibits features similar to that of the temperature-dependent pmr spectrum. An interpretation fully consistent with Manning's three-structure hypothesis and with the pmr and carbon-l3 nmr observations can now be put forward. Both cis-trans interconversion and bridge-terminal carbonyl interchange must occur in the temperature range above -35°C, since only one carbonyl resonance and one cyclopentadienyl resonance are observed. The cis and trans isomers both undergo bridge-terminal carbonyl interchange, but with different activation energies and hence at different rates. As the temperature is raised from -90°C, where no carbonyl interconversion occurs, the trans tautomer, less stable in the polar dichloromethane, begins to undergo carbonyl interconversion whereas the cis tautomer remains intact. At -35°C the cis tautomer begins to undergo bridge-terminal carbonyl interconversion; this temperature is also the one at which cis-trans isomerization of the cyclopentadienyl rings commences. Carbon-l3 nmr is thus the only technique, available at present, that is capable of providing data adequate to the task of solving the solution behavior of (I) and its homologs. 63 A year after we published the above data, a 65 (and sometime mechanism was proposed by Cotton gt gi. later by Roberts gt gi,66)that explains the rate dif- ferences observed for carbonyl interconversions of the cis and trans isomers of (I) and of some related complexes. The salient points of this mechanism, illustrated in Figure 12, now follow. Two bridging carbonyl groups open simultaneously, each moving to a terminal position on a different metal atom. Accompanying this step is what- ever degree of twist or deformation is necessary for the open species to pass intotfluamost easily accessible, staggered rotamers 12A, 12B, 12C. Since rotamer 12A has C2h symmetry, bridge opening renders the four carbonyls homotopic and random pairwise closing of two ggti carbonyls completes carbonyl interchange. However, inter- conversion of the cis isomer 24A requires, in addition, rotations of 120° about the metal-metal bond. These are diagrammed as (a) 24A212B212C224A or (b) 24A212BZl2A2 12CZ24A. Pathway (b) has two advantages. It avoids a transition state in which the bulkier organic ligands are eclipsed and it provides a mechanism by which cis-trans isomerization and carbonyl interconversion are accomplished in concert. Roberts gt gi.66 performed a computer-simulated lineshape analysis of the temperature-dependent carbon-l3 spectrum of the carbonyl region of (I) and calculated the activation energy for cis-trans isomerization to be 11.7 i 1.0 kcal/mole. 64 mmmeQEoo pmumamu paw makconHMU HmquAamcmwp- CV owHHMDmEHp CH m mwcmnoumucw Hzconumo paw cowummfiumEOmH mcmuuamwo How Emwcmnowa < .NH muswwm <§ 1: ‘.’ % 3N 65 The possibility of other mechanisms which des- cribe carbonyl interchange processes has been discussed by Cotton. Mechanisms such as a one-for-one carbonyl 65 interchange, or interconversion via a triply-bridged 72 have been discussed to some extent. These intermediate variations of the Cotton-Roberts mechanism have found little application thus far. 7 1....‘5 if Magnetic resonance studies of complexes similar in structure to (I), but with various phosphine, phosphite, or isocyanide substituents in place of carbonyl groups have also been reported. The mechanism outlined above was based partially on Cotton's results from a proton nmr study of the di(methylisocyanide) derivative of (I).65 The methylisocyanide ligands were seen to interchange in the same manner as the carbonyl groups. An X-ray crystallographic study of this complex revealed that the isocyanide ligands occupy the bridging positions in the 67 solid state. Studies of other isocyanide complexes reveal that they all undergo ligand scrambling.68’69 Similar nmr studies were conducted on the triethylphosphite66 7O derivatives of (I). These and triphenylphosphite complexes, too, were shown to be fluxional, and in the solid state, the latter complex was seen to have two car- 71 The Cotton-Roberts mechanism has even bonyl bridges. been shown to be applicable to the ligand scrambling pro- cess in the dimethylgermanium-bridged derivative of (I), l(OS-CSH5)2F62(CO)3Ge(CH3)2].72 66 Experimental All preparative procedures and manipulations were carried out under an atmosphere of prepurified nitrogen. Solvents were deoxygenated prior to use. All chromatographic procedures were performed by using Woelm neutral alumina, activity grade I (ICN Pharmaceuticals). [(nS-C5H5)Fe(CO)2]2 (I). A modified version of the method suggested by King and Stone73 was used by synthesize complex (I). A mixture of 1.5 ml Fe(CO)5 (Ventron Corp.), 1.5 ml heptane, and 15 m1 dicyclopentadiene was refluxed until cessation of gas evolution. The reaction mixture was allowed to slowly cool to room temperature, after which the precipitate which formed was filtered and re- crystallized from dichloromethane-pentane. [(nS-CH3C5H4)Fe(CO)2]2 (II). A similar method was used to synthesize this complex. A mixture of 1.5 ml Fe(CO)5, 1.5 ml heptane, and 15 ml purified74 di(methylcyclo- pentadiene) (Aldrich Chemicals) was refluxed until evolution of gas ceased. The reddish-brown reaction mixture was cooled to -15°C overnight and the resulting precipitated material was filtered and washed with pentane. The crude material (80% based on Fe(CO)5) was chromato- graphed by using dichloromethane eluant, and, after solvent evaporation on a rotary evaporator, the material was recrystallized from chloroform-pentane. [Calculated for C16H14Fe204: C: 50.3%, H: 3.8%; found: C: 50.1%, H: 3.8%.] 67 [(nS-C9H7)Fe(CO)2]2 (IV). The indenyl complex (IV) was similarly synthesized by refluxing 1.5 ml Fe(C0)5, 1.5 ml heptane, and 15 ml distilled indene (Aldrich Chemicals) until gas evolution stopped (gg. 6 hours). The resulting reddish-brown solution was cooled slowly to room tempera- ture and then poured into 60 ml of cold (0°) pentane. The resulting precipitate (representing 63% yield based on Fe(CO)5) was purified in a manner identical to that described for (II) above. [Calculated for C22H14Fe204: C: 58.2%, H: 3.1%; found: C: 57.5%, H: 3.5%.] [(nS-C9H11)Fe(CO)2]2 (III). The tetrahydroindenyl complex (III) was obtained by catalytic hydrogenation of an ethanolic solution of (IV) over Adams' catalyst at atmospheric pressure.75 This complex was purified by recrystallization from ethanol. [(nS-C5H5)Ru(CO)2]2 (V). Four different methods were utilized for the synthesis of (V) during the course of this research. Initially, the method of Fischer and Vogler76 was used to prepare the material. Anhydrous RuI3 was converted to the uncharacterized polymeric [Ru(CO)212]n by heating at 220°C in a stream of carbon 77 This material monoxide until iodine ceased to evolve. was treated with NaCSH5 in tetrahydrofuran and the product was recovered from the reaction mixture. A second method involved carbonylation of RuCl3 under 10 atmospheres of carbon monoxide pressure, and reaction of the resulting [Ru(CO)3C12]2 with NaC5 5.78 A recently-reported synthesis 68~ of this material from dicyclopentadiene and Ru3(CO)12 in heptane proved to be the most facile route to this 79 The fourth method was developed in our complex. laboratory in order to circumvent the laborious steps required in the first two methods and the use of the extremely expensive Ru3(C0)12 required in the third method. The first step of our synthesis is formation of the uncharacterized chlorocarbonylruthenium complex, Ru(CO)nClm, by refluxing a 2-ethoxyethanol solution of hydrated RuCl3 under a stream of carbon monoxide.80 To the yellow solution is added a tetrahydrofuran solution of NaCSH5 (200% excess based on RuCl3). The mixture is stirred at 80°C eight hours. After solvent removal at 3 mm, the residue is extracted with dichloromethane and the concentrated extracts are chromatographed by using dichloromethane eluant. Rotary evaporation of the sol- vent leaves a residue which is sublimed at 110° (0.3 mm). We believe that this method is superior to the three established procedures if both cost and time are con- sidered. A procedure for the rather inexpensive synthesis of Ru3(C0)12 gig the reduction of the chlorocarbonyl- ruthenium species with zinc has recently appeared,80 but our method utilizes the chlorocarbonylruthenium species directly; there is no need to form the Ru3(C0)12. [(nS-CH3C5H4)Ru(CO)2]2 (VI). This complex was synthesized from in a manner identical to that described for (V) by heating the chlorocarbonylruthenium complex with .. .‘O s c . — .- n .u .— 4 o ... \ o h .u..\ .. . - ‘0... ~1- (I) , ‘-v-\. H.“ it"; n ‘ , v , n , ' I 69 Na(CH3C5H4). Purification of this material was effected in a manner identical to that described for (V) above. [Calculated for C16H14Ru204: C: 40.51%, H: 2.91%; found: C: 40.75%, H: 2.95%.] [(nS-C9H11)Ru(CO)2]2 (VIII). The method of Humphries and Knox79 for the preparation of (V) was adapted for the synthesis of (VIII). A mixture of 1.0 g Ru3(CO)12 and 10 ml distilled indene in 50 ml heptane was refluxed 24 hours without exclusion of air. The hot solution was filtered and the precipitate was placed in a sublimation apparatus at 110° (0.3 mm) for four hours in order to re- move any unreacted Ru3(CO)12. The residue was then chromatographed by using 50% dichloromethane-50% pentane to elute the orange band. The solvent was removed and the product was recrystallized from dichloromethane-pentane. [Calculated for C22H14Ru04: C: 48.35%, H: 2.56%; found: C: 48.02%, H: 2.44%.] [(nS-C9H11)Ru(CO)2]2 (VII). The final ruthenium complex was synthesized gig the catalytic hydrogeneration of (VIII) in tetrahydrofuran over PtO2 at atmospheric pres- sure. The resulting solution was filtered, the solvent removed, and the residue recrystallized from dichloromethane-pentane. Since this complex is as yet un- reported in the literature, its properties are shown in Table 1. [Calculated for C22H22Ru204: C: 47.65%, H: 3.97%; found: C: 47.30%, H: 3.86%.] 70 Table 1. Physical Data for the Previously Unreported Complex [(n5-09H11)Ru(CO)2]2 (VII)a’b Melting Point: 123-4°C 6(c0)°, cm-lz 1986(17.9); 1954(15.4); 1835(2.1); 1809(10.0) lH nmr,d 6: 4.90 (area 1); 4.88 (area 2); 2.56 (area 4); 1.87 (area 4) a for carbon-l3 nmr data, see Table 8. b for mass spectral data, see Table 14. C in tetrahydrofuran d in CD013 71 [(nS-CSH5)2FeNi(CO)3] (IX). The mixed metal complex (IX) was prepared from nickelocene81 and Fe(CO)5 and was purified according to the method of Tilney-Bassett.82 [(nS-CSHSNi(CO)]2 (X). The nickel dimer (X) was purchased from Ventron Corporation and was purified by sublimation. Two other complexes were synthesized for eventual use in this research, but for reasons to be given were not utilized. Their synthesis follow: [(nS-(CH3)5C5)Fe(CO)2]2. The permethylated iron dimer was synthesized from Fe2(CO)9 (Ventron Corp.) and 5-acetyl-l,2,3,4,5-pentamethylcyclopentadiene (from hexamethylbicyclo 0.2.0 hexadiene and m-chloroperoxybenzoic acid)83 according to the method of King and Efraty.83’84 We were unable to enrich this material with 13CD, as will be elaborated in the next section. [(nS-CH3C5H4)2FeNi(CO)3]. A synthesis of the methylcyclopentadienyl homolog of the iron-nickel dimer, (IX), was also attempted. Reaction of the green liquid 1, l'-dimethylnickelocene with Fe(CO)5 in refluxing benzene yielded a brown reaction mixture. Following the procedure of Tilney-Bassett outlined for (IX),82 the mixture was chromatographed on a 24 in. alumina column and four bands developed. A green band eluted rapidly; it proved to be unreacted 1,1'-dimethylnickelocene. A pink band followed which was demonstrated by an infrared spectrum to be the l,1'-dimethyl homolog of the dinickel dimer (X). The third band, brown in color, was isolated 1"? v .n—v-a .y—r . . vs~‘- -v~-~ o... . .FH, bvd. Aq‘a "vu V‘qn 1" ' v...‘v 72 and treated as the desired product. The final band eluted was reddish-brown in color and was demonstrated to be complex (II). The brown material isolated from the third band proved to be so air-sensitive that neither a satisfactory analysis nor a pmr spectrum could be obtained. Inasmuch as the methylated iron and nickel complexes were obtained as byproducts, it is virtually certain that the brown product was indeed the methyl homolog. During the course of this synthesis, a new scheme for obtaining l,l'-dimethylnickelocene was used. This product was obtained by refluxing a tetrahydrofuran solution of NaCH3C H4 and Ni(NH3)6C1281 overnight. The solvent was 5 removed, the residue dissolved in dichloromethane and filtered, and the concentrated solution was chromato- graphed. The green band which rapidly developed was eluted, the solvent removed, and the green liquid remain- ing was utilized in the synthetic scheme. l3CO Enrichment. Complexes (I), (II), and (IV) were synthesized from Fe(CO)5 which had been enriched to gg. 33% in 13C0 by stirring for two hours a mixture of 1.5 m1 heptane and 1.5 ml Fe(CO)5 in an atmosphere of 90% carbon-l3 enriched carbon monoxide (Mound Laboratories, Monsanto) (250 ml at 27°C) together with 10% palladium on activated charcoal.85 After two such enrichment processes the mixture was filtered and the heptane-Fe(CO)5 mixture was used in the synthesis. Previous attempts to remove pentane from a pentane-Fe(CO)5 solution through distillation - . . .. s :15...— 73 yielded a green product, indicating that the Fe(C0)5 was being converted to Fe3(CO)12. Heptane was chosen be- cause its boiling point nearly matches that of Fe(CO)5. Neither (II) nor (IV) could be induced to exchange with 13CO directly, either in the presence or absence of the catalyst or at elevated temperatures; only decomposition products were obtained. Attempts to enrich the permethylated iron dimer directly yielded brown solutions which exhibited infrared spectra with no bridging carbonyls. Complex (III) was synthesized fron enriched (IV). Complex (V) was enriched by stirring a dichloro- methane solution of the complex in an atmosphere of en- riched carbon monoxide for two weeks at ambient tempera- ture and laboratory (fluorescent lamp) illumination. Sublimation of the enriched material effected purification. Complexes (VI), (VII), and (VIII) were enriched in a similar manner; however, the time was reduced to 72 hours and the enrichment was performed in the dark. Extensive decomposition of the products resulted when the enrich- ment process was performed in the presence of light. Complex (IX) was enriched in an identical manner to that described for (V) above. Chromatography of the enriched product yielded traces of (I), nickelocene, and Ni(CO)4 as well as the enriched complex. Complex (X) was enriched by stirring a dichloromethane solution overnight in an atmosphere of enriched l3CO.86 74 Acquisition and Treatment of Spectral Data Proton nmr spectra were obtained by using either a Varian A56/60 or HA-lOO spectrometer. Temperatures were measured by inserting a copper-constantan thermo- couple directly into a solvent-filled tube in the nmr probe and reading the temperature from a Doric Trendicator 400 digital output. Samples were allowed fifteen minutes of equilibration time. Infrared spectra were recorded by use of a Perkin—Elmer 457 spectrophotometer; both nujol mull and solution media were used, and polystyrene film was used to calibrate the spectra. Mass spectra were measured by an Hitachi RMU-6 Spectrometer by Mrs. R.L. Guile of this department. Carbon-13 nmr spectra were measured by using a Bruker HFX-lO spectrometer modified for multinuclear measurements as described by Traficante gt gi.6 and equipped with a Nicolet 1083 computer with 12K of memory, a Diablo disk memory unit, and a Nicolet 293 I/0 controller. The spectra were obtained at a frequency of 22.63 MHz. Temperature calibration was achieved by reading values from athermometer inserted directly into a solvent—filled sample tube in the nmr probe. Temperatures were observed to vary no more than : 1°C. All samples were prepared under an atmosphere of prepurified nitrogen by using suitably degassed solvents. Samples to be used in the temperature range above +9°C were prepared by using dibromomethane or toluene with 20% ‘ ; ... F1 «la 6.. 75 l,2-difluorotetrachloroethane (PCR, Incorporated) added to serve as a fluorine-l9 lock. Samples studied in the temperature range +9°C to -l30°C were prepared by using dichlorofluoromethane (Genetron 21, Matheson Gas Products), a compound which serves as both fluorine-l9 lock and sol- vent. Spectra obtained below -l30°C employed samples dis- solved in gg. 50-50 mixtures of dichlorofluoromethane and difluorochloromethane (Freon 22, Matheson Gas Products), with the former compound serving as a lock. All samples were 0.03 M in ttig(acetylacetonato)chromium (III) (re- crystallized from benzene-pentane). This reagent serves to reduce the lengthy carbonyl carbon-l3 Tl relaxation times.7’87 Samples prepared for use in the temperature range -l40°C to +70°C contained in addition 5% carbon disulfide, present as an internal chemical shift reference. All carbonyl chemical shifts are reported in ppm downfield from TMS, with the conversion factor being éobserved = 6 + 192.4. CS 2 The spectra in this section were measured over a 5000 Hz spectral width. Optimum signal-to-noise was obtained by using between 1024 and 2048 transients per spectrum, and by multiplication of the resulting free induction decay by an exponential function with a negative exponent. The resulting artificial line broadening (.2 to .4 Hz) was kept to a minimum. Computer-simulated spectra were generated by using 88 the lineshape simulation program EXCHSYS. This program ,- 11' 76 simulates nmr spectra by performing the calculation F ~ ~ ~ -« ~d1-Kll/r KlZ/T Kl3/I... l R /T -a -R /T R /T... l 1(0))(1 Re(11,12,13,..) 21 2 22 23 K31/T K32/T -a3-K33/I.. l L. \ J where I(w) is the intensity of a resonance at angular frequency w, IR is the intensity of resonance k, dk = i(wk - w) + l/Tzk, where wk of the kth intensity is to be calculated, and T2 k relaxation characterizing the linewidth of the k resonance in the absence of exchange; Rij are the elements of the exchange matrix R which specifies is the angular frequency resonance, w is the frequency at which the is the transverse th matrix the is transfer of magnetization among the sites, and T the mean pre-exchange lifetime. The input parameters for this calculation are the following: the frequencies and linewidths of the resonances in the slow—exchange limit, their relative intensities, the intensities of the resonances as exchange occurs, values of the mean pre-exchange lifetime (I) describing the rate of interchange, and the exchange matrix that specifies the details of the mechanism by which the magnetization is transferred among sites. Each chemically nonequivalent set of atoms is designated as a site. Superposition of calculated and observed spectra taken in 77 the range between the slow- and fast—exchange limits per- mits estimates of the mean pre-exchange lifetime I, and therefore the first—order rate constant k = l/I. An Arrhenius plot of the calculated rate constants yielded a value for the activation energy Ea. The other activation parameters were obtained from an Eyring plot of the rate constant data. A plot of ln(k/T) versus l/T yields a straight line with slope -AH#/R and intercept ASE/R + ln(k/h). A calculation performed in this manner insures that AHI and A35 are temperature independent. From the relation AG76 = AHI - TASI, the quantity AG§986 may be calculated. The thermodynamic parameters for the equilibrium between the bridged and nonbridged isomers of the complex [(nS-C5H5)Ru(CO)2]2 (V) were determined from the temperature dependence of the equilibrium con- stant measured from chemical shift changes as a function of temperature. The thermodynamic parameters describing the equilibrium between the cis and trans isomers of the iron complexes (I)-(IV) were determined from the tempera- ture dependence of the relative populations of the two isomers. The relative population of each isomer was obtained either by electronic integration of the spectra or by cutting out the peaks and weighing them. A plot of ln K versus l/T, where K = [trans]/[cis], yields a straight line with slope -AH/R and intercept AS/R. The values of AG298° may be calculated from the AH and AS values as described above. "”‘”3 78 The exchange matrix which was used to describe the carbonyl site exchange in the trans isomers of complexes (I)-(IV), and for the complexes (V)-(IX), is given by the general formula _§_ l-X 1-x where the fractional populations of site A and site B are l-X and X, or I = (I 18) = (1-X,X). The matrix used A! to describe the three-site exchange process observed in the cis isomers of complexes (I)-(IV) is given by the general formula I '91 0 T .SC ' T + .50 T + .DC R = .ST -T .5T , T 5C -1 0 T + .5c T 4 .SC where C represents the relative population of the cis isomer and T the relative population of the trans isomer (T = 1 - C). In order to simulate the temperature dependent chemical shift of the ruthenium complex (V), an exchange matrix was constructed to accurately reflect the relative populations of the bridged and nonbridged isomers. At -l3l°C a bridging carbonyl resonance was detected at 60.0 ppm and a terminal carbonyl resonance was detected at 6.42 ppm (with respect to carbon disulfide). The relative 'r'V' ‘ '1‘ 1’.“- ' . .nen 'pr ..ov"‘" ,..~.-A-v- .« .v-~.b~“ ,n '14- DD .44.. ' ...... .. _ ’i r- . 5.6.1.“... I "vnfiww VO-y'gen . (I) i‘ us. An. v.55. ~ C! .- h“ ’nn u... — 79 population of the nonbridged species was calculated from the relation 6 33.21(l-x) + 6.42x, where x re- observed = presents the fraction of nonbridged isomer and the quantity 33.21 is the average of the bridging and terminal carbonyl resonances. These calculations incorporate these assump- tions: 1) the bridging and terminal carbonyl chemical shifts are taken as temperature independent except for exchange effects and 2) the chemical shift of the terminal carbonyls in the nonbridged isomer is identical to that of the terminal carbonyls in the bridged structure. The exchange matrix was then constructed by defining IA as the relative population of the nonbridged species and IB as the relative population of the bridged species. The matrix thus constructed yielded simulated spectra which are congruent with the observed spectra. 80 ——M \ / Ill M=Fe VII M=Ru 8 / . /;Q\\C”hh 0C 0 IX 0 M/——C\M/CO C)c/ \g/fl‘cuz, II M = Fe VI M =Ru o fl/8\M/C \ / IV M = Fe VIII M =Ru fl 8 ./ \ . hh;:;;;flflu o g X Figure 13. The [(nS-dienyl)M(CO)2]2 complexes utilized in the study of solution dynamics outlined in Chapter 2 81 Results Proton and carbon-l3 nmr spectra of gg. 0.2 M solutions of the ten dimetallic nS-dienylmetal carbonyls (see Figure 13) utilized in the present study have been obtained at numerous sample temperatures between -160° and 100°C. The wide separations of chemical shifts and the facile, line broadening chemical exchange processes observed for these metal-bonded carbonyls required, for resonance detection within a reasonable amount of time, the preparation of 1300 enriched compounds and the use of several different solvents. The spectra obtained in this study are displayed in Figures 14 through 23, with the computer simulated Spectra shown in the appropriate place. Tabulations of carbonyl chemical shifts measured and sol- vents employed at various temperatures are provided in Tables 2 through 11. In every instance, the more downfield of the two or more signals detected at low temperatures were assigned as bridged carbonyl resonances. Higher sample temperatures resulted ultimately in broadening of all carbonyl resonances, followed by coalescence, except for the easily decomposed indenyl complexes (III), (IV), into a single averaged line which for the iron systems and for (VIII) remained chemical shift invariant with temperature. In contrast, the averaged resonances of the ruthenium com- plexes (V)-(VII) were seen to move upfield as sample ‘40] 32 55' 37' DWI..-” 22° 8‘ 0' qz'um#—*.pH~h—unuu-'\up W I JWL... 22° W _52. HL‘LL “‘V‘Jfé.“ :“T‘V £438. Figure 14. Variable temperature carbon-13 nmr spectrum of the carbonyl region of [(nS-C5H5)Fe(CO)212 (I) 83 Table 2. Chemical Shift and Rate Constant Data for T (0C) saverage abridge aterminal K(best fit) +55° 239.4a +39° 239.4a +21° 239.1a -20° 241.4b 272.5a 210.3C 133e -32° 241.7b 272.8C 210.0C 409 -40° 241.9b 273.2C 210.0C 206 -51° 242.1b 273.5C 210.0C 4 x 105f,6.06e -60° 242.2b 273.9C 210.1c 2.5 x 10Sf -70° 242.4b 274.1C 210.2C 1.0 x 10f -80° 274.4C 210.2C 4 x 1o4f -90° 274.7C 210.2C -110° 275.3C 210.3C ~126° 275.9d 210.4d a average of all carbonyls of both isomers b average of trans isomer C bridge, terminal carbonyl resonance of cis isomer d probably bridge, terminal carbonyl resonances of both isomers e rate constant for cis isomer carbonyl interchange f rate constant for trans isomer carbonyl interchange 74’ .. 4W «W11..- -LW -u° A I I 4WD-.. (.11.... Figure 15. Variable temperature carbon-l3 nmr spectrum of the carbonyl region of 5 [(n -CH3C5HA)Fe(CO)2]2 (II) 85 Table 3. Chemical Shift and Rate Constant Data for 5 [(n -CH3C5H4)Fe(CO)2]2 (II) T (°C) 5.2119128? 112%? aterminal K(best fit) +74° 240.4a +59° 240.6a +48° 240.6a +36° 240.6a 267° -14° 242.4b 274.2C 210.5C 123° -24° 242.6b 274.7C 210.6c 40° -38° 242.8b 275.0C 210.6c 4 x 105f, 13 33° -48° 243.0b 275.0C 210.7C 2 x 105f I -57° 243.2b 275.6c 21o.7° 8 x 1o4f —69° 243.4b 276.0° 210.8C 24,390f -81° 276.3C 210.8C -90° 276.5C 210.9C -102° 277.1c 210.1° -117° 277.2d 211.0d -125° 277.4d 211.1d a average of all carbonyls of all isomers average of trans isomer bridge, terminal carbonyl resonance of cis isomer probably bridge, terminal carbonyl resonances of both isomers rate constant for cis isomer carbonyl interchange rate constant for trans isomer carbonyl interchange C3 -34‘ 1111.11 425° 441' Figure 16. Variable temperature carbon-13 nmr spectrum of the carbonyl region of [(05'C9H11)F€(C0)2]2 (III) 87 Table 4. Chemical Shift and Rate Constant Data for 5 [(n -C9H11)F€(CO)2]2 (III) abridge 274. 275. 276. 276. 276. 276. 277. 277. 277. 277. 278. 278. 278. 278. T (°C) (Saverage +1° 243.1a -7° 242.8a —14° 243.6a -24° 243.8a -34° 243.9a -43° 243.9a -52° 244.0a -61° -72° -85° -96° -110° -125° -l41° a average of trans isomer b C both isomers d e 211. 211. 211. 211. 211. 211. 211. 211. 211. 211. 211. 211. 211. 211. 6terminal K(best fit) 200d 74d 15d 4 x 105°, 8d 2.5 x 105° 1.6 x 105° 5 x 104° 2 x 1046 bridge, terminal carbonyl resonances of cis isomer probably bridge, terminal carbonyl resonances of rate constant for cis isomer carbonyl interchange rate constant for trans isomer carbonyl interchange ‘w'hfifihflhfiwnuhuuaflmyghu -40. 3¢”ilflflfll‘flpbluflduu&*pu “'4F*HF“*“'"***'*VA'*' :::WW 111.31... Mt ::W 1...... -l'. ‘20. -402' Figure 17. Variable temperature carbon-13 nmr spectrum of the carbonyl region of [(nS-C9H7)Fe(CO)212 (IV) 89 Chemical Shift and Rate Constant Data for Table 5. [(n5-69H7)Fe(co)2]2 T (°C) saverage °gggggg +26° 240.3a 267.1b +23° 240.3a 267.2b +19° 240.0a 267.1b +15° 240.1a 267.1b +11° 240.0a 267.2b +9° 240.0a 267.4b +5° 240.0a 267.4b -2° 240.7a 268.7b -11° 240.8a 268.9b ~20° 240.8a 269.3b -300 269.5b -400 269.9b -490 270.1b -6o° 270.6b -70° 271.0b,272.5C -86° 271.3b,272.8C -102° 271.6b,273.2C a b average of trans isomer (IV) 5terminal 210. 210. 210. 210. 210. 210. 210. 210. 210. 210. 210. 210. 210. 210. 210. 210 210. 3b,211.5C .2b,211.6C 3b,211.6C K(best fit) 400d 200d 114d 61d 40d 18.5b 2.5 x 105° 1.5 x 105° 6.67 x 104° 2.5 x 104° 40 bridge, terminal carbonyl resonances of cis isomer bridge, terminal carbonyl resonances of trans isomer rate constant for cis isomer carbonyl interchange rate constant for trans isomer carbonyl interchange v‘ w —— “82. 1.1.1 1.1.1 9. 4* —33' ”Li -50. __v L... -‘AL—v—v—L Figure 18. Variable temperature carbon—l3 nmr spectrum of the carbonyl region of [(ns-CSHS)Ru(CO)212 (V); the highest field peak is CS2 kquf 91 OH x N No.0 de.mNN .Nw- mow x m.H $6.0 om.mmm omen moa x m.q No.0 om.mmm com: 00H $O.NH o¢.NNN owmu pm.mNN 66m- pH.NNN damn 00H x mm.m Nm.ma om.HNN com: po.HNN Oman NOH $5.0N 0H.omm on: pm.omm oH+ 50H x N Nm.mm ow.wHN oom+ pq.mam 0NN+ pq.wHN omm+ Nm.mm nm.mmm oom+ pH.NHN 0Hm+ pm.mam 6Hm+ NH.Ho nm.mom owm+ Nm.m© nm.wom oom+ Auflm ummnvx mHmCHEumuw mmmpflpno mp660udmu pmwpflunucoc N we wum>m© Auov H m A>V «.mxoovsmxmzmo-ch. pom cofludnfluumwm umEOmH paw .mUmQ ucmumcou mumm umanm HmoHEono .o manme 92 . .i. “14: . HIIIIIIIIII COHumadono pom uxmu mom m mumEomH Loon mo mmucmCOmmu Hzconpmo HmcHEHou .mwpflun zanmnoua m coHusHOm mawuuflcoumom CH p conusaom Namzo an d COHDDHOm owSHOD SH 3 mmocmcomww Ham mo maxconumo Ham mo mwmpm>m m m.me m.NmN coma: ono w.me q.mmm damau m.mma m.mmH m.HmN owHHu Auww ummflvm HmaEHmuw mm wwuflm mMmEousmu Umwpwuflucoc N mm mum>mm Aoov H m A.ecoev e dfiede WWW» _,,. Wk» 43° -40' 40? ‘46. 3» WM 111.-.... Figure 19. Variable temperature carbon-l3 nmr spectrum of the carbonyl region of 5 . . [(n -CH3C5H4)Ru(CO)2]2 (VI), the highest field peak is CS2 Table 7. T (°C) -23° -28° -33° —40° -46° -54° -98° -105° -121° 94 Chemical Shift and Rate Constant Data for [(nD-CH3C5H4)Ru(CO)2]2 (VI) Egveragea 6bridgeb Sterminalb K(best fit) 224.8 1.11 x 106 224.9 7.7 x 105 225.1 4.2 x 105 225 2 2 x 105 225.3 1.33 x 105 225.5 8.33 x 10° 254.5 200 255.1 255.6 average of all carbonyls of all resonances b probably bridge, terminal isomers of both isomers \x) L3 58" WM W 4° 1...). 10' -60° 6.5’ 1- 402° Figure 20. Variable temperature carbon-13 nmr spectrum of the carbonyl region of 5 . . [(0 -C9H11)Ru(CO)2]2 (VII), the highest field peak is CS2 Table 8. T (°C) +100° +71° +58° +55° +41° +35° +28° +20° +17° +10° +6.5° +4° -52° -60° -71° -85° -95° -ll6° toluene solution b C chloroform solution CHFC12 solution 96 Chemical Shift and Rate Constant Data for 5 K(best fit) 5averaged 6bridgee aterminale 220.7a 221.0a 222.4b 221.5a 222.7b 2 x 107 221.73 221.6b 222.0a 223.3b 4 x 105 223.4b 2 x 105 223.5b 1.33 x 105 223.8b 8 x 104 250.7 199.7 100 251.1 200.0 40 251.6 200.1 5.5 252.1 200.1 252.7 200.2 252.9 200.2 d average of bridge, terminal carbonyls resonances of all isomers probably bridge, terminal carbonyl resonances of all isomers 97 WM 58' W :w'uw-uun~1unw~wqun 20'-~u~qu»~eu.._...1__ .._....I. -500 -55' ~65. 1...]. 1.4.1 1.-...- Figure 21. Variable temperature carhon-l3 nmr spectrum of the carbonyl region of [(nS-C9H7)Ru(CO)2]2 (VIII); the highest field peak is CS2 Table 9. 1? (°C) 89° 68° 56° 35° 20° -46° -55° -65° -75° -97° -lll° average of all carbonyls of both isomers 98 Chemical Shift and Rate Constant Data for [(nS—C9H7)Ru(CO)2]2 (VIII) 6average .5 219 219. 220. 220. 220. WNNCD 247. 247. 248. 248. 249. 250. itridgeb I-‘-I-\\II—'\IU1 6terminalb 198. 198. 198. 198. 198. 198. DDb-I-‘b UI' K(best fit) 33 x 107 3.8 x 107 5.25 x 107 8.4 x 105 2.6 x 105 55.6 13.33 probably bridge, terminal carbonyl resonances of both isomers 99 .141 I)-.. IP‘igure 22. Variable temperature carbon—l3 nmr spectrum of the carbonyl region of (05-C5H5)FeNi(CO)3 (IX); the higher field peak is C32 Table 10. [(nS-CSH T (°C) §average ~14° 237.8 ~34° 238.2 -43° 238.5 -57° 238.7 -910 -102° -117° -130° Chemical Shift FeNi(CO)3] 100 6bridge 254.2 254.5 254.8 254.9 (IX) 6terminal 209. 209. 209. 209. and Rate Constant Data for K(best fit) 106 1.33 x 105 6.67 x 10° 1.67 x 104 200 40 .1“: 25' "J 1.....1 L wW-ve- -..- J Figure 23. 101 L..1 71......» Variable temperature carbon-l3 nmr spectrum of the carbonyl region of [(05-C5H5)Ni(c0)12 (X); the higher field peak is CS2 102 Table 11. Chemical Shift Data for [(nS-csasmimonz (X) T (°C) 6 +25° 225.5 0° 226.2 -38° 226.8 -77° 226.1 —84° 227.4 -98° 227.8 103 Itemperatures were raised. This chemical shift dependence was utilized for complex (V) to determine thermodynamic parameters for the equilibrium between bridged and non- bridged tautomers present in solution. Similarly, the integrated intensities of carbonyl resonance lines measured as a function of temperature for (I)-(IV) were employed to obtain the thermodynamic parameters for the cis-trans equilibrium between structural isomers of the form 24A224B. Analyses of nmr lineshapes were performed to obtain rate constants and activation parameters for the structural interconversions detected. Carbon-l3 nmr spectra measured in this study for the iron carbonyls (I)—(IV) are reminiscent of those pre— viously communicated7 by us for (I) and, as such, provide evidence for two separate carbonyl exchange processes. the most facile of which can be viewed in Figure 15 for the methylcyclopentadienyliron complex (II) as occurring between -125° and —35°C, in Figure 16 for the tetra- hydroindenyliron complex (III) as occurring between -110 and ~24°C, and in Figure 17 for the indenyliron complex (IV) as occurring between -102° and -9°C. This low tempera- ture process is clearly seen for (IV) in Figure 17 to average two equally intense bridged (273.2 ppm) and terminal (211.6 ppm) carbonyl resonances. Studies at even very low temperatures, gg. -l60°, failed to resolve similar, Se . - parate bridged and terminal resonances for complexes t‘_ 104 (I)-(III), although careful examination of the signals observed for these complexes at -125°C reveals that they are asymmetric. Activation energy data reported below leave little doubt that, at the temperatures utilized in this study, separate signals should, in principle, be seen. It is likely that viscosity broadening, which be— low -120°C limits resolution to gg. 1.0 ppm, as measured from the carbon disulfide reference line, prevents their detection. The source of the lower temperature exchange pro- cess is easily understood. Cotton gt gi,63'64 have re- ported that it is possible to alter the ratio of the two separated cyclopentadienyl resonances seen for (I) at low temperatures by increasing solvent polarity, and have argued that the more polar cis form 24A would reasonably be expected to be more stable in the more polar medium. Similarly, if less polar solvents than dichloromethane or dichlorofluoromethane are used to obtain carbon-13 nmr spectra of the carbonyl region of (I)-(IV), the intensity of the low temperature averaged carbonyl resonance in- creases in favor of the other as yet stationary carbonyl signals. The lower temperature carbonyl averaging process is therefore assigned to trans isomer bridge-terminal carbonyl interchange. The higher temperature exchange process, which occurs for (II), (III), (IV) above -l4°, -10°, and +1l°, respectively, results in the broadening and coalescing “‘W'H‘Ln Isl: . Figure 24. 105 A more realistic depiction of the solution structures of (nS-dienyl) metal carbonyls and related complexes 107 of all carbonyl resonances. Above 40°C, solutions of the indenyl complexes decompose; however, spectra of the two remaining iron systems taken at and above this temperature display the growth of an averaged, chemical shift invariant resonance found for (II) at 240.2 ppm. Concurrent with this process, both the proton and carbon-l3 signals of the cyclopentadienyl ligands of (I) are bbserved to broaden and coalesce into a sharp line. Similar, though not so straightforward data are obtained for the more complex ligand spectra of compounds (II)-(IV). This set of spectral observations can be handily accounted for by postulating that, at the higher sample temperatures, simultaneous cis-trans isomerization and cis isomer carbonyl interconversion become energetically allowed.65’66 Spectra recorded for the ruthenium complexes (V)- (VIII) may be compared with those obtained for the homologous iron complexes. Low temperature spectra again evidence the presence of both bridged and terminal car- bonyls, which, as illustrated for these complexes in Figures 18-21, are seen to broaden and coalesce at higher temperatures into a single sharp line, thereby providing evidence for only one bridge-terminal carbonyl inter- change process. Yet another feature of the high tempera- ture carbonyl spectra of the cyclopentadienylruthenium complex (V) is of particular interest. As sample tempera- tures are raised, the bridge-terminal averaged carbonyl resonance shifts monotonically upfield towards the terminal 108 carbonyl chemical shift region, without appreciable line broadening, from the averaged position which is a shift from 223.2 ppm at -50° (dichlorofluoromethane) to 208.7 ppm at 96°C (toluene). The chemical shift of the averaged signal was also seen to vary with solvent polarity at a given temperature. The graph in Figure 25 illustrates the behavior of the chemical shift of the averaged resonance of (V) as a function of both temperature and solvent. Only relatively small shifts of the averaged resonance are observed for complexes (VI), (VII) and none from (VIII). The marked chemical shifts seen for (V) at various temperatures and.invarious solvents indicate that we are able to detect the population redistribution be- tween open and bridged isomers of the form 24E,F¢24G,H as has been postulated from ir spectral studies.56 Concurrent, temperature dependent proton nmr studies have been performed for the several ruthenium complexes (V), (VI), (VIII). As reported previously,65 even at —110°C only one sharp pmr signal may be observed for the cyclopentadienyl carbonyl (V) and, under similar conditions, we could detect only one methyl proton and carbon resonance for complex (VI). However, pmr spectra of the indenyl complex (VIII) are more informative. The spectrum in Figure 26a is typical of those obtained above -30°C in a variety of polar solvents, such as dichloro- fluoromethane, dichloromethane-d2, or acetone-d6. Spectral 109 A>vm_mxoo.emxm:mu - rilr .. . . 1:. CV_ . mucm>aom mooflum> CH pow musumpdeou .m> DMHLm HmoHEoLo we uon .mN deemed m 86:. oo. om om ow 0m 0 ow - o? om u om - _ d _ q _ _ _ _ _ _ .I I. DON ..L../.T IT - ed 1 I... 9m 1 1 cum I 1 mum 2:26.34 I .N c2350 0 1 3028.4 1 0mm _ [h b I]; _ _ _ _ _ _ (same/(0)039 -‘m"air IF h" 1‘“ I E‘s; Figure 26. 110 Temperature dependent proton nmr spectrum of [(nS-C9H7)Ru(CO)2]2 (VIII). Spectra (a)-(c), 50% CDZClZ-CS2 solution; (d) CD2C12 solution 111 (0) I2’ \mmm K180 IIZb (15)-54° trans cis bnn U»... ....\x .on v.5 “to .j 112 features detected are easily assigned by reference to those 89 for the analogous (ms-indenyl)2Ru. The cyclo- reported pentadienyl ring protons form an AZB pattern, JAB = 0.5 Hz, (vA-vB) = 11 Hz, while one relatively broad, slightly structured resonance is observed for the protons on the aromatic six-membered ring. In less polar solvents such as 50% carbon disulfide-50% chloroform-d the J 1' AB/AVAB ratio becomes gg. 1.0, giving rise to a complex multiplet. At -54°C, in polar media, (see Figure 26b) both the aromatic and cyclopentadienyl resonances are coalesced into two indistinct multiplets. Below -70°C, the aromatic‘ resonances are narrowed somewhat into one relatively sharp line at 732 Hz and a broader line centered at 670 Hz. Simultaneously, the AZB multiplet is converted into a well defined, unsymmetrical five line pattern, as illustrated in Figure 26c,d. The character of the low temperature pmr spectra is strongly influenced by solvent polarity. In dichloromethane-d2 solution, the integrated intensity ratio of the two aromatic resonances is 1.58:1, whereas when the solution is diluted to 25% carbon disulfide, that ratio becomes 1.22:1. The intensity of the most upfield cyclopentadienyl line may also be observed to increase markedly upon addition of carbon disulfide. This signal becomes the dominant spectral feature of the cyclopenta- dienyl ring proton region at -70°C when the least polar solvent available to us, namely 50% carbon disulfide-50% chloroform-d1, is used. 113 Two interpretations of the temperature dependent pmr data were initially considered: a) as temperature is lowered, the rate of cis-trans isomerization is decreased and separate resonances characteristic for ligands on each isomer are seen, or b) rotation of the indenyl ligand about the ligand-metal bond slows and this in some complex manner alters the pmr spectra. Possibility (b) can be safely discarded on several grounds. First, activation energies for rotations of symmetrically-bonded nS-CS ring moieties are typically very small,90 on the order of l kcal for ferrocene.91 The temperature required for coalescence of the cyclo- pentadienyl ring pmr signals 11 Hz apart for (VIII) was -54°C, which, by assuming a reasonable pre-exponential frequency factor leads to an estimated Ea < 10 kcal, a figure far in excess of that required for this type of ring rotation. (A more complete analysis will be given below.) Furthermore, ligand pmr data obtained under con- ditions of slow ligand rotation are expected to be complex A2B2’ AZXZ’ gtg. and ABC, ABX, gtg, spectra which should exhibit solvent independent, integral intensity ratios of separated resonances. Our observations are entirely at variance with these expectations. We must account for the observed pmr spectra in terms of possibility (a). It is well established that solvent polarity affects the equilibrium ratio of cis-trans 114 . 63,64 isomers. We are therefore led to assign the sharp aromatic ring signal to the cis isomer, and correspondingly, the broad multiplet which increases in intensity in non- polar media is attributed to those protons of the trans isomer. The intense, most upfield cyclopentadienyl ring resonance is the only signal to similarly increase its intensity in solutions containing carbon disulfide. It must therefore be assigned as an A3 signal arising from all three protons on the trans isomer five membered ring; the AZB multiplet, the A2 portion of which is clearly observed at 574, 577 Hz in Figure 26c, remains for those protons of the cis isomer. These assignments can be tested for internal consistency by examining the relative inte- grated intensities of the several sets of well-separated spectral lines in Figure 26c. The cis-trans isomer ratio of 1.22:1 obtained from integration of the aromatic ring proton data may be used to predict an intensity ratio of 1.69 for the two separated five-membered ring multiplets if indeed the cis B resonance is overlapped by the intense trans A3 line. The measured ratio of upfieldzdownfield multiplets is 1.62, in good agreement with our signal assignments. A temperature-dependent cmr study of the iron- nickel system evidences only one bridged-terminal carbonyl interchange process, as seen in the spectra in Figure 22. As illustrated by the spectra in Figure 23, no such car- bonyl interchange was detected for the related nickel- I". nu. _‘-___ 1 I 115 nickel dimer (X) in the carbon-13 nmr study, nor did a pmr study of this system reveal any dynamic behavior. Finally, we note a temperature dependence of the chemical shifts of the bridging and terminal carbonyl resonances after the slow-exchange limit is reached. As can be seen by examination of Tables 2-11, some of the chemical shifts of the carbonyls, especially the bridging carbonyls, are quite large; for instance, the bridging carbonyl resonance of complex (II) shifts 3.2 ppm down- field in going from -l4°C, where it was first detected, to -125°C. Concurrently, the terminal carbonyl resonance is shifted downfield only 0.6 ppm. The resonance of the ruthenium complexes were not subject to such sizeable temperature shifts, but this is probably because these resonances were not detected until very low temperatures were reached, and consequently their behavior could be monitored over only a small temperature range. Integrated signal intensities of the widely separated resonances ascribed to the cis and trans isomers for the iron systems (I)-(IV) were compared as described in the experimental section to determine cisZtrans equilib- rium constants. The thermodynamic parameters obtained are exhibited in Table 12. The accuracy of the reported data is somewhat limited due to the small temperature range over which suitable spectral information could be obtained. Overlap between Freon lock and ligand carbon resonances precluded their use in similar studies. The data reported 116 Table 12. Thermodynamic Parameters for Cis-Trans Equilibria a b a COMPOUND AH AS AG298 I 1 3 i 0 5 1.0 i 0 5 l 0 i 0.3 II 2.4 i 0 2 2 2 t 0 2 l 8 i 0.2 III 2.5 i 0 5 1 7 i 0 6 1 7 i 0.3 IV 3.5 i 0.1 3.3 i 0.1 2 5 i 0.1 a kcal/mole. b 811. 117 are consistent with those previously determined for (I) 64 with the differences attributable to from pmr studies, the fact that our experiments were performed in somewhat polar Freon solvents. For the equilibrium between the bridged and non— bridged isomers of the ruthenium complex (V), the follow- ing thermodynamic parameters were obtained: AH = 2.6 kcal/mole, AS = 6.9 e.u., and AG298° = 0.5 kcal/mole. To justify our assumption of the terminal carbonyl chemical shift value, we note that resonance positions measured for the other terminal ruthenium carbonyls are very near to those assumed for the non-bridged structures. Additionally, terminal carbonyl chemical shifts reported for several other organoruthenium carbonyls in general do not differ markedly from our assumed chemical shift.5 Attempts to analyze the bridged Z nonbridged equilibria of the ruthenium systems (VI)-(VIII) were unfortunately thwarted by the small temperature dependence of their bridge-terminal averaged resonances. Lineshape analysis of the temperature-dependent cmr spectra for the nine fluxional complexes allowed computation of activation parameters as described pre- viously in the experimental section. The activation param- eters acquired in the manner described are tabulated in Table 13. Computer analysis of the pmr spectra of complex (VIII) proved unrewarding because of both the complexity 118 .Dm c .MHoE\0mox m 00.0 0 0.0 0.0 H 0.0 0.0 H 0.0 0.0 0 0.0 0.0 0 0.00 00 0.0 0 0.0 0.0 0 0.00 0.0 0 0.00 0.0 H 0.00 0.0 0 0.00 000> 00.0 H 0.0 0.0 0 0.00 0.0 0 0.00 0.0 0 0.00 0.0 0 0.00 00> 0.0 H «.0 0.0 0 0.0 0.0 0 0.00 0.0 0 0.00 0.0 0 0.00 0> 00.0 H 0.0 0.0 H 0.0 0.0 H 0.0 0.0 H 0.0 0.0 0 0.00 > 0.0 0 0.00 0.0 0 0.00 0.0 0 0.00 0.0 0 0.00 0.0 0 0.00 00060>0 0.0 H 0.0 0.0 0 0.0 0.0 0 0.00 0.0 0 0.00 0.0 H 0.00 Amem0ev>0 0.0 0 0.00 0.0 0 0.0 0.0 0 0.00 0.0 0 0.00 0.0 0 0.00 00060000 0.0 H 0.0 0.0 H 0.0 0.0 H 0.0 0.0 H 0.0 0.0 H 0.00 0060000000 0.0 H 0.00 0.0 0 0.0 0.0 0 0.00 0.0 0 0.00 0.0 H 0.00 0006000 0.0 0 0.0 0.0 0 0.0 0.0 0 0.0 0.0 0 0.0 0.0 0 0.00 000000000 0.0 H 0.00 0.0 0 0.0 0.0 0 0.00 0.0 0 0.00 0.0 0 0.00 000600 0.0 0 0.0 0.0 0 0.0 0.0 H 0.0 0.0 0 0.0 0.0 0 0.00 00000000 0mm00 0000 0100 6000 0 060 02:00:00 mommmooum mwcmnoumucH Hmconumo HmCHauoeumwwH0m 00w mumuoamumm COHum>00o< .ma manme 119 of the spectra and the small chemical shift differences measured between the spectral lines. It is possible, fortunately, to extract some information from the coalescence temperature (gg. -54°C) of the sets of cyclopentadienyl resonances in order to test whether cis- trans isomerization and bridge-terminal carbonyl inter- change occur at similar rates. By assuming the Arrhenius frequency factor measured for carbonyl interchange, an activation energy may be calculated for the cis-trans interconversion. Since K = 11 Hz = exp(26.6:1.5)exp(Ea/RT) at coalescence, we can calculate Ea (cisitrans) to be gg. 14.9i0.4 kcal/mole, in excellent agreement with E8 = 14.5i0.5 kcal/mole measured for bridge-terminal carbonyl interchange. The results of a mass spectral investigation of the ruthenium complexes V-VIII are tabulated in Table 14. Naturally-occurring ruthenium is comprised of seven isotopes, with the mass number and the percent natural abundance of each isotope as follows: 96 (5.51%); 98 (1.87%); 99 (12.72%), 100 (12.62%); 101 (17.07%); 102 (31.61%), and 104 (15.85%). Since each molecule of these complexes contains two ruthenium atoms, ruthenium- containing fragments exhibit complex mass spectra. The numbers quoted in the table are with respect to the isotOpe of mass 101, since the atomic weight of ruthenium is 101.07. 120 Table 14. Mass Spectral Data for the Ruthenium Complexes (V) — (VIII) Complex (m/e); (relative abundance) V 446(32); 418(20); 390(8); 362(8); 334(100); 206(36); 280(20); 231(4) v: 474(24); 446(21>; 418(20); 388(73). 360(100); 332(14); 306(12); 280(12); 259(6) VII 552(13); 496(18); 468(19); 434(100); 406(13); 331(11) VIII 546(14); 462(19); 434(100); 406(20); 380(20); 331(8) 121 For complex (V), a molecular ion appears at m/e 444, and is seen to lose four carbon monoxide groups in a stepwise manner to yield the most intense peak at m/e 334. Then stepwise loss of two fragments with m/e 26 is the next major feature observed. The fragments of m/e 26 are probably acetylene, or C2H2 moieties. A very small peak due to ruthenocene (m/e 231) was also detected. The mass spectrum of complex (VI) exhibits a molecular ion at m/e 474. Stepwise loss of two carbon monoxide groups is followed by loss of an m/e 30 fragment. The most intense ion m/e 360 appears after loss of the final carbon monoxide group. Following the appearance of the most intense ion, successive loss of two m/e 30 fragments is again observed. Only a very small peak due to l,l'-dimethylruthenocene (m/e 259) is observed. A molecular ion is also observed in the mass spectrum of complex (VII), with m/e 552. Simultaneous loss of two carbon monoxide groups results in the appearance of an ion at m/e 496. Then a mass loss of m/e 34 follows. The most intense ion then makes an appearance at m/e 434; this ion exhibits the same m/e as that observed in complex VIII. However, the appearance of the most intense ion of complex VII is followed by loss of the final carbon monoxide group to yield a peak with m/e 406. A small peak at m/e 331, corresponding again to (C9H7)2Ru+, is also observed. 122 A molecular ion at m/e 546 is observed for complex (VIII), which is followed by an ion at m/e 462, indicat- ing that three carbon monoxide groups are lost simul- taneously. Following the loss of the final carbon monoxide group, the most intense ion appears at m/e 434. Successive loss of two CzHé groups from the indenyl ring is then the final important step. A small amount of (C9H7)2Ru+ is detected at m/e 331. The single most important feature of the mass spectra of all four ruthenium complexes is that the ‘majority of the ions observed contain two ruthenium atoms, an observation which is in sharp contrast to the mass spectrum of the diiron complex I, in which no ions con- taining two iron atoms are observed. This reflects the relative strength of the metal-metal bonds in the two dif- ferent homologous series. That the integrity of the ruthenium—ruthenium bond is maintained under conditions of a mass spectrum experiment indicates that it is consider- ably stronger than the iron—iron counterpart.92 123 Discussion The results of this investigation clearly demon- strate that structural interconversions in solution of dinuclear nS-dienyliron and -ruthenium carbonyls are very general and facile processes. Both cis-trans isomeriza- tion and intramolecular bridge-terminal carbonyl inter- change could be examined by carbon-13 and/or proton nmr techniques. The trans structure of the iron systems (I)- (IV) is seen to interconvert carbonyls at different rates and with very different activation energies than the corresponding cis isomers. Carbonyl interchange for the cis isomer is always slower than for trans and, moreover, is always accompanied by simultaneous cis-trans structural interconversion. In contrast, concurrent proton and carbon-l3 nmr studies of the ruthenium homologs (V)-(VIII) show that in these systems both cis and trans isomers simultaneously undergo the isomerization 24E224F and inter- change bridged and terminal carbonyls. The sizeable tem- perature dependent upfield chemical shifts detected for the bridge-terminal averaged carbonyl resonances of (V)- (VII) suggest that non-bridged structures such as 24G,H are appreciably more stable in solution than for the corresponding iron system tautomers. These spectral results offer strong confirmatory evidence for the nmr interpretations we have previously presented7 for (I) and, as well, support the mechanistic explanation of these data recently put forth by Cotton 124 65 et gi. and Roberts gt gi.,66 and outlined in the introduction to Chapter 2 and in Figure 12. In order to test the general applicability of this mechanism, we per- formed experiments designed to answer the following questions: i) Can the (presumably steric) barrier to rotation be altered by changing the nS-dienyl ligand? ii) Does changing the inductive character of the ligand have a measurable and understandable effect on structural interchange rates? iii) Will the mechanistic considera- tions apply to structural interconversions of homologous ruthenium systems? To answer these questions, the dinuclear molecules (I)-(VIII) were chosen for study since they offered systematic variation of two properties, extent of alkyl substitution and ligand size. The portion of space which must be subtended by the rapidly rotating metal-bonded n5- dienyl moiety increased from (I) + (IV) and alkyl sub- stitution similarly increases the electron donating pro- perties of the cyclopentadienyl ring for (I) + (III). Activation parameters recorded in Table 13 demonstrate that decreasing the bulkiness of the ligand has a remarkable lowering effect on the free energies (A6298°) and enthalpies (AHI) of activation measured for cis isomer carbonyl interconversions in the iron system. Parallel observations of temperature dependent pmr spectra for (I)-(IV) leave little doubt that this process is accompanied by'andintegral to cis-trans isomerization. h___._ I u 125 Activation energies for the two processes in (I) agree to 64’66’93 This evidence is within a few tenths of a kcal. strongly supportive of the proposed mechanism in that, at least for (III), (IV) and probably for (I), (II), the rate determining step for cis—trans interconversion must be attributed to rotation over the steric barrier in the Cotton-Roberts mechanism. It is also interesting to note a that the large excursions in AH with ligand size are accompanied by parallel changes in activation entropy (A85). Inspection of molecular models for cis-(IV) reveals that the two ligands may not freely rotate against each other. The phenyl ring protons in the most eclipsed con- former would overlap by gg. 1.3 A, and indeed, would still interact strongly in the rotamers in which the ligands are gauche. Passage to the trans rotamer would allow free ligand rotation and could account for the increased entropy. The pmr spectrum of the trans bridged isomer of (IV) and (VIII) does in fact contain a collapsed A3 signal for all three protons on the cyclopentadienyl ring, in- dicating that they experience similar, averaged environ- ments. The large variation in activation parameters is not, however, seen for iron system trans isomer carbonyl inter- conversions. Instead, AHI and A85 increase in the series (I) + (III) as alkyl substituents are added to the five- membered ring. This can be understood in terms of a picture developed by Braterman for bonding in these 126 complexes.94 Bridged metal carbonyl bonds are thought of as having, in addition to a metal-metal bond, two three center components, as illustrated in Figure 27. One, 27a, is formed by overlap of the filled ligand 0 orbital with two empty metal 0 orbitals and the second, 27b, is constructed from back donation from two filled metal non- bonding g orbitals to empty ligand 0* orbitals. The three center bonding concept for bridging carbonyl systems gains credence in light of the statement by Cotton that "bridging carbonyl groups never occur unless the bridged metal atoms are formally bonded to each other."95 Electron donating alkyl substitutents on the cyclopentadienyl ring would tend to increase its electron density which might then be transported to the metal carbonyl dative 0 bond, since carbonyl 0* orbitals tend to act as electron sinks. This would make bridge-breaking more difficult and could be the source of the small increase in AHI seen for car- bonyl interconversion of trans (I)-(III). As discussed on pages 45-51 of this dissertation, the chemical shifts of metal-bonded carbonyls are quite sensitive to the inductive properties of the other metal- bonded ligands. The carbon-l3 nmr data evidence downfield shifts for both bridged and terminal carbonyls in the sequence (I) + (III), (V) + (VII), and an upfield shift is observed for the indenyl derivatives. The latter shift 96 is not unexpected since it is known that the indenyl 127 [no 9 4?, O Figure 27. Description of a bridging M~C0 bond. (a) illustrates the overlap of a filled CO 0 orbital with two empty metal 0 orbitals; (b) illustrates the back donation from two filled metal g-orbitals to empty 0“ CO orbitals 128 ligand is slightly electron withdrawing. Furthermore, the chemical shift of the decomposition product of the decamethyliron dimer, [(nS—(CH3)5C5)Fe(CO)2]2, almost certainly the chloride, (ms-(CH3)5C5)Fe(CO)2C1, is found at 214.1 ppm, which is the lowest field terminal iron carbonyl observed in this laboratory. The systems in this study do exhibit carbonyl chemical shifts that reflect the in- ductive ability of the cyclopentadienyl ligand. The observed shifts for (I)-(III) are consistent with our interpretation of the effects of alkyl substitution on rates of carbonyl interconversion. By using the thermodynamic data obtained for the observed structural interconversions, a schematic repre- sentation of enthalpy versus reaction coordinate may be constructed for the iron system (Figure 28). Our work provides clarification of the several already published, differing portrayals63’64’93’97 of these processes in that AH3, the barrier to rotation of non-bridged conformers, is shown to be sizable and to determine the rates of cis- trans isomerization for (III), (IV), and probably (I), (II). The figure is rigorously correct only for (I). The enthalpy difference AH2 between the bridged and non-bridged isomers was obtained for (I) and (V) by Noack56 from ir studies. Manning and coworkers reported that ir bands assigned to non-bridged species of (I) and (V) may also be observed for (II), (VI), but that in- 58-60,98 tensities are sharply reduced. No such bands could N_ononmAmzmu-mtv_ CH c00mcm>coduosc0 Hapsuusth pow mum50p0ooo c00uomou .m> >mrmce 0m00cm0oa mo Empwmmc 000mEm£om .wm murm0m 129 130 be seen for (IV) or (VIII). Our ir results also follow this pattern, as only barely detectable non-bridged isomer absorptions were seen for complex (VII), and these absorp- tions are solvent polarity dependent. It is apparent that alkyl substitution on or aromatic fusion with the cyclopentadienyl ring decreases the stability of non-bridged conformers. This indicates that AHZ values increase somewhat in the sequences (1) + (IV) and (V) + (VIII). Alkyl substitution in the iron systems certainly increases both the enthalpy for the cis-trans equilibrium (AHI) and the barrier to trans isomer carbonyl interconversion (AH6). At least a portion of the increase in AH6 can probably be attributed to the greater AHZ, since AH4 values are estimated to be < 0.2 kcal/mole.64 Barriers to bridge opening are assumed to be identical in both isomers in Figure 28. Results of our nmr investigations of the nS-dienyl- ruthenium carbonyls reveal a striking difference in their solution behavior as compared to the iron systems in that only one bridge-terminal carbonyl interconversion may be observed in the carbon-13 nmr spectra. Concurrent, tempera- ture dependent proton nmr studies of (VIII) demonstrate that bridge-terminal carbonyl interconversion and cis- trans isomerization occur simultaneously. The presence in solution of sizeable quantities of non—bridged isomers for (V)-(VII) again suggests these species as intermediates in the structural interconversion processes and renders 131 probable the mechanistic pathways described above. Comparison of activation parameters for cis-trans isomerization (cis data for the iron systems and the ruthenium parameters in Table 13) shows that for the ruthenium systems, the bulkier ligands change AH]6 and A85 appreciably, but not nearly so dramatically as for iron. Increased ASI values still indicate some release of steric compression in the transition state, but the barrier to rotation of non-bridged tautomers is markedly reduced. The increase of AHI in the sequence (V) + (VII) in fact parallels similar values for trans isomer car- bonyl interconversions of the iron homologs. Alkyl sub- stitution effects on rate for (I)-(III) parallel (V)-(VII). The above evidence and the observed simultaneous isomeriza- tion and carbonyl interconversion suggest strongly that bridge Opening to non-bridged rotamers is the rate de- termining step for the ruthenium system structural inter- conversions. A reaction coordinate diagram may be constructed to portray the energetics of structural interconversions for (V)-(VIII). The actual thermodynamic values repre- sented in Figure 29 are those for (V) and include our measured value for AHZ. While precise values for AH1 could not be obtained, proton nmr spectra for (VII) and (VIII) leave little doublt that both cis and trans isomers are present in about the same quantities and have the same solvent polarity dependencies as measured for the A>0 NHNAcuvzmfimImUImgc_ c0 co0ccm>coo0o0CH 0&0303300m new cLoC0C0ccc cowuommu m: >wuwcw Hawucmuca mo Empwm0p owumfiocum .mm m05w0m . . . 0 — M014 4 . ONHNIQ /3. >3.\ Q 81.0.0 /\ \ ..... 1.. fl 3a >3. ONHOIQ 133 homologous iron systems. The exact barrier to rotation AH5 cannot at present be determined, but for simultaneous isomerization and carbonyl interconversion to occur, it must certainly be smaller than the enthalpies required for bridge opening in both isomers. For purposes of the diagram, equal activation enthalpies for opening cis and trans isomers were assumed. Now that solution structural interconversion studies have been performed for a number of isoelectronic (ms-dienyl)2M2(L)2(L')2 complexes (M = Cr, L = L' = NO), (M = Mn, L = NO, L' = C0), (M = Ru, Fe, L = L' = C0), as. well as for numerous iron systems in which one carbonyl was replaced by an isocyanide or a phosphite, it seems profitable to consider more carefully the exact nature of the barrier to rotation of the non-bridged conformers by examining the structural and activation energy data in Table 15. The tabulated Ea values for [(n5C5H5)Mn(CO)(NO)]2 (XI) and [(nSC5H5)Cr(NO)212 (XII) were obtained by a more detailed consideration of previously reported nmr data93 in terms of the Cotton-Roberts mechanism. In their analysis of the structural interconversions in solutions of these complexes, Marks gt gi. proposed that direct cis- trans isomerization could be effected directly through the non-bridged intermediate obtained by concerted, pairwise bridge opening of either isomer. Rather, rotation about the M-M bond is also required, as has been outlined above 0063 0000 m .mump mm 00 mocmuommm p .mmpauosuum zmunx ou modemummmm o .0xmu mom .mmepomcoo pmcmao mwp00£ cw UCOfl HmumE-Hmuoe ecu unoem £000m000 pom AHmoxv wumcm cowum>wuu< c .pcmeH wC0wU00c u m .HmuoE n E m ,4 - 1:] OJ 11. -I II omm 00.0 + w.0 . 0.0w m0.m A>V :AouvdmA -m mo: ch. 600 00 0 0 0 00 0 00 000.0 000 00000000000000-000. 0 600 0.0 H 0 00M 0 00 000.0 0000 00002000000200 00 o- 000. 0 600 0.000 0.00 000.0 00000 000200000 00 0- 000. 000 000 02-0-20 2-20 02:00:00 maxcocumo Home: H>mepumc oflcouuoonOmH 0000300 mucm0m> mEom pom muma xwumcm 5000m>00o< paw 000300500m .mH 00309 135 for carbonyl interconversions. Similar considerations apply to the pmr data for complex (XII). Several factors which determine the magnitude of the barrier to rotation of the non—bridged conformers can easily be extracted from Table 15 and Figures 28 and 29. Both the enthalpy difference between bridged and non- bridged structures (AHZ) and the size of the steric barrier (included in AHS) are important. The AH2 value for complex (I) is nearly twice that of (V), which partially accounts for the decreased rotational barrier observed for the ruthenium systems. The barrier to rotation AH3 was measured, and was seen to increase from 7.1 to 28.8 kcal/mole upon substitu- tion of indenyl for the cyclopentadienyl ligand in the iron system. Probably only a small portion of this dif- ference can be attributed to AHZ, so there is at least an SE- 15 kcal/mole increase in the steric barrier. Yet, in the ruthenium complexes, upon similar substitution, less than 7 kcal total increase in E8 is observed, and a portion of this, at least 1-2 kcal, must be due to the larger AHZ. This marked difference between the iron and ruthenium systems seems surprising at first glance because the change in M-M bond distances from complex (I) to complex (V) is only 0.2 A. It may be that the Ru-Ru bond is sub— stantially longer in (VIII), but this seems highly un- likely. At present, its crystal structure has not been determined. Alternatively, the steric barrier may simply l 4 .1 | 'I ' 7‘ a.» 136 be unusually sensitive to even small changes in M-M bond distances. The M-M bond distances for the non-bridged complexes [(nS-C5H5)M(CO)3]2, M = Cr, Mo, W, decrease as Cr(3.281 A) > Mo(3.235 A) > W(3.222 A); yet, changes in A0398, sizeable Cr(12.l) < Mo(lS.O) < W(16.2).99 Since bond distances measured for both cis and trans (I) are identical to within experimental error, it is unlikely that non-bonded repulsion interactions used to explain the above Group VIB carbonyl data can be important for (I)—(VIII). Steric considerations alone certainly cannot account for the increased barrier to rotation measured for (XII) > (XI) > (I). Bond distances are measured to increase regularly with decreasing metal atomic number, leading us to expect a corresponding, regular decrease in the steric barrier. Increases in strength of metal- nitrosyl versus metal—carbonyl bonds could, as suggested by Marks gt £1.93 account for larger Ea values for trans isomer interconversions, but cannot account for the in- creased rotational barrier, assuming no change in mechanism. Rather, we must suspect that AHZ increases substantially in order (I) < (XI) < (XII). Unfortunately, no AHZ data are available for the latter two complexes. The very large changes in AHZ which would be required to explain the Ea data in Table 15 lead us to speculate that yet another factor may be of importance. The ability of the first row transition metals to form metal-metal dn-d7T 137 bonds is expected to increase with decreasing nuclear charge for isoelectronic systems. The order of dn-dTr bond strength predicted is Cr > Mn > Fe, in the order of decrease in rotational Ea values. An inspection of the d orbitals in the non-bridged conformer reveals that dn—dTT overlap is indeed possible for (I), (XI), and (XII). No evaluation of the possible importance of directional M-M bonds is now possible, if, indeed they even exist. No independent studies relevant to this speculation are evident to us. Further work is necessary to test the validity of this speculation. The iron-nickel dimer (IX) undergoes carbonyl interconversion. In this case, proton nmr spectra are, however, unchanged with temperature, in agreement with ir 98 which indicate the presence of only the trans studies isomer in solution. In this case, a mechanism involving concerted bridge making and breaking is required to effect carbonyl interchange, rather than the Cotton—Roberts mechanism, since the absence of cis-trans isomerization precludes a bridge opened structure and interconversion of carbonyls via the pathways depicted for the other di— metallic carbonyls examined here. A mechanism was proposed by Cotton et al.65 to account for the carbonyl interconversion of the singly- bridged complex (nS-C5H5)2Rh2(C0)3 . As depicted in Figure 30b, this mechanism involves passage of the structure through a triply-bridged intermediate. The 11*: ”*1 Figure 30. 138 Some mechanisms which might effect carbonyl interchange in single—bridged complexes. (a) unbridged intermediate; bridged intermediate; (b) triply- (c) a concerted mechanism 139 140 mechanism preferred by Lewis gt gt.100 for a phosphite- substituted derivative of the rhodium complex is given in Figure 30c; this mechanism involves the concerted opening of the bridging carbonyl and closing of one of the terminal carbonyls. The synchronous mechanism seems to be more believable. From our data we cannot deduce the mechanism by which carbonyl interchange is effected in the iron-nickel dimer but we can note that the activa- tion energy and the other activation parameters for the process are of the same order as those which proceed ztg the Cotton—Roberts mechanism. Finally, a short statement on the chemical shift dependence of the slow-exchange resonances for these complexes would seem in order here. Two experimental observations which eliminate a few speculations are a) the shift is observed both in the presence and the absence of the shiftless relaxation reagent, tttg(acetylacetonato) chromium(III) ; and b) the shift is observed in the non- fluxional nickel-nickel dimer (X). It is knownlOI’102 that bridging carbonyl ligands are better Lewis bases than their terminal counterparts. The downfield shifts of both the bridging and terminal carbonyl resonances, with the bridging carbonyl resonance being shifted to a greater degree, can then be rationalized by postulating that as the temperature is lowered, solvent-ligand inter- actions become stronger. The solvent molecules interact rm- -— \ .- g1 8 141 more strongly with the bridging carbonyls than with the terminal carbonyls because of their basicity. At higher temperatures, the interaction becomes less important be- cause the carbonyl groups begin to scramble. 142 Suggestions for Subsequent Research Some of the developments in the course of this study suggest that its scope be extended. This section outlines some thoughts about further developments along the line of this research project. A carbon-13 and pmr study of the permethylated iron dimer (a carbon-l3 nmr study was attempted during the course of this research) would certainly be in order. From the reported solution ir of this complex,64 it is apparent that the trans structure is maintained in solu- tion. This observation, together with the fact that the. permethylated cyclopentadienyl ring is extremely bulky, should make a study of this molecule quite interesting. Since the complex itself does not undergo exchange with 13CO, it would be necessary to synthesize it using free enriched Fe2(CO)9, which in turn would have to be synthesized from enriched Fe(CO)5. Since this latter 103 this compound reaction can give yields as high as 90%, would not be difficult to synthesize. Another complex of immediate interest is the complex (nS-CSHA-CH(N(CH3)2)z-nS-C5H4)Fe2(CO)4, which we have christened "the tied-back dimer." The crystal 104 reveals a molecule similar structure of this complex in structure to the cis isomer of (I); the molecule is locked into the cis configuration by a two-carbon bridge which links the two cyclopentadienyl rings. A carbon-13 nmr study of the carbonyl region of this complex would 143 prove interesting because, if indeed the carbonyl groups were observed to be scrambling, a modification of the Cotton-Roberts mechanism would be in order. In this study we utilized only one complex with an electron withdrawing ligand, the indenyl ligand. The perchloroiron dimer, which has not as yet been reported in the literature, would be of considerable interest, since the pentachlorocyclopentadienyl ligands would be extremely electron withdrawing. A possible route to this complex is the following: Reduction of hexachlorocyclo- pentadiene with stannous chloride affords l,2,3,4,5— 105 which upon treatment with pentachlorocyclopentadiene, thallium ethoxide at -78° in pentane yields the white pro- duct thallium pentachlorocyclopentadienide. Reaction of this material with Nil2 in acetone yields the green 106 perchloronickelocene. By following the method of Tilney-Bassett for the synthesis of the iron-nickel dimer 82 it may be (IX) by reaction of nickelocene with Fe(CO)5, possible to obtain quantities of the desired product plus quantities of the perchlorinated iron-nickel and nickel- nickel dimers through an analogous route by using perchloronickelocene and Fe(CO)5. Investigations of the possible fluxional behavior of adducts of (I) with Lewis acids such as triethylaluminum are feasible. Reaction of complex (I) with excess A12(CH2CH3)3 yields a red solution from which crystals of 144 I(US-C5H5)Fe(CO)2]2-2Al(C2HS)3 may be isolated.101 The ir spectrum of this complex shows a decrease of 112 cm-1 101 in the stretching frequency of the bridging carbonyl. Of interest here is whether the carbonyl groups can still interchange when a Lewis acid is coordinated to the bridgir@;carbonyls, and whether cis-trans isomerization can .qM mpnwwm .ne- iii-jig... 154 WWI- em J W 422° Figure 35. Variable temperature carbon-l3 nmr spectrum of the carbonyl region of norbornadienetri- carbonyliron(0); the high field peak is CS2 . J « .\ —fl' -93’ “W ' "ABET” l it... Figure 36. Variable temperature carbon-13 nmr spectrum of the ring carbon region of norbornadienetri- carbonyliron(0); the highest field peak is TMS .520 k... -46' Figure 37. Variable temperature carbon-l3 nmr spectrum of the carbonyl region of (SP)Fe(CO)3 157 .30. ' .5. -50. h 402. Figure 38. Variable temperature carbon- 13 nmr spectrum of the carbonyl region of (SP)Ru(CO)3 the highest field peak is CS2 ‘13‘*“b.’-_._-.- _ “ Fl 158 The spectra of the two SP complexes are shown in Figures 37 and 38. The spectra are not of high quality, but it is evident that from a single resonance at room temperature the spectrum becomes more complicated as the temperature is lowered. 159 Discussion The carbon-l3 nmr spectrum of 1,3,5-cycloheptatriene- tricarbonyliron(0) as a function of temperature was re- ported while we were in the process of obtaining similar spectra.120 However these authors were content to end their experiment at the point at which the basal carbonyl resonance was not resolved into two separate resonances. We believed that our nmr system could achieve sufficient resolution to separate the basal carbonyl resonances if indeed the basal carbonyls are nonequivalent. Our con- fidence was justified, as the spectra plainly illustrate. We are unable to speculate any further concerning the mechanism of carbonyl rearrangement without con- current rearrangement of the ring carbons. It appears as if the mechanism involving simultaneous rotation of the carbonyl groups and the diene moiety, accompanied by small bending motions of the diene moiety, can neither be challenged nor fortified by our data. We have merely demonstrated that the three carbonyl groups within this complex are nonequivalent, and that in the future chemists' investigations of similar complexes should include studies at much lower temperatures than have previously been used. The norbornadienetricarbonyliron(O) complex is believed to possess the square pyramidal geometry featured in the previously-discussed complex and yet, even at -122°C, only a single relatively narrow resonance is fir—“7‘ nth-fl 160 observed. In the light of the results obtained from our studies of the cycloheptatriene complex and from pre- vious studies performed by other workers, we can only conclude that this molecule is extremely fluxional even at very low temperatures. The norbornadiene ligand is a 1,4-diene; it can conceivably span the axial-equatorial sites of a trigonal bipyramidal intermediate (the inter- mediate invoked in a Berry-type mechanism) more easily 1“ than can the 1,3-diene moiety of 1,3,5—cycloheptatriene. 1_ Since this complex is exceedingly more fluxional (”more fluxional" in the sense of being characterized by a much lower activation energy) than the cycloheptatriene complex, our data indicate that a Berry-type mechanism is perhaps the mechanism of choice. This mechanism is pictured in Figure 39. The 2-vinylphenyl(diphenyl)phosphine ligand is bidentate; it has been shown to coordinate to the central metal atom through the tricoordinate phosphorus atom and 128 The through the olefin moiety on the styrene ring. iron and ruthenium complexes of interest can conceivably exhibit trigonal bipyramidal or square pyramidal geometries, 128 indicate that the structure but solution ir studies is trigonal bipyramidal. Figure 32 depicts the three geometrical isomers which in principle can exist for these complexes. Bennett gt gt.128 suggested that isomer 32a,b, or c is the structure of the molecule, but they reported that cis and trans proton-phosphorus coupling 161 (D D {Zig‘ —- B A. | C. %) nucleus, it would seem most likely that benzeneselenol is relaxed primarily through the chemical shift anisotrOpy interaction. The chemical shift difference between diphenyldi- selenide and benzeneselenol was found to be 312.8 ppm, with the diphenyldiselenide resonance being the more downfield resonance. The C6HSSe- moiety apparently directs more electron density around the selenium than does a single proton. The chemical shift of the proton directly bonded to the selenium in benzeneselenol, measured at 138 is indicative of the fact that there 1.35 ppm from TMS, is nothing unusual about the electronic properties of the C6H5Se- moiety. In order to determine whether or not the absence of dipole-dipole relaxation in selenium compounds is a generally-occurring phenomenon we examined the selenium spectrum of dimethylselenide, which, although it has no protons directly bonded to the selenium does exhibit 139 An undecoupled spectrum of proton-selenium coupling. dimethylselenide was obtained over a spectral width of 1000 Hz; after 900 transients all seven components of the expected septet could be seen. ZJSe-H is 10.74 Hz. Figure 44 shows the free-induction decay of the unde- coupled spectrum, its Fourier transform, and the selenium-77 nmr spectrum of dimethylselenide obtained by Lardon. The comparison of Lardon's spectrum, obtained Figure 44. 183 (a) Free-induction decay (FID) of the selenium-77 nucleus in (CH3)28e; (b) the Fourier transform of this FID; (c) the selenium-77 spectrum of (CH ) Se obtained 3 2 by Lardon (a) (b) (C) 185 from a spectrometer equipped with cw instrumentation, with the spectrum we obtained vividly demonstrates the superiority of FTNMR techniques. The nuclear Overhauser experiment was performed and, once more, the result is that no significant NOE enhancement was observed at room temperature, where n = 0.036. The spectra are shown in Figure 45. A tempera- ture-dependent T1 study was performed, and the results are given in Table 18 and illustrated in Figure 46. In the case of dimethylselenide the trend in T1 versus tem- perature indicates the dominance of the spin-rotation. relaxation mechanism; as the temperature increases, the T1 relaxation time decreases. From the microwave spectrum 140 it was shown that there is internal of dimethylselenide rotation of the methyl groups as well as overall rotation of the molecule. The barrier to internal rotation was determined to be 1500 i 20 cal/mole. Either the internal rotation of the methyl groups, the overall molecular rota- tion, or a combination of the two processes is capable of providing a pathway by which the nuclear spins can return to equilibrium. Conceivably the latter process is the most important case, since the microwave spectrum exhibits a fine structure attributal to the coupling of the two rotational modes. Our findings from this series of experiments with Selenium-77 nmr are 1) the nuclear Overhauser effect is, in the molecules we have examined, not of major importance, Figure 45. 186 (a) Selenium-77 nmr spectrum of (CH3)ZSe with no proton decoupling; (b) spectrum with proton decoupling and in the presence of Cr(acac)3; (c) spectrum with proton de- coupling, no Cr(acac)3 present; note that all the integrals are approximately equal 188 Table 111 77Se Tl Data for (CH3)ZSe T (°C) T1(sec) 5° 5.6 18° 5.0 27° 4.5 (n‘= .036) 40° 3.2 T. (39C) Figure 46. 189 2'07-1-1il-1-ni‘ IO 20 30 40 50 T(°C) Plot of T1 vs. temperature for the selenium-77 nucleus in (CH3)ZSe 2.1 i 190 since relaxation processes other than the dipolar-dipolar mechanism dominate the T1 relaxation. 2) Selenium Tl relaxation times may be quite long; and 3) the use of a shiftless relaxation reagent such as tttg(acetylace- tonato)chromium(III) can effectively reduce long selenium relaxation times. We have demonstrated that the future of selenium nmr should be viewed optimistically, once required instrumentation becomes generally available. Indications are that those chemists interested in selenium will eventually be able to utilize selenium nmr techniques in experiments much easier to effect than carbon nmr experiments, which are at present routine. Attempts were also made in our laboratory to ob- serve magnetic resonanee signals of the two spin -% tellurium nuclei, tellurium—123 and tellurium-125. The isotope of mass 123 has only 93% of the sensitivity of carbon-13 at constant field, whereas the isotope of mass 125 has a sensitivity 12.8 times that of carbon-13 at a constant magnetic field (and is therefore 4.2 times as sensitive as selenium-77!). However, the only tellurium magnetic resonance signal we were able to obtain was the tellurium—123 resonance of dimethyltelluride. Its free- induction decay and its Fourier transform spectrum are 2 shown in Figure 47. The JTe-H was measured to be 17.08 Hz. The spectrum shown resulted from 1000 transients, and even with this number of scans only five of the components of v (b) Figure 47. The tellurium-123 nmr spectrum of (CH3)2Te; (a) the free-induction decay; (b) the Fourier transform spectrum of (a) 192 the septet were resolved. The extreme air sensitivity, coupled with the truly unbearable stench of this com- pound, prompted our abandonment of further experimenta- tion with this compound. We were unable to find the tellurium—125 resonance for this compound, and as stated previously, our attempts to locate tellurium signals of either isotope in other tellurium compounds were fruit- less. CHAPTER 5 THE EFFECT OF TRIS(ACETYLACETONATO)CHROMIUM ON CARBON-13 T1 RELAXATION TIMES Background This chapter deals in a rather limited fashion with the effect of an added shiftless relaxation reagent upon the T1 relaxation times of carbon nuclei. Initial use of tttg(acetylacetonato)chromium(III) in carbon-l3 nmr was focused on its ability to undermine the nuclear Overhauser enhancement observed when protons coupled to carbon atoms are decoupled.87'134»141-142 This observed "short-circuiting” of the nuclear Overhauser effect is attributed to an electron-nuclear dipole-dipole relaxation mechanism dominated by the large magnetic moment of theelec- tron (see Equation 57). The electronic magnetic moment is gg. 103 times that of the proton. As the paramagnetic molecules diffuse in solution, their motion induces tre- mendous fluctuating local magnetic fields which in turn rapidly relax the excited protons. Because the proton spins are relaxed as rapidly as they are excited, the carbon nuclei are not influenced by the energy transi- tions. Due to the magnitude of the electronic magnetic moment, the attenuation of the dipole-dipole relaxation 193 194 process by the factor l/r6 apparently does not dominate the influence of the unpaired electrons.. In most in- stances involving organic substrates, all carbon atoms are relaxed to the same degree by the relaxation reagent.141 A dramatic demonstration of the influence a shiftless relaxation reagent possesses is seen in the carbon spectrum of the carbonyl cluster carbide [Rh6(CO)15C]-2. The structure of this cluster compound is a bare carbon atom surrounded by an octahedron of rhodium atoms, which are in turn bonded to terminal and bridging carbonyl groups. The central carbide carbon is buried deeply within the metal—carbonyl framework, and yet a carbon—13 nmr spectrum obtained from a Fourier transform experiment was able to be measured in a few thousand pulses when the solution was doped with tttg(acetylacetonato)chromium(III).143 It has subsequently been discovered, however, that when intermolecular interactions of any sort are present between various carbon nuclei and the relaxation reagent, these carbon nuclei are selectively relaxed at a faster rate. Levy gtgt.144’145 measured the T1 relaxation times of carbon nuclei in noninteracting substrates such as benzene or toluene as well as those in substrates such as phenol, anisole, or benzonitrile, compounds which possess groups which can perhaps attract the relaxation reagent. These workers observed that in various mixtures of these compounds the noninteracting carbons were indeed affected to a lesser extent by the presence of the relaxation 195 reagent than those carbons in the vicinity of the inter- acting site. When the interaction between substrate and relaxation reagent is through hydrogen bonding, the re- laxation times of the various carbon nuclei are observed to vary as l/r6. However, there is no angular dependence 142'146 In principle, then, a relaxation reagent observed. may be utilized to calculate bond distances and to make assignments of resonances. We undertook a study of the carbon T1 relaxation times in mixtures of benzene and cyclohexane as a function of added tttg(acetylacetonato)chromium(III) or tttg- (acetylacetonato)iron(III), in order to determine whether these molecules might indeed be able to interact weakly with the relaxation reagent and in order to compare the relative effectiveness of two different relaxation reagents. The results of this study follow. 196 Experimental Solutions were prepared from degassed benzene and cyclohexane in the following manner: aliquots from a stock solution of 0.1 M relaxation reagent (Cr(acac)3: Research Organic/Inorganic Chemicals; Fe(acac)3: Ventron Corp.) in benzene were measured and transferred ytg pipette and diluted with sufficient benzene, cyclohexane, and hexafluorobenzene (PCR, Inc.) to form solutions con- sisting of 40% benzene, 40% cyclohexane, and 20% hexa- fluorobenzene (by volume) and a known concentration of relaxation reagent. T1 relaxation times were measured by the 180°-I-9O° method described in Chapter 4. Optimum signal-to-noise was achieved with 4 scans per I value, and a total of twenty-seven values and three infinity values (i.e. T = 5 x T1) were used per T1 run. Spectra were recorded and their intensities were directly measured by hand, and the T1 calculation was carried out as outlined previously. The 180° pulse was measured as 70 usec. I." I‘alfor. '.-.r‘jq Ev 197 Results and Discussion The results from this series of experiments are illustrated graphically in Figures 48-53. Plots of T1 versus [Cr(acac)3] for benzene and cyclohexane, Figures 48 and 49, yield curves qualitatively similar to that observed previously for selenium T1 relaxation times and plots of UTI versus [Cr(acac)3] and [Fe(acac)3] are linear, as would be expected. However, the difference in slope between the iron and chromium relaxation reagent graphs, Figures 50, 51 and 52, 53, should be noted. As would be expected from the relative squared ionic magnetic moments, “:ff’ of the two metal ions, the plot of 1/T1 against [Fe(acac)3] shows a steeper slope than the similar plot for Cr(acac)3. It is apparent that tttg(acetylacetonato)iron(III) is a more efficient re- laxation reagent than tttg(acety1acetonato)chromium(III). Were efficiency the only criterion, certainly the iron relaxation reagent would be used exclusively. However, the chromium relaxation reagent is preferred in this laboratory because Cr(III) is kinetically inert, whereas Fe(III) is not strictly so, and because the presence of iron oxide on the surface of some samples which con- tained tttg(acetylacetonato)iron(III) has been noted. The plots shown in Figures 50 and 51 clearly in- dicate that the benzene carbon Tl relaxation times are decreased more rapidly than cyclohexane carbon T1 re- laxation times in the presence of a relaxation reagent. 198 It is believed that this is a manifestation of a weak interaction, probably through hydrogen bonding, between the benzene and the relaxation reagent. The cyclohexane molecule most likely has only a minimal interaction with the relaxation reagent, hence it is less—strongly in- fluenced. Inasmuch as this experiment involved symmetric, cyclic molecules, no further conclusions can be extracted from the data, but certainly experiments involving molecules with many functional groups, some interactive with the relaxation reagent, other noninteractive, should prove interesting. Other possible variations of this experiment include the use of more highly substituted ligands, such as 3,5-pentanedione or 2,2,6,6-tetramethyl- 3,5-heptanedione, for the synthesis of relaxation reagents. These bulkier ligands would increase the distance of closest approach between the paramagnetic metal ion and the carbon atoms in the substrate. 199 Figure 48. Plot of T1 vs. concentration of Cr(acac) for the carbon—13 nuclei in cyclohexane 200 IL l—H .03 .04 [Cr (acac)3] Figure 161 Plot of T1 vs. concentration of Cr(acac)3 for the carbon-l3 nuclei in benzene 201 be» In. l l l l l l J l I .005 .OIO 0|?) .0 20 [Cr(acacl3] Figure 50. Plot of (T1)'l vs. concentration of Cr(acac)3 for the carbon-l3 nuclei in cyclohexane 202 1 1 J 1 I 1 1 1 1 .005 .OIO .0l5 .020 [Cr(acac)-5,] Figure EH_ Plot of (T1).1 vs. concentration of Cr(acac)3 for the carbon—13 nuclei in benzene 203 I 1 l I I .5 - -‘ .4 K: .3 .2 .l l l l l L .00I .0025 .005 .0075 .0l [Fe(acoclil Figure 52. iPlot of (T1)-1 vs. concentration of Fe(acac)3 for the carbon-13 nuclei in cyclohexane 204 .5 .4 I: .3 2 .2 . I L l l l J .OOI .0025 .005 0075 .0l [Fe(acacla] Figure 53. 'Plot of (Tl)-1 vs. concentration of Fe(acac) for the carbon-l3 muclei in benzene CHAPTER 6 DINUCLEAR COBALT CARBONYLS IN SOLUTION. A CARBON-13 MAGNETIC RESONANCE EXAMINATION OF THEIR STRUCTURES, EQUILIBRIA, AND STRUCTURAL INTERCONVERSIONS Background The structure and behavior of C02(CO)8 (XIII) in solution, as well as its structure in the solid state, have been the focus of numerous papers in the chemical literature during the past three decades. In 1947 Anderson147 speculated that the diamagnetism and dimeric nature of C02(CO)8 might be accounted for in terms of a structure with bridging carbonyls and a cobalt-cobalt bond. He based this conjecture on the known structure of Fe2(CO)9, which had been reported eight years earlier.148 149 Syrkin and Kyatkina proposed that, from the standpoint of the electronic distribution within the molecule, there is no reason to eliminate the structure featuring eight terminal carbonyl groups and a cobalt-cobalt bond. The solution structure of C02(CO)8 was subsequently 150 who utilized infrared, studied by Nyholm gt gt., visible, and untraviolet spectrosc0py. The infrared spectrum clearly evidence the presence of bridging carbonyl 7': . groups, as did the ketone-like n + n transition observed in the ultraviolet spectrum. These authors then considered 205 206 O | o C%;\. C£D.—av"C£> ,’E:>CQ7”’(L 0"TT—CK) C- C) C 2v C2 h 0 O o O C: C C C C) C) C\ | CO I I |° I C: C c; C C) C) c) c) D 2h 02h W\\‘\. :,C:\ ,I’CICK> AC0 ”III’ \\\‘ D2h Figure 54. Several structures postulated for C02(CO)8 (XIII) 207 the structure to be one with one of the following symmetry groups: D2h’ C2h’ C2v’ D3h’ and D3d' The infrared spectra they obtained made the choice of the correct structure a difficult one, but these authors preferred the D2h structure. See Figure 54. Wender gtgt.151 reported the infrared spectrum of C02(CO)8 in solution and in the gas phase and concurred that bridging carbonyls are present. These authors con— sidered but rejected the possibility that in solution two forms of the molecule were present. They did obtain direct proof, though, that the bridging to terminal carbonyl, ratio is 2:6 by treating C02(CO)8 with acetylene. The spectrum of the resulting compound, C02(CO)6C2H2, in- dicated that the bridging carbonyl infrared absorptions had disappeared, while the absorptions corresponding to the terminal carbonyl groups remained. The molar absorptivity measured from the infrared absorptions of the bridging carbonyls of C02(CO)8 matched quite well the extinction coefficients observed in organic carbonyl groups. From their data, these authors proposed a structure based on trigonal bipyramidal coordination about the cobalt atoms with the cobalt atoms and the bridging carbonyl groups being coplanar. This structure is shown in Figure 54. The structure of a related acetylene complex C02(CO)7(C4H202) (XIV) (empirical formula C02(CO)9C2H2) 208 was obtained by Mills and Robinson.152 Its characteristic structural feature is that the two metal atoms and the two bridging carbons are not coplanar. In the light of that evidence, these authors suggested that perhaps specula- tions about the structure of C02(CO)8 should be based on the absence of two cobalt atoms ceplanar with the two bridging carbonyls. See Figure 55. In a study primarily designed to test the resolu- tion of their infrared Spectrophotometer, Bor and Mark0153’154 investigated some solutions of C02(CO)8. They considered the structure proposed by Mills and Robinson to be quite likely correct, but stated that infrared spectroscopy pgt gg was incapable of distinguishing between differing structures of C2V symmetry. By using newly developed calcium fluoride optics, rather than the conventional sodium chloride optics, Cotton and Monchamp155 undertook an ir investigation of solutions of C02(CO)8 and were able to attain superior resolution as compared to all previous studies. The observations recorded by Cotton and Monchamp are essentially identical with those of Bor and Marko: the results obtained from ir spectrosc0pic studies reflect the symmetry of the species of interest. As a consequence, ir studies cannot be used to discern differences in structure among molecules with identical functional groups 156 157 and identical symmetry. Bor and Noack then reported the infrared spectrum of C02(CO)8 along with spectra of 209 1‘ —n 3— /; II _\_.<'/ / \\ o \ - — — n o ____n/ ? n n o I; O \ n— I -\ I I / I I O-n~—-_ Figure 55. Crystal structure of CoZ(CO)7(C4H002) (XIV) 210 several other metal carbonyls by using the improved ir optical system. Bor reported that some of the terminal carbonyl bands appeared to be doublets under the higher resolution. A major breakthrough in interpreting the results obtained from infrared studies of C02(CO)8 was announced by Bor in 1963.158 By using the results he obtained from a high resolution infrared study, and by citing the fact that C02(CO)8 exchanges its carbonyl groups quite readily with free 14COlsg’160 he postulated that two configura— tions of the molecule exist in solution. He based his hypothesis on the observation that the intensity ratio of the doublets he had observed earlier is solvent de- pendent. Shortly thereafter Noack161 presented evidence obtained from a temperature-dependent infrared study of C02(CO)8 in pentane. The spectra obtained at room temperature and at -105°C were conspicuously different. Bridging carbonyl absorptions were present at both tem- peratures; however, only three of the terminal carbonyl absorptions remained with their original intensity at -105°C. The other terminal carbonyl absorptions were considerably weakened in intensity. Noack interpreted the results to indicate that two configurations of C02(CO)8 exist in solution in a temperature-dependent equilibrium. The low temperature form possesses bridging carbonyls and the high temperature form possesses only terminal carbonyls, 211 o o 0“ 8 /(’O I :l; \ oc—Co< >Co—CO :co—Co o—C C/ C \ /C o 0 Co 06’ O A 8 Figure 56. Solution equilibrium postulated by Noack and Bor for C02(CO)8 (XIII) 212 as is depicted in Figure 56. He proposed that the high temperature form was linear and similar in structure to the isoelectronic [F€2(CO)8]2’.162 Bor163 immediately and independently announced the same conclusion, based on his analysis of the solvent dependence of the intensities of some of the terminal carbonyl infrared absorptions. His conjectures concerning the structure of the open isomer parallelled those of Noack. The temperature dependent changes in intensity of the terminal carbonyl infrared absorptions of C02(CO)8 in solution were studied in greater detail by Noack.164 From variations of absorption intensities he calculated that the open structure is less stable than the bridged structure by AH = 1.3 kcal/mole, and that the entropy difference between the two isomers is +5 e,u, He also determined that, at room temperature, a pentane solution of C02(CO)8 is comprised of 43% A and 57% B, whereas at ~105°C there is present 84% A and 16% B. During the same period when Bor and Noack were elucidating the behavior of C02(CO)8 in solution, Klug 165 communicated its solid state structure. It 9.221. was found to be a distorted Fe2(CO)9 structure exactly like that predicted by Mills and Robinson; the cobalt atoms and the two bridging carbonyls are not coplanar. See Figure 57. 213 (XIII) Crystal structure of C02(CO)8 Figure 57. 214 The cobalt-59 nmr spectrum of C02(CO)8 in solu- tion has been measured by Lucken gt gt.31 166 and by Haas and Sheline. Only one very broad signal was observed (10.41 gauss in benzene, 6.35 gauss in pentane), rather than two resonances, which would be expected if the two isomers present are distinct and non-interconverting. It was stated by Lucken gt gt. that perhaps the resonance from one of the isomers (and they believed the bridged form was the likely candidate) was too broad to be seen. Recent carbon—13 nmr studies of the polynuclear 67 68 cobalt carbonyl C04(CO)121 and the related RhCo3(CO)12l demonstrate that the carbonyl groups within these molecules are fluxional in solution in the manner predicted by Cotton169 in 1966. The fluxional process in the former compound is stopped on the nmr time scale at -lOO°C and at -85°C in the latter. The only carbon-l3 nmr study of a dinuclear cobalt carbonyl complex to date is the study 170 of (H5C6CCC6H5)COZ(CO)6- a complex which contains no bridging carbonyls. Even at —125°C a single resonance is observed in the carbonyl region, indicating that all terminal carbonyls are undergoing some type of scrambling 171 of some phosphine process. A carbon-13 nmr study derivatives of methinyltricobalt enneacarbonyl, H3CCCo3(CO)8PR3, has established that these cluster com- pounds exist in solution as mixtures of isomers and that they too undergo bridge-terminal carbonyl interconversions. 215 From evidence obtained from carbon-13 nmr studies of polynuclear cobalt carbonyl complexes we concluded that C02(C0)8’ (XIII) should likewise exhibit bridge-terminal carbonyl interconversions of the type shown in Figure 11. In continuing our investigations of bridge-terminal carbonyl interchange processes in dinuclear metal carbonyl complexes, we examined the carbon-13 nmr spectrum of (XIII) and discovered that, even at -164°C, our present experimental low temperature limit, only a single, relatively narrow resonance is observed, indicating that if bridge-terminal carbonyl interconversion does occur, the process is scarcely affected by this low temperature. Several complexes structurally-similar to C02(CO)8 were then utilized in our studies in the hope that the sub- stituents would slow the rate of carbonyl interconversion and therefore make it amenable for study by carbon-13 nmr. The bridge—substituted complexes C02(CO)7(C4H202) (XIV) and C02(CO)7Ge(C6H5)2 (XV) and the terminally-sub- stituted complexes C02(CO)6DPPM (XVI) (DPPM = bis— (diphenylphosphino)methane) and [C02(CO)7]2DPPA (XVII) (DPPA) = bis(diphenylphosphino)acetylene were synthesized for this purpose. These complexes are pictured in Figure 58. The fluxional behavior of these complexes is observ- able in the carbon-l3 nmr spectrum, and the slow-exchange limit could be reached for most of these complexes. The processes through which carbonyl interconversion is effected are novel. 216 ’9 "fr—i 4’ d2 OCHCKC’O (:o o ‘68 Co ‘\ /’ ‘\../ z” - /’ OC-Co— Co-CO 00— Co Co -CO 0 O O o o XIV XV 0c 8 (MP 0C 8 Co \ / \ / \ / \ / OC—Co o—P 2 OC—Co Co—CO / C \ \c/ \ C O C C o P¢2 O O O \ L- C\\\ 42 XVI XVII 0 o c 0 c \ /C\ / OC—Co Fe / \8/ OC Figure 58. The dinuclear cobalt carbonyl complexes utilized in the study of solution dynamics outlined in Chapter 6 217 Experimental Dicobalt octacarbonyl(XIII) and C02(CO)7(C4H202) (XIV) were purchased from Strem Chemicals and were purified by recrystallization from pentane. C02(CO)7Ge(C6H5)2 (XV). The procedure of O'Brien gt gt.172 was used to synthesize the cobalt-germanium complex from diphenylgermane (Columbia Organic Chemicals) and C02(CO)8. The complex was purified by recrystallization from pentane. C02(CO)6DPPM (XVI). The DPPM complex was synthesized from ttg(diphenylphosphino)methane (Strem Chemicals) and C02(CO)8 according to the method of Chia and Cullen.173 Purification was achieved through chromatography with dichloromethane elution and recrystallization from dichloromethane-pentane. [CoZ(CO)7]2DPPA (XVII). The synthesis of DPPA complex from C02(CO)8 and bis(diphenylphosphino)acetylene (Strem Chemicals) was performed according to the procedure 174 established by Ng and Carty. Recrystallization from dichloromethane-ethanol yielded the product. (nS-C5H5)Fe(CO)2Co(CO)4 (XVIII). The method of Joshi and 175 Pauson was used to prepare the iron-cobalt complex from (nS-CSHSFe(CO)2C146 and NaCo(CO)4. This very unstable I material was then recrystallized from dichloromethane- pentane. The dense, reddish-brown liquid (nS-C5H5)C0(CO)2 (XIX), similar to bromine in appearance, was purchased from Ventron Corporation and used without further purifica- tion. 218 13CO Enrichment. Solutions of (XIII) and of (XIV) in dichloromethane were stirred overnight under an atmosphere of 90% carbon-l3 enriched carbon monoxide in order to 160 Enrichment of enrich the complexes to gg. 60% 13CO. (XV), (XVI), and (XVII) was accomplished by synthesis of these complexes starting with enriched C02(CO)8. The unstable complex (XVIII) was enriched under an atmosphere of enriched carbon monoxide overnight at —15°C. Enrich- ment of (XIX) was accomplished in the same manner as described for (XIII) and (XIV).86 All instrumental procedures are identical to those described in Chapter 2 of this dissertation. Solvents employed in the temperature dependent nmr studies were the same as thoseemployed in the study of the iron dimer complexes described in Chapter 2 unless otherwise indicated. All carbon-13 nmr samples were 0.03M in tttg(acetylace- tonato)chromium(III). In retrospect, the addition of the relaxation reagent to the cobalt samples was un- necessary, since the carbonyl carbons are relaxed primarily through interactions with the quadrupolar cobalt nuclei. Nevertheless, addition of the relaxation reagent does not alter the spectral results. 219 Results Carbon—l3 nmr spectra have been obtained for the six dinuclear cobalt carbonyls (XIII + XIX) over a tem- perature range +50°C to -164°C. Each of these molecules is observed to be fluxional and at higher temperatures, since all bridged and terminal carbonyls are seen to be completely averaged. For complexes (XIII) and (XVIII), a single averaged resonance was observed even at the lowest temperatures, as shown by spectra pictured in Figures 59 and 66. For complexes (XIV)-(XVII), however, the Spectra were observed to change markedly with tempera- ture, as illustrated in Figures 60, 61, 63, and 65. In the low temperature limit of each set of spectra, carbonyl resonances were detected in the region 229-241 ppm down- field of TMS and in the region 193—203 ppm from TMS. The resonance of complex (XIX), which has only terminal car- bonyls, was detected at 204.6 ppm. We therefore assigned the resonances in the downfield spectral regions to the bridging carbonyls by basing our assignments on the magnitude of the observed chemical shift of the terminal carbonyls in (XIX) plus observations that 1) integrated intensity ratios become meaningful when this assignment is made, and 2) in other dinuclear metal carbonyl systems examined thus far, the far downfield resonances may always be assigned to the bridging carbonyls. A single carbonyl resonance was measured for octacarbonyldicobalt in dichlorofluoromethane even down Figure 59. 220 Variable temperature carbon—13 nmr spectrum of C02(CO)8 (XIII); the higher field peak is C82 8. .22. -57° ~85° -415’ 2221 3 v1.17 :‘I‘N~ 5.45 fr" . o, 1 n ,q 1' 81’ W ' I” .g-.' " 222 Table 19. Chemical Shift and Linewidth Data for C02(C0)8 (XIII) T (°C) 6CD 02° +8° 201.8 43.1 -20° 202.5 27.7 -57° 203.2 12.7 -85° 203.2 8.8 -115° 203.9 8.3 -127° 204.1 7.3 a full-width at half height in Hz :n‘n— — hr” 223 Figure 60. Variable temperature carbon-13 nmr spectrum of the carbonyl region of Coz(CO)7(C4H202) (XIV); the highest field peak is CS2 n-13 titsaf eak is 55 222211 4° -23° ~33° A k ' -53° 400° [Ix-4' Figure 60. 223 Variable temperature carbon-l3 nmr spectrum of the carbonyl region of C02(CO)7(CAH202) (XIV); the highest field peak is CS2 1354 a U) (*3 Li) I l r 7“ w N' 5.5,... r‘ \ '1‘". -O!_H .V / ~‘-“‘ . "b' 224 ~ 26° 40 v v 400‘ 225 Table 20. Chemical Shift and Linewidth Data for C02(CO)7(C2H402) (XIV) T(°C) 6average 6bridge 6terminal v%a +47° 200.4 61.6 +26° 200.3 . 54.9 +4° 200.7 85.4 -5° gg.200.5 140.0 -12° 192.0 -18° 134.0 -20° gg.198.9 -23° gg.l98.0 102.0 -30° gg,196.9 97.1 -37° complex -42° complex -44° complex -52° 193.6b 229.8 197.0 -590 193.5b 230.0 197.2 -64° 193.5b 230.1 197.1 -69° 230.1 197.1,193.5 -78° 230.5 l97.2,l93.6 -89° 230.6 197.2,193.6 3o.5°,25.0,36.6 -100° 230.3 196.7,196.2,193.1 19.5°,ca 18.3 a full width at half height in Hz average of four terminal carbonyls C width of bridging carbonyl resonance _]02. M It...» Lt... . A‘— ‘ 1,. “W—-';: ‘4‘: *2... .132 -82' m A: —‘;—‘r“~ w___ v 54-145. Figure 61. Variable temperature carbon—l3 nmr spectrum of the carbonyl region of C02(CO)7Ge(C6H5)2 (XV); the highest field peak is CS 2 227 NxmmcoccuerovN quOHl Um A>xv 00 W0 COHwNM H%COQHQU NEH MO Eflhuuwgm REC MHICODHQU .Nc champs Table 21. T (°C) +4° -l8° -38° -62° -82° -93° -102° -114° —122° -125° -132° -145° -l64° a Chemical Shift and Linewidth Data for (XV) C02(CO)7Ge(C6H5)2 Gaverage abridge 204.5 204.6 204.8 205.0 complex complex complex complex complex complex 237.4 236.8 gg.237.0 228 full width at half height in Hz 0 C N\*‘ terminal 200.1 200.1 gg.200.0 34.8 21.4 17.1 14.6 28.7 K.) N \D I I I 4L1 Figure 63. Variable temperature carbon-l3 nmr spectrum of the carbonyl region of highest field peak is CS2 t i CC 405' 230 o.qca- he AH>xv .mflmmcoveNmoemflmmcov.cfloovmoo mo cowwmh ahconumo mnu mo Eduuomam MEG mH-copnmo .38 605042 231 AH>xv o.NmH o.woa m.oq e.mm .em.m2 2meacAoecNoo Hmafihmuc m.oqm m.o¢m H.00N ©.mmm H.omm m.wmm o.wm~ weeps w no.mHN nm.HHN no.mHN nm.mHN Q¢.MHN aq.mam n¢.mam 0.0HN m.cHN m.oHN m.©HN m.oHN n.0HN mwmnm>m 060 meta 200036602 cam emetm 26605620 omNH- cowa- omoau 6mm- 6mm- com- amo- 6mm- enq- oom- ema- om- 60+ Auov H .mm cfieme 232 mazcoppmo HmcHEumu Hmum>mm LDHB ammonumo wcflwpflun mco mo mwmum>m m.meH .cm.me m.Hom.c.mON Hmcfieumuw Hmconumo wcflwcwun 0 mo M> em as etwect was; 06 teens 23:0 6 «.mmm.mm .m.oem.mm N.mmm.00 .H.Hem mm H.46N emceece onQEoo nm.mom mwmum>m@ A.ucoov o¢©HI ommHI OWN-HI 8L H mm manme 233 or ...;\...‘_".p._,/ILILfl"¥65’ 40' =24 J 42‘ v—Vvv 24' 407° 4? MI.- 424' LLLIEII t L 424' -53' Figure 65. Variable temperature carbon—13 nmr spectrum of the carbonyl region of [C02(CO)7]2[(C6H5)2PCCP(C6H5)2] (XVII); the highest field peak is C82 Table 23. T (°C) +63° +40° +24° +5° .70 -21° -36° -53° -65° -80° -96° -107° -124° a CHFCl2 234 Chemical Shift Data for [CoZ(CO)7]2 DPPA (XVII)a 5average 6bridge éterminal 222.1b 206.2b 213.9C 197.4C 214.0C 197.6C 214.7 197.8 212.8 197.4 211.3 197.5 210.6 197.5 211.1 234.7 197.6 211.3 234.8 197.6 235.4 197.5 236.0 197.6 236.7, 2 2.4 complex 236.9, 2 2.6 206.4, 200.0, 199.2, 197.9, 193.8 solvent unless stated otherwise Tetrahydrofuran solvent c CH2C12 solvent 235 Figure 66. Variable temperature carbon-13 nmr spectrum of the carbonyl region of (nS-C5H5)Fe(CO)2Co(GD4 (XVIII) in dichlorofluoromethane; the highest field peak is CS2 13 91111 91.21 . 1112.90.11 3.5L”. 236 I b51- LI 5L. O A — w v— 8' -31' 4o Figure 66. 235 Variable temperature carbon-13 nmr spectrum of the carbonyl region of (nS-C5H5)Fe(CO)ZCo(CO)4 (XVIII) in dichlorofluoromethane; the highest field peak is CS2 It... I... .111 _ Ml... .. .AVJ @410 L125. ,WLL I ' A... ....J .w. wtg‘tv ‘ 237 Table 24. Chemical Shift Data for I(HS-C5H5)Fe(CO)2Co(CO)4] (XVIII) in Various Solvents Sgttggt t_t:tt 6C0 (average) CHFCl2 +8° 211.9 -6° 212.7 -18° 213.0 -31° 213.5 -40° 214.1 -53° 214.8 -63° 215.5 -81° 216.3 -96° 217.5 -110° 217.9 -115° 218.5 -125° 218.6 CH2C12 +8° 212.1 -37° 213.7 -60° 214.7 CHCl3 +8° 212.4 -55° 214.6 C32 +8° 210.0 -30° 210.4 —45° 210.7 —64° 211.2 -87° 211.9 CHZBr2 -46° 213.9 Acetone +8° 212.8 -22° 213.5 -60° 214.7 238 Figure 67. Ambient temperature carbon-13 nmr spectrum of the carbonyl region of (ns-C5H5)C0(CO)2 (XIX) the high field peak is CS2 239 to -164°C. This resonance was observed to undergo a measurable chemical shift as the sample temperature was lowered; at 24°C, the signal is located at 201.8 ppm, whereas, at -130°C, the resonance was observed to have shifted 2.2 ppm downfield. As tabulated in Table 19, the linewidth of the carbonyl resonance decreased appreciably with decreasing temperature. The crystal structure of (XIV), determined by 152 shown in Figure 57, shows that each Mills and Robinson, cobalt atom is bonded to three terminal carbonyl groups with the Co-Co bond being spanned by one bridging carbonyl and by the five-membered lactone ring, with the C02(CO) (lactone) ring being non-planar. The low temperature carbon-l3 nmr spectra of (XIV) are consistent with this structure (see Figure 60 and Table 17). Below ~100°C, three resonances which may be assigned to the three chemically-different terminal carbonyls are observed at 196.7, 196.2, and 193.1 ppm. A bridging carbonyl reson- ance, with only 50% of the integrated intensity of each of the terminal carbonyl resonances, is detected at 230.3 ppm. As the temperature is raised, the two highest field terminal resonances are observed to broaden, whereas the third terminal carbonyl resonance apparently does not undergo any such broadening at this temperature. Con- currently, the bridged carbonyl resonance broadens and, at gg. -53°C, the two terminal resonances are averaged 240 and the bridged carbonyl resonance is quite broad. Near -30°C the remaining unaveraged terminal carbonyl begins to take part in the averaging processes, and near room temperatures, only a single resonance is observed at 200.3 ppm, which is the precise weighted average of the four low temperature resonances. Apparently there are at least two distinct carbonyl interchange processes de- tected; there is a high temperature process which averages all seven carbonyls, and a low temperature process which averages only four of the terminal carbonyls with the single bridging carbonyl. Only a single distinctive feature was observed in the proton nmr Spectrum of (XIV). Two doublets, with chemical shifts at 6.01 ppm and 7.48 ppm, and a coupling constant of 5.0 Hz, were observed at ambient temperature, and these signals remained virtually unchanged as the temperature was lowered. 172 and from From reported ir spectral evidence, the nmr data reported herein, there is sufficient evidence that the structure of complex (XV) is that shown in Figure 68, a structure which is analogous to that determined for (XIV). The second bridging ligand is a diphenylgermanium moiety, and the four-membered ring is assumed to be non- plgnar in the complex. The carbon-13 nmr spectra of the carbonyl region of this complex are shown in Figures 61 and 62. At -l64°C, two resonances of exactly 1:6 intensity are detected at 200.1 and 236.7 ppm, respectively. Again Figur- 241 Figure 68. Probable str ‘t ° uc urc of C02(L.O)7Ge(C6H5)2 (XV) the downfi carbonyl. able to t1 carbonyl 1 that even which is . viscosity more than resonance A carbonyl begin int resonance the termi imPOSed - becomes I resonanca terminal exchange SPEetruH 242 the downfield resonance was assigned to the bridging carbonyl. That there is only one large resonance attribut- able to the terminal carbonyls implies that the terminal carbonyl resonances may be coincidental, or, more likely, that even at -164°C there is some process in operation which is averaging the terminal carbonyls. Solvent viscosity at this temperature limits resolution to no more than gg. 10 Hz, as measured from nmr lock carbon resonance linewidths. As temperature is raised to —l32°C, the bridging carbonyl and four of the terminal carbonyls are seen to begin interchange. At -125°C, the bridging carbonyl resonance is completely broadened into the baseline, while the terminal region consists of a very broad signal super- imposed upon a sharp resonance. At -ll4°C, when averaging becomes more rapid, the broadened terminal carbonyl resonance is seen to shift out from under the remaining terminal resonance, which now is just beginning to undergo exchange broadening. The computer simulation of the -102°C spectrum, Figure 69, demonstrates that the intensity ratio of the two signals at 206.6, 200.6 ppm is 5:2. As the temperature is raised further, the two broad lines broaden further and eventually coalesce at -62°C into a sharp signal at 205.0 ppm which remains unchanged in higher temperature spectra. The chemical shift value of the averaged resonance observed above -62°C is very near the weighted average of the low temperature limiting re— sonance S . 243 Figure (El Computer-generated carbon-l3 nmr spectra of C02(CO)7Ge(C6H5)2 (XV) at -102° C; (a) peaks in 4:3 ratio; (b) peaks in 5:2 ratio; (c) measured spectrum 244 E 3.. . I I I@ :tra 3f 3) 383315 245 In contrast to complexes (XIV), (XV), which are bridgesubstituted derivatives of C02(CO)8, complex (XVI) is formed by diphosphine substitution of two terminal carbonyls. This complex was first synthesized by Chia 173 in order to contrast the structure of a and Cullen molecule containing a chelate with a small "bite angle" to the structures of molecules containing chelating ligands which had longer chains. As depicted in Figure 70, there are two possible isomers which can be visualized; one, 70A has the chelate ligand coordinated in a symmetric fashion. This isomer therefore contains four chemically distinct carbonyls, with two equivalent bridging carbonyls. Isomer 70B, on the other hand, places the chelate un- symmetrically, thus rendering all six carbonyls non- equivalent. The present carbon-l3 nmr study indicates that only one isomer, B, is found in solution. At -164°C, the lowest temperature attainable with the present temperature controller, the spectrum shown in Figure 64 was observed. In the bridging carbonyl region, there are two discernable resonances; the re- sonance at 253 ppm appears to form a doublet, and the re- sonance at 240.5 ppm appears to be a poorly defined triplet, 2 with J = 8.5 Hz. The separation between the "doublet” C-P peaks is nearly 100 Hz. Two broad lines centered at 206, 201 ppm are observed in the terminal region of the spectrum. Integrated intensities reveal a bridgedzterminal 246 O C P P P P CO V A B Figure 70. Possible isomers of C02(CO)6[(C6H5)2PCH2P(C6H5)2] (XVI) O P C (3C:ZE!E::CI) (3f;:§!gz:CC) 0C CO OC p A 8 Figure 71. Possible isomers of [C02(CO)7]2[(C6H5)2PCCP(C6HS)2] (XVII) 247 carbonyl ratio of 1:2. Upon increasing the temperature to -153°C, the terminal carbonyl resonances are seen to broaden. At —l38°C, the downfield bridging carbonyl resonance has broadened beyond the limits of detection, whereas the upfield bridging carbonyl resonance is main- tained. The broad, averaged terminal carbonyl resonance is seen at -125°, -120°C to broaden and shift downfield, an effect due undoubtedly to a process which averages it with the bridging carbonyl. The resonances eventually sharpen until, at -95°C, the bridging carbonyl resonance, now a well-defined triplet, and the terminal averaged resonance are seen at 243.6, 213.5 ppm, respectively. The signal at 213.5 ppm is very near the average position of the one bridging and two terminal carbonyl sets seen in the -164°C spectrum. Saddling the large averaged resonance are two small lines, which, if they are the two outer components of a doublet of doublets from splitting of the terminal carbonyl on the phosphorus substituted cobalt. The large averaged resonance is not precisely between the two satellite peaks. Spectra obtained at successively higher temperatures are characterized by the collapse of the multiplet structures of both resonances, broadening of the remaining bridge resonance with concurrent broadening of the averaged resonance, and finally coalescence of the two resonances into a single resonance which then sharpens as the temperature is raised. At ambient temperature the resonance appears at 216.9 ppm, which is, within 248 experimental error, the weighted average chemical shift of the low temperature spectra. The structure of complex (XVII) is as yet unknown, but ir studies174 indicate the presence of both bridging and terminal carbonyls. The structure pictured in Figure 58 is that suggested by the workers who initially synthesized the complex. As was the case with complex (XVI), the possibility of two different isomers must be considered. The phsophorus atom Of the DPPA ligand may substitute for a terminal carbonyl to give a symmetric isomer 71A or an asymmetric isomer 71B, when the steric requirements of the bulky diphenylphosphine group are neglected. The present carbon-13 nmr study indicates the presence of only one isomer, the asymmetric one, 718, and inspection of a molecular model of this complex reveals that interaction of the diphenylphosphine moiety with the bridging carbonyls establishes serious doubt that isomer 71A could be formed. The spectra of complex (XVII), collected in Figure 65, clearly indicate fluxional behavior. At -124°C, seven resonances are detected, thus indicating the presence of only 71B. The bridging carbonyl resonances are seen at 242.6 and 236.9 ppm, and the terminal region consists of resonances at 206.4, 200.0, 199.2, 197.9, and 193.8 ppm. The bridging carbonyl resonance at 242.6 ppm and the terminal resonance at 206.4 ppm are the lowest field resonances thus far observed for cobalt carbonyls. As the temperature is raised, the downfield bridge broadens 249 and disappears, and the terminal resonances begin to broaden and coalesce, until, at -96°C, only two resonances are seen at 236.0 and 197.6 ppm. The terminal resonance continues to sharpen, and at -65°C, it appears to be quite narrow. A new resonance appears at gg. 211 ppm at this temperature; it begins to sharpen with concurrent broaden- ing and ultimate disappearance of the bridging carbonyl resonance, until, at -36°C, only two resonances at 197.5 and 210.6 ppm are detected. As the temperature is in- creased, the downfield resonance is seen to broaden and to reappear at 214.7 ppm, whereas the upfield resonance begins to broaden. At +5°C in dichlorofluoromethane solvent, the two resonances are found at 197.8 and 214.7 ppm. A change in solvent at this point dramatically affects the chemical shifts of the carbonyl resonances; at +24° in dichloromethane, the resonances are shifted up- field 0.3 ppm, and in tetrahydrofuran at +63°C, the reson- ances are shifted over 8 ppm upfield. At this temperature in tetrahydrofuran, the downfield resonance is seen to be narrow, in contrast to the broad appearance of the up- field resonance. Based on the results of a crystal structure determination of the complex (nS-C5H5)Fe(CO)2Co(CO)3 176 P(CH3)(C6H it can be surmised that the structure 5)2: of complex (XVIII) is characterized by a nonplanar Fe(CO)2Co ring, with the cyclopentadienyl ring trans to 250 the puckering. As indicated in the spectra shown in Figure 66, no resonances due to bridging and terminal carbonyls were observed at the lowest temperatures, and moreover, the single averaged resonance observed is still very narrow at the lowest temperature. The spectra are characterized by a relatively broad resonance which exhibits a temperature-dependent chemical shift and a narrow resonance which (in dichlorofluoromethane) remains chemical shift invariant at 212.4 ppm. It is not known at this time whether the narrow resonance is due to a non- scrambling terminal iron carbonyl in the complex, or an impurity with a terminal iron carbonyl. Since the low temperature limit was not reached, the assignment of this resonance is not at this time of utmost importance. Since it has been demonstrated by ir studies that this molecule exists in solution in an equilibrium of the type depicted 177 it was decided to examine the effects of in Figure 11, different solvents on the chemical shift of the averaged carbonyl resonance. The results of this study may be seen in Table 24. At a temperature of +8°C, it can be seen that the resonance is shifted upfield in the less polar carbon disulfide solution, (210.0 ppm), and that it shifts downfield slightly when more polar solvents are used. Interestingly enough, there seems to be no correlation between the downfield shift and solvent polarity; in chloroform, where e = 4.81, the chemical 251 muco>H0m mo zuowum> m Cw quovooNAoovomAmmmu-ch How ousumuanou .m> Amwmuo>mv 000 86:. ON 0 ON. 9v- Om u 00.. 00 7 ON.- n d _ _ . _ _ _ _ «2:32.698 4 .I OCC.-bud I 228562265 0 ocufioEEoEuB . 53825 O 85:23 0 p b _ _ — — _ _ CNN AHHH>xv mo no.3 .Nm mudwfim 0.!- _ L EN 1 N_N I. MM§ ¢;mw mfl A a m 6 SN m- 9N 252 shift value is 212.4, but in acetone, with e = 20.70, the resonance appears at 212.8 ppm (6 is the solvent dielec- tric constant). Figure 72 is a graph of chemical shift versus temperature for this complex in a variety of solvents. One distinctive feature of all the cobalt carbonyl spectra is that, at room temperature, the single, averaged resonance is very braod, and as the temperature is lowered, the resonance sharpens somewhat before it commences broaden- ing due to the slowing down of the carbonyl exchange pro- cesses.’ Both types of broadening are due to the scalar re- laxation interaction, with the former referred to as scalar relaxation of the second kind and the latter as scalar re- laxation of the first kind. Since the cobalt nucleus is relaxed extremely efficiently by the quadrapolar relaxation interaction, it follows that T1 = T2 = TS, where TS is the correlation time which modulates the relaxation rate of the cobalt nucleus. The expression which describes the scalar relaxation interaction interaction in terms of the relaxation rate of spin I coupled to a spin S is R = R2 = (AZ/3)S(S + 1)-S + TS/l + (ml - wS)2T§, (75) where A is the spin-spin coupling constant in units of radians per second. In this discussion, the spin S re- fers to the cobalt nucleus (S = 7/2) and spin I refers to carbon. From Equationfifl) it is apparent that as R; decreases, the width of the resonance decreases; hence, as the temperature is lowered, TS, or the relaxation time of the cobalt nucleus, must be decreasing. 253 By recalling Equation 58, (where I now refers to the spin of the cobalt nucleus), it is seen that the three variables which may affect the magnitude of RQ are the asymmetry parameter, n, the quadrupole coupling constant (or more precisely, the electric field gradient), or the molecular correlation time, TC. Since neither the asymmetry parameter nor the electric field gradient would be expected to vary significantly with temperature, we can be confident that the relaxation time decreases with decreasing temperature as the result of an increasing molecular correlation time. That the molecular correlation 178 time is temperature-dependent is known, and the de- pendence is given by TC = Toexp(Ea/RT), (76) where E8 is the activation energy of the reorientation of the molecule under study. Therefore the sharpening of the averaged resonances with decreasing temperature arises from the temperature dependence of the quadrupolar relaxation rate of the cobalt nucleus, which in turn interacts with the carbon nuclei through the scalar re- laxation mechanism. Another remarkable observation in this series of spectra is that the averaged resonance of complexes (XIV), (XV) is significantly broader than the averaged resonance of complexes (XVI), (XVII) at room temperature. The 254 latter two complexes are substituted in terminal posi- tions whereas the former two complexes are bridge-sub- stituted. Again we must refer to Equation 58 in order to determine the relative contributions of the three variables to the observed linewidth differences. Unfortunately, the specification of one of the variables as the major contributor to the linewidth difference is not so straight- forward. We may first consider the asymmetry parameter and the electric field gradient contributions. Certainly, a bridging ligand other than a carbonyl group would perturb the electric field gradient around each cobalt atom by virtue of its size and electronegativity characteristics. Unquestionably, the terminal phosphine substituents would perturb the electric field of the cobalt atom, but the averaged resonances observed result from carbonyl carbons which are under the influence of both a relatively asymmetric field gradient on the one cobalt atom and a relatively symmetric field gradient on the other. It is possible that the narrower averaged resonances result from the fact that only one of the cobalt atoms, rather than both, exhibits a relatively large asymmetric field gradient. However, the effect of the correlation time cannot be ignored. The bridging ligands in complexes (XIV), (XV) most likely do not open to form a non-bridged species, whereas there is nothing to prevent the opening of the bridging carbonyls in complexes (XVI), (XVII) to form non-bridged species. The rotational correlation times 255 in the relatively rigid bridged species would be more pronounced than those in the systems which are capable of Opening and therefore experiencing more degrees of freedom. It is tempting to assign the linewidth dif- ferences between the two types of complexes to the dif- ferences in correlation times, but to adamantly maintain that this is the sole contributing factor would be only a speculation. 4‘. . 1... pl‘.\ 256 Discussion The results of this carbon—l3 nmr investigation of dicobalt carbonyls in solution demonstrate that carbonyl interconversions in these complexes are also facile and general processes. Carbonyl interconversion was con- clusively demonstrated for the substituted complexes (XIV)-(XVII), since discrete resonances, assignable to bridging and terminal carbonyls, were detected at very low temperatures. Although we were unable to reach the slow-exchange limit of complexes (XIII) and (XVIII), it is nonetheless certain that the single resonances ob- served for these complexes at all temperatures result from a carbonyl averaging process, and not to an accidental overlap of bridging and terminal carbonyl resonances, since the bridging resonances in complexes (XIV)-(XVII) always appeared far downfield (at least 30 ppm) from the terminal resonances. Because these complexes are demonstrably fluxional, we must attempt to explain the process by which structural interconversions are effected. Fluxional behavior of the methyl groups has been reported in the cobalt complex [(Ge(CH3)2)Co(CO)3]2, where the Ge(CH3)2 moieties function as bridging ligands.179 This phenomenon is explained non-mechanistically in terms of a "butterfly” motion of the Ge(CH3)2 groups about the cobalt-cobalt axis which interconverts the methyl groups. In a mechanistic sense 257 this operation is viewed as a converse Berry mechanism in which the square pyramidal coordination about each cobalt atom is converted during the flapping process to an intermediate with trigonal bipyramidal coordination about each cobalt atom, with the pseudo—threefold axis being in the plane of the Co-Ge(CH3)2-Co ring. Such a mechanism was employed to rationalize the fluxional be- havior of the iron complex Si(CH3)2FeZ(CO)7, which features two bridging Si(CH3)2 groups and a bridging carbony1,180 Job and Curtis181 have reported that C02(CO)7Ge(CH3)2 exhibits a single line in the methyl region of the proton nmr spectrum. These workers have proposed that the inter- conversion might be accomplished th a bridge-terminal carbonyl interchange in which the new Co-CO-Co bridge is formed on the opposite side of the Co-Ge-Co plane. This process would necessarily average the terminal carbonyl groups. A previously investigated complex which would seem pertinent to our study is [(nS-C5H5)2Fe2(CO)3Ge(CH3)2],72 a complex with a bridging Ge(CH3)2 moiety. However, the carbonyls in this complex were observed to be static on the nmr timescale even at +160°C. The Cotton—Roberts mechanism, and hence the proposal that the molecule interconverts carbonyls ytg a nonbridged intermediate possessing a terminally-coordinated dimethylgermylene moiety, was used to rationalize the fluxional behavior of the -CH3 groups. The interconversion rate of the 258 carbonyl groups was presumed to be so slow that the car- bonyl groups appear to be static. In considering mechanisms by which the carbonyls of these complexes are interconverted, we must examine initially the applicability of the Cotton-Roberts mechanism. 65 and Roberts66 As described more fully in Chapter 2, Cotton independently proposed that complexes similar to [(nS-C5H5)Fe(CO)2]2 (I) undergo carbonyl interconversion through pairwise Opening of the bridging carbonyls to form a non-bridged intermediate, followed by reclosing of the terminal carbonyls to form bridges. The key point in determining the applicability of this mechanism is naturally to ponder the possibility of the non-carbonyl bridges opening in complexes (XIV) and (XV), and to weigh the possibility that the phosphorus ligands in complexes (XVI) and (XVII) can occupy bridging positions. If it is assumed that the bridging ligands in complexes (XIV), (XV) can open to form a nonbridged intermediate, the focus of this must be the geometry of the intermediate, and then the manner in which it might rearrange. An intermediate with trigonal bi- pyramidal geometry, like the open structure in Figure 56, could, in principle, provide a pathway for a "five—two" process (five carbonyls exchanging, two stationary). The apical CO might very well remain unaveraged in an interconversion process which involves rotation about the 259 cobalt-cobalt bond and reformation of the bridging ligands. However, a lineshape analysis based on the re- arrangement of a trigonal bipyramidal intermediate (this technique is described below) yielded spectra that did not match the observed spectra closely. Therefore, rearrangement of carbonyls through a trigonal bipyramidal, nonbridged intermediate is not the process observed. The alternate five-coordinate Stereochemical configuration about the cobalt nuclei in the nonbridged intermediate is the square pyramidal configuration, with the cobalt-cobalt bond forming the apical element. There is no "five-two" process imaginable which could proceed through such an intermediate. The possibility that the bridging ligands in the first two complexes can open to form terminal groups appears to be remote. A three-coordinate carbon atom in a lactone ring is without precedent. There is a precedent for the opening of a germanium bridge to form a nonbridged intermediate in [(nS-C5H5)2Fe2(CO)3Ge(CH3)2], but this process was observed at higher temperature; indeed, the rate of this interconversion process is so slow that even at +160°, discrete bridging and terminal carbonyl reson— ances were observed. Not only are the temperatures at which our observa- tions were made too low to provide sufficient energy for the non-carbonyl bridging groups to open, there are other 182 factors to be considered as well. Poe gt gt. studied 261 hindrance. Therefore, appeal to mechanisms other than the Cotton-Roberts carbonyl interchange mechanism becomes necessary, since the basic tenet of that mechanism, that pairwise opening of bridging ligands initiates the process, cannot be satisfied in these complexes. Solutions of complexes of formula [(nS—C5H5)M(CO)3]2, where M is Cr, Mo, or W, yield proton nmr spectra which indicate that interconversion among trans, cis, and gauche rotamers is a facile process.99 Since these complexes have only terminal carbonyls, the stereochemistry about each metal atom closely matches that of the cobalt atom to which the bridging carbonyl breaks in complexes (XIV), (XV). Although Cotton gt gt. discussed the possibility that carbonyl and cyclopentadienyl rearrangement may pro— ceed through an intermediate with one or both halves of the molecule assuming a trigonal bipyramidal configura- tion, these authors favored a rotational pathway similar to that described for the rotamers in the Cotton-Roberts mechanism.99 However, the idea that rearrangement may proceed through a trigonal bipyramidal intermediate formed the basis of several carbonyl interconversion mechanisms considered in this dissertation. In order to deduce the process(es) by which carbonyl interchange accomplished in the substituted dicobalt carbonyl complexes, it was decided to simulate several series of spectra by modeling several different 262 exchange mechanisms. These results were then compared to the measured carbon-l3 nmr spectra. The spectra obtained from complex (XIV) were treated in this manner initially, because among the several complexes it alone was resolved at low temperatures into the predicted number of resonances, and there appeared to be no viscosity broadening of the resonances at the lower temperatures. After preliminary simulation of approximately ten possible rearrangement mechanisms, the list of possibilities was narrowed to four processes. The basic figure of the molecule utilized in this scheme of analysis is depicted in Figure 73A. The black sphere which serves as a bridging ligand in the figure merely represents the bridging lactone ring or diphenylgermanium ligands which are presumed to maintain their bonding to both cobalt atoms throughout the carbonyl interchange processes. Henceforth this ligand shall be referred to as the "handle." The exchange processes to be described can be characterized by a 4 X 4 exchange matrix based on Figure 73A, or a 7 X 7 exchange matrix which is based on a model with one set of terminal carbonyls labelled E, F, and G. The simulated spectra obtained from a 4 X 4 exchange matrix were practically identical to those obtained through use of the 7 X 7 exchange matrix. The 4 X 4 exchange matrices were used whenever possible, since the cost of running a program using a 4 X 4 matrix is only gg. 30-40% of the cost required to run a program based 263 D B D CACo/——-—.\C{—A C-\Co/.\C o/f-D s/ \C/ \o s/ \3/\ \A Figure 73. Schematic depiction of Mechanism 1. All seven carbonyls are scrambled in a polytopal rearrangement process through a trigonal bipyramidal intermediate. 264 on the corresponding 7 X 7 matrix. However, selected spectra were matched with both sets of matrices to in- sure that identical results were obtained. The exchange matrix used in the calculations was of the following form: IA 'B 'C 'D' Carbonyl A is the bridging carbonyl and carbonyl B, C, and D form the terminal carbonyl set as shown in Figure 73A. Carbonyl interchange mechanisms which can conceivably permute the carbonyl groups and the matrices which describe these mechanisms are now considered. The first carbonyl interchange mechanism which was considered, designated as Mechanism I, can be des- cribed as a polytopal rearrangement which scrambles all seven carbonyl groups, a mechanism most likely operant at the higher temperatures. Consider the structure of the dicobalt carbonyl complexes: in the solid state structure each cobalt atom is approximately square pyramidal (neglecting the cobalt-cobalt bond), where terminal carbonyl D in Figure 73 serves astjmzapical element. As depicted in Figure 73, the initial step of this inter- change mechanism is the Opening of the carbonyl bridge, A, to either side of the molecule; simultaneously, the cobalt-cobalt bond breaks. The breaking of the metal- ‘metal bond is postulated in order to maintain a five- 265 coordinate intermediate, but more importantly, it insures that the effective atomic number (EAN) rule is obeyed by the cobalt atoms; were the metal-metal bond to remain intact, the cobalt atom to which the bridging carbonyl opens would be surrounded by 37 electrons and the other cobalt atom would be complemented with only 35 electrons. Since the EAN rule is maintained "perhaps 99% of the time"185 by metal carbonyl systems, it does indeed appear necessary to postulate the metal-metal bond scission. The cobalt atom toward which the bridging carbonyl opens is seen in Figure 73B to be trigonal bipyramidal, with carbonyls B and C and the handle occupying the equatorial positions and carbonyls A and D occupying the equatorial positions of the trigonal bipyramid. This trigonal bi- pyramidal intermediate can then rearrange in the following manner: carbonyl A can undergo an axial + equatorial interchange with B or C, and carbonyl D can interchange with C or B, processes which yield intermediates 73C or 73E. The axial carbonyl in a position to do so then closes to form a bridging carbonyl, which can then Open to either side and continue with the scrambling process. Concurrent with bridging carbonyl reformation is the re-~ establishment of the cobalt—cobalt bond, which implies that the cobalt atoms are once again square pyramidal. Mechanism I may then be described as an inverse Berry type mechanism, since the scheme square pyramidal + 266 trigonal bipyramidal + square pyramidal describes the process. The exchange matrix utilized to describe this process is .5 -1 0 .5 .5 0 -1 .5 _0 .5 .5 -1 The second process considered is somewhat related to that effected by Mechanism 1. From the spectra of complex (11) at -47°C and of complex (III) at -102°C, it is apparent that only five of the seven carbonyls are interconverting, whereas two others appear to remain static. Mechanism II, which effects the interconversion of four terminal carbonyls with the bridging carbonyl, is pictured in Figure 74. The initial step is the simultaneous opening of the bridging carbonyl to one side and the scission of the cobalt—cobalt bond. In configura- tion 74H, it can be seen that carbonyls A, C, and D form the equator of a trigonal bipyramid, with carbonyl B and the handle forming the apices. Rotation of :120° about the B-Co-handle axis is then accomplished, and in the pro- cess carbonyls A, C, and D are scrambled, whereas carbonyl B retains its identity. Repetition of the process on either side of the molecule completely scrambles four of the terminal carbonyls and the bridging carbonyl. This process would be expected to dominate at lower temperatures, where rotation about the B-Co-handle axis would require 267 -f5Co’//J.\\Cgf:Cl B/'I\§A/’/’ ‘\ G D\ °\ 1’ ;c C— , C ,I’ 3/ AZ’N Figure 74. Schematic depicition of Mechanism II. Two pairs of terminal carbonyls and the bridging carbonyl are scrambled 268 a minimum amount of energy. The exchange matrix which describes this process is -1 0 .5 .5 K = o o o o .5 o -1 .5 .5 0 .5 -1 Another process which scrambles only five of the carbonyls is described by Mechanism III in Figure 75. After bridge opening and cobalt-cobalt bond scission, terminal carbonyl D can reclose to form a bridging car- bonyl, a process which rearranges, or "rocks" the other terminal carbonyls in the following way: in order to accommodate the closing of carbonyl D to form a bridge, the molecule must undergo simultaneous deformation. This deformation, which is more or less a torsional motion, interchanges terminal carbonyls C and D on the cobalt atom to which the original bridging carbonyl A did not open, and it interchanges carbonyls A and C on the other side; during this rearrangement process, however, neither terminal carbonyl B experiences a change of environment. The net effect is the scrambling of the bridging carbonyl with four of the terminal carbonyls, two from each cobalt atom. The exchange matrix for this process is given by O 0000 O O ma2£0£pmo 030 MO 0>Hw moHnEmwom axconumo HOCHEMOO oco mo mchoHo cam ammonumo mewwcwun ecu mo wcflcoao couuoocou .HHH Emflcmnooz mo cofluowaoc OHumEmnom .mn ouswflm 9 < U m,< m m m m / \ / x / < \ clone/\V/oolo II ol 6\ 8 IO III ul oo\ /8 lo m o M y/o\ /o o\ /o\ /o 270 Finally, a mechanism which would probably proceed at higher temperatures is one which describes a totally random carbonyl interchange, and is described using a 4 x 4 matrix of the form -1 .33 .33 .33 .2 -l .4 .4 .2 .4 -l '.4 .2 .4 .4 -1 In order to fit the computer-generated spectra to the measured spectra, it was necessary to make several adjustments in the input parameters of the program. At the lower temperatures, we perceived no problem with line broadening due to the cobalt quadrupole moment, since it is effectively decoupled by the lower temperatures by a process termed "thermal decoupling.” However, at higher temperatures (beginning around -50° for complex (II)) there is noticeable broadening due to interaction with the cobalt quadrupole moment. In order to simulate spectra in this temperature region and above, it became necessary to introduce different T2 values into the input parameters at each temperature. The most successful computer-simulated spectra for the low temperature processes considered, together with the measured spectra, are shown in Figure 76. The processes described by Mechanisms II and III match the Ineasured spectra closely at low temperatures. However, only one mechanism, that which describes the concerted one-for-one carbonyl interchange process, Mechanism III, 1‘.“ ill-Ls Figure 76. 271 The measured carbon-13 nmr spectrum of C02(CO)7(C4H202) (XIV) together with computer- generated spectra based on Mechanism II (left) and Mechanism 111 (right) 272 b)SPECTRA OF Co(CO)7(C4 HZOZXII) 47° C52 26° 0) RANDOM c) CONCERTED EXCHANGE(5-2) -300 EXCHANGEIS-Z) -----:§7:J\k -005_ J‘\ 3144...;5? 00325 A w .0025 A A A A A __J\ -52° .I\J\ _/\.ous A 03:25 A .Jx... -5go - M $03125 it A 5.3- III II II .005 A .075 -69° 40°75 _. . .1 431 -- I1 _j\ 5 .1 -1; III JL-5 III 230 2|O I90 273 is seen to generate spectra in the region -59°C to -37°C that closely match the measured spectra. The spectra generated by the process described by Mechanism II show too much separation between the terminal carbonyl re- sonances in this intermediate temperature range to match the measured spectra. Figure 77 presents an alternate view of Mechanism III. It can be seen that five successive applications of the process that Mechanism III describes completely averages carbonyls A,C,C',D,D', whereas carbonyls B,B' remain unaveraged, thus neatly accounting for the observed Spectra. The calculated value for the activation energy of the carbonyl interconversion process is roughly 15.5 i 2.5 kcal/mole for complex (XIV). The high tempera- ture process may be governed by the polytopal rearrange- ment process, described by Mechanism I, or the totally random exchange process described by Mechanism IV. It is impossible at this time to choose between the two, since the simulated spectra do not discernably differ. It is assumed that the carbonly interconversion process in the germanium-bridged complex (XV) is governed by the same mechanisms. Figure 69 definitely indicates that a 5-2 process is in operation at the lower tempera- tures. Because the slow-exchange limit is below -l64°C, we were unable to accurately simulate spectra for this complex. By assuming a pre-exponential coefficient of 274 HHH Emwcmzom: zn vmuoommo ucoEmwcmupmmp map mo 3mH> mumcumua< .Nm musmwm .o .u a .o a .m < m ol/oo\o/oo\l< u ol/ou\ /oU\Ju u ul Vubouru \ /u\ M \ /o\ / \ /o\ / .m u < o .o .m/ \.u/ \m .U/ \Q/ \< .m/ \