THE PREPARATION AND CHARACTERIZATION OF SOME MIXED-UGAND COMPLEXES 0F TlTANIUM (HI) Thasis for the Degree of Ph. D. MECHEGAN STATE UNIVERSITY GLEN RIC-HARD HOFF 2970 LIBRAP. 1’11 Michigan State - . University This is to certify that the thesis entitled The Preparation and Characterization of Some Mixed-Ligand Complexes of T1tan1um(III) presented by Glen Richard Hoff has been accepted towards fulfillment of the requirements for & degree in 21516—11533! Major professor ifiv DateWJir F1 70 0-169 ABSTRACT THE PREPARATION AND CHARACTERIZATION OF SOME MIXED-LIGAND COMPLEXES OF TITANIUM(III) By Glen Richard Hoff The complexes of titanium(III) chloride with the coordinating solvents acetonitrile, tetrahydrofuran, dioxane, and isopropanol were investigated. The solvent adducts TiCl3-nL were prepared by the direct reaction of the solvent with titanium trichloride or with many of the titanium(III) complexes. Addition of tetraethylammonium chloride to the solvent adducts suspended in the solvent yielded (C2H5)uN[TiClu°2L]. Compounds of the type TiCl3°L and the chloroalkoxo complex Ti2Cl3(C3H7O) were obtained by heating the solvent adducts under vacuum. A new class of compounds of the type TiC13L2L', containing two different non-halide ligands, were prepared by the addition of ligand L' to the adduct TiClB-nL suspended in the solvent L and by the dissolution of TiCl3-L' in the solvent L. The reactions and inter— conversions of all these compounds were studied, especially with regard to non-halide ligand substitution. Glen Richard Hoff The infrared spectra of the complexes generally showed the shifts of absorption bands associated with coordination of the solvents. Far infrared spectra were too broad to provide useful information about the symmetries or bond strengths of the compounds. Two absorption bands were present in the visible spectra which indicated that all the compounds contained some elements of tetragonal or lower symmetry. Jorgensen's rule of average environment was obeyed in the complexes of the type TiCl3L2L'. The frequencies of the absorption bands increased with increasing chloride content of the complexes in contrast to Jorgensen's rule. Magnetic studies over a range of temperatures indicated normal paramagnetism for most of the complexes and several ligand field parameters were obtained. As was expected, the distortion and t2g electron delocalization decreased and the magnetic moment became more dependent upon tempera- ture as the chlorideznon-halide ligand ratio was increased. No correlations between several ligand field parameters were found in complexes of the type T1013L2L'. The compounds TiCl 'CH CN and TiCl 'CHHBO were antiferromagnetic and 3 3 3 TiZCl3(C3H7O)3 was diamagnetic. The electron spin resonance Spectra were measured at 297 and 77°K. The spectra indicated that compounds of the l . , 1 types TiCl3L2L and (02H5)uN[TiClu 2L] as well as T1013 and TiCl3-HC3H8O had cis configurations. In some complexes the separation of the t °2CuH802 2% sublevels was either too small or Glen Richard Hoff too great to permit the observation of more than a single absorption peak even at 77°K. Since most of the complexes had an average g value close to 1.89 it was difficult to confirm the existence of any definite correlations of the g value with ligand field parameters. It appeared that dis- tortion of the complexes and the approach of the g values toward 2.00 increased in the same order as the ligand field strengths of the complexes. Several of the substances prepared changed color as they were cooled to 77°K. These thermochromic materials had charge-transfer bands which narrowed and.shifted to higher frequencies to produce the color change. In addition, when some of the thermochromic substances were cooled to 77°K new peaks appeared in their esr spectra. It is felt that the thermochromism and unusual esr spectra of some of the substances is somehow related to the presence of both titanium(III) and titanium(IV). THE PREPARATION AND CHARACTERIZATION OF SOME MIXED-LIGAND COMPLEXES OF TITANIUM By Glen Richard Hoff A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1970 t/‘I 1‘ ' i" " . '4... ‘3 2 '3 (HI-w.” {~1,.‘..,'_",‘w -,.,.~' 1"“ za‘“ " ‘f’/ To my wife, Janet, and our family ii ACKNOWLEDGMENT The author wishes to extend his thanks and appreciation to Professor Carl R. Brubaker, Jr. for his interest and guidance during this investigation. Financial support from the National Science Foundation was deeply appreciated. iii TABLE OF CONTENTS Page ACKNOWLEDGMENTS. . . . . . . . . . . . . . iii LIST OF TABLES . . . . . . . . . . . . . . V LIST OF FIGURES. . . . . . . . . . . . . . Vi INTRODUCTION. . . . . . . . . . . . . . . 1 EXPERIMENTAL. O O O O O O O O O Q 0 O O O 19 Materials. . . . . . . . . . . . . 19 Analytical Methods. . . . . . . . . . . 22 Experimental Apparatus and Technique. . . . . 23 Preparation of the Compounds . . . . . 2A Spectroscopic Measurements . . . . . . . 31 Magnetic Moment Measurements . . . . . . 32 RESULTS AND DISCUSSION . . . . . . . . . . . 35 Preparation and Reactions of the Complexes. . . 35 Infrared Spectra . . . . . . . . . . . 39 Electronic Absorption Spectra . . . . . . . “8 Magnetic Moments . . . . . . . . . . . 50 Electron Spin Resonance Spectra . . . . . . 68 Thermochromism . . . . . . . . . . . . 72 BIBLIOGRAPHY. O O O O O 0 O O O O O O O O 75 iv Table II. III. IV. VI. LIST OF TABLES Page Low Temperature Baths for Magnetic Suscepti- bilities . . . . . . . . . . . . . . 33 Electronic Absorption Spectra of Compounds. . . A9 Magnetic Properties of Titanium(III) Complexes . 51 Ligand Field Parameters Calculated from Magnetic Data 0 o 0 Q 0 Q 0 0 o o o o 67 The g Values from esr Spectra of the Complexes . 7O Thermochromic Titanium Complexes . . . . . . 72 Figure 20. LIST OF FIGURES Page Relative energy of the titanium(III) d orbitals in octahedral, tetragonal, and trigonal environments. . . . . . . . . . . . . 6 Magnetic behavior of typical (1) paramagnetic and (2) antiferromagnetic substances as a function of temperature . . . . . . . 1A Infrared spectrum of TiCl3‘3CH3CN. . . A0 Infrared spectrum of TiCl3-3CQH80. . . A0 Infrared spectrum of TiCl3s2CuH802 . . . Al Infrared spectrum of TiCl3oAC3H80. . . Al Infrared spectrum of TiCl3(CuH80)2(CHBCN) . A2 Infrared spectrum of TiCl3(CuH802)2(CH3CN). A2 Infrared spectrum of TiCl3(CuH802)2(CuH8O). A3 Infrared spectrum of TiCl3(CAH802)2(C3H80)' A3 Infrared spectrum of TiCl3-CH3CN . . . . AA Infrared spectrum of T1C13°CAH8O . . . . AA Infrared spectrum of TiCl3«CuH802. . A5 Infrared spectrum of TiZCl3(C3H7O)3 . . . A5 Infrared spectrum of (C2H5)uN[TiClu°2CH3CN] A6 Infrared spectrum of (02H5)uN[TiClus2CuH80] A6 Infrared spectrum of (C2H5)4N[T1Clu-2CNH802] A7 Graph of “err vs. £3 for TiCl3~3CH3CN . . 5A Graph of “eff Vs. §$ for TiCl3~BCuH80 . . 55 Graph of “err vs. §$ for Tici3~2CuH802 . . 56 vi Figure Page 21. Graph of “eff vs. €$~for TiCl3‘AC3H8O . . . . 57 22. Graph of ”eff vs. §$-ror TiCl3(CuH80)2(CH3CN). . 58 kT 23. Graph of vs. -for TiCl3(CuH8O2)(CH3CN). . 59 ueff 2A. Graph of ”eff vs. -for TiCl3(CAH8O2)2(CAH8O) , 60 25. Graph of —— for TiCl3(CuH8 (C3H80) . 61 TI kT T" kT “eff VS' A! O2)2 kT 26. Graph of “eff vs. IT for TiCl3.CuH8O2 . . . . 62 kT At kT T'- kT IT 27. Graph of —— for (C2H5)uN[T1Clu.2CH3CN] . 63 ueff VS' 28. Graph of for (C2H5luNETiClus2CuHBOJ . 6A ueff vs. 29. Graph of for (C2H5)4N[T1Clu'2CuH8O2]. 65 ueff VS' 30. Representative electron spin resonance spectra . 69 vii INTRODUCTION Ever since the beginning of coordination chemistry more than 170 years ago, chemists have been concerned with theories about the bonding in transition metal complexes. As the theories are deveIOped and refined they are more able to account for many of the properties of these coordi- nation complexes. Compounds of transition metals which have only one d electron provide the simplest of systems with which to test bonding theories as measurements such as the optical and electron spin resonance spectra and magnetic behavior can be correlated with the behavior of the lone d electron. The symmetries of the complexes generally are easy to predict and are only slightly distorted from ideal point group symmetry. Comparisons between molecular orbital parameters calculated from theoretical considerations and those obtained experi- mentally are especially useful for compounds in which the symmetry is lower than octahedral. The d1 transition metal complexes studied in this laboratory include those of titanium(III),l vanadium(IV),2 niobium(IV),3 molybdenum(V),u and tungsten(V).5 There are several advantages in the use of titanium(III) instead of other metals in a study of d1 transition metal complexes. The properties of titanium(III) 1 compounds are generally sufficiently different from those of the other oxidation states of titanium so that the com- plexes can usually be isolated and characterized. In addition, titanium(III) does not form the oxo compounds which make up a large part of the chemistry of vanadium(IV), molybdenum(V), and tungsten(V). In most complexes with monodentate ligands the titanium atom is octahedrally coordinated. In those cases where the ligand:metal ratio is less than 6:1 one or more of the ligands often bridge two titanium atoms to complete the coordination sphere. ’ Thus characterizations can be based upon distortion or reduction of octahedral point group symmetry. However, titanium(III) compounds are very susceptible to attack by oxygen and water and previous work in their characteriza- tion has been limited. Titanium(III) chloride reacts with a number of coordinating solvents to form complexes of the type TiCl3°nL (n = l-A). Few of the compounds are more than sparingly soluble in the solvent from which they were made and none will dissolve in any other solvent without reaction. The titanium atom appears to be six-coordinate in all of the complexes. In alkyl cyandides titanium trichloride reacts 6-8 to form TiCl ~3RCN (R = methyl, ethyl, n-propyl). 3 Titanium(III) chloride coordinates with three acetone molecules but only two diethylketones or benZOphenones.8’9 The complex HTiClu-CquOO is made from diethylether saturated with HC1.l Titanium trichloride adds three molecules of tetrahydrofuran(THF) but only two of dioxane and one of those is reported to be eliminated with prolonged heating and stirring in dioxane.8’9 TiCl3°l.5L is formed in 1,2—dimethoxyethane or ethylene glycol dimethyl ether.9’10 While titanium(III) chloride reacts with ammonia to form the white hexammine complex,11 it adds only two trimethylamines and undergoes solvolysis with methylamine 12,13 or ethylamine. Under appropriate conditions titanium .2,5dimethylformamide,lu 16 -l.5bipyridine. All three trichloride forms TiCl3 TiC13-2bipyridine,15 and T1013 chlorides are ionic in the amine complexes TiCl -3L 3 (L = ethylenediamine, propylenediamine) and TiCl3- 2diethylenetriamine.7’l7 TiCl3-Aethylenediamine has been 7,18 reported but its existence is questioned. In alCohols titanium forms complexes of the type TiCl -AL (L = 3 methanol, ethanol, is0propanol, sec—butanol, cyclohexanol) 19,20 in which one of the chlorides is ionic. Titanium(III) bromide and iodide also react with coordinating solvents to form compounds analogous to those of the chloride.6’18’20’22 There are several alternative methods to prepare complexes of titanium(III) halides. Addition of the ligand to titanium trichloride suspended in another solvent is used to prepare compounds containing alcohols, alkoxide 1,16,23 ions, nitriles, and amines. In many cases an inter- mediate complex such as TiCl -2(CH3)3N or HTiClu-Et20 is 3 used to prepare compounds not obtainable by direct 6,2A synthesis. Trimethylamine is displaced from TiCl3-2(CH3)3N in dioxane and the very soluble TiCl3'3dioxane is formed.9 Tetraethylammonium halide added to TiX3-3CH3CN (X = Cl, Br) in acetonitrile yields the polymeric (C2H5)uN[TiXu°2CH3CN].25 The intermediate TiBr3-2(CH3)3N, used to prepare a variety of titanium(III) bromide complexes,is prepared by the reduction of 22 titanium(IV) bromide with trimethylamine. TiCl -nCH OH 3 3 (n = A,5) is made by the electrolytic reduction of titanium(IV) chloride in methanol.2u Several complexes eliminate non—halide ligands when heated in a vacuum and compounds of the type TiCl3-2L (L = NH 11:1“,24 25,26 3, THF, DMF), TiCl °L (L = THF, CH3CN), and (C2HS)AN[T1XAJ 3 (X = Cl, Br) are formed.25 Complexes in whichthe titanium atom is coordinated to six identical, non-bridging ligands generally exist only in solution. Systems of the type [TiL6]n containing water, halides, alcohols, urea, acetonitrile, and amines 27-30 have been characterized. Solid compounds which have been isolated include the hexammine complex,5 - 25 (CSHSN)3[T1X6] (X Cl, Br), and Ti(CH30N)6(I3)3 which has been prepared by the oxidation of powdered titanium metal with iodine in acetonitrile.31 Titanium(III), having only one d electron, has a 2D ground state. In an octahedral ligand field this ground state is split into two levels (Figure l). The upper level, 2 Eg, is separated from the ground 1evel,2T2 by an energy A0. g! The orbital degeneracy in each of these two levels is lifted by components of lower symmetry. Generally the ligand field symmetry about the titanium atom is reduced from cubic (Oh) to tetragonal (DAh) or trigonal (D3) although other symmetries are also possible depending upon the nature of the ligands. Trigonal distortion, caused by elongation or com— pression along the C3 axis of the octahedron, leaves the 2BE state of the cubic field unchanged but splits the 2T state into a 2E and a 2A1 state, with an energy 28 separation 5 Tetragonal deformation, caused by com— 1' pression or elongation along the CA axis, splits the 2T283. ground state into an upper 2E8 and a lower 282g level separated by an energy 61. The 2Eg state of the octahedral field is split into a 2A level, and a 2B level, lg lg separated by an energy 62. The ordering of these two levels, shown in Figure l, are those which are most 8,28,32 widely used. They are the reverse of those used by McDonald gt al. which must be considered when spectrochemical data are compared.7 The 2T + 2Eg transition in octahedral titanium(III) 2% complexes falls between 13,000 and 20,000 cm-l. The inten- sity of the peak (e’b3 to 12) is consistent for transitions which are Laporte forbidden but which appear because of 2E 2 S E c > ------- < 10 Dq A0 E 10 Dq 2E ( )2 ‘ ’- - \\\‘ a ,/’< 1 ,’ 1 2g D3 Oh trigonally regular distorted octahedron octahedron Figure 1. Relative energy of the titanium(III) 2B ,I' f I 1g //’ 6 \ 2 \\ l \\_ 2A 1% 10 Dq ,,( )2E ,’ f S \ 6 \. 1 \\ A L 23 2S Duh tetragonally distorted octahedron d orbitals in octahedral, tetragonal, and trigonal environments. 33 vibronic coupling. However, the expected sharp symmetrical ligand-field peak is not seen. The actual peak is broad and asymmetric, showing a distinct doublet structure which indicates that the ligand field must contain at least a tetragonal component. As previously indicated, a tetragonal deformation will split both the ground term (2T2g) and the upper term (2Eg) so that two 2 2 2 2 transitions, 82% + A18 and 82g + Blg’ would be expected. The tetragonal distortion is considered to be 3A,35 the result of the Jahn-Teller effect, which removes the degeneracy of the original cubic field orbitals, and 1 for titanium(III) is on the order of 1,000 to 3,000 cm- complexes. The difference in the energies of the two transitions corresponds to the separation energy 62 of the upper orbitals. Because data concerning the concommitant splitting of the ground level are not obtainable from optical spectro— scopy, the usual practice for a d1 system is to make no correction in either the upper or ground states for lower- symmetry components in the octahedral ligand field.36 The energy A0 of the 2T + 2Eg transition is taken to be that 28 of the more intense upper band of the asymmetric ligand field band. The energy A0 is also equal to 10 Dq, where Dq is the cubic field splitting parameter. To facilitate comparisons of ligand field strengths when the ligands are varied, the practice of ignoring lower ligand-field symmetries also extends to tetragonal and trigonal complexes. Although the perfect octahedral field no longer exists, the energy of the most intense d-d transition is set equal to 10 Dq. In tetragonal complexes this corresponds to the 282% + 281g transition. It can be seen from figure one that trigonal complexes have only one upper state. No amount of trigonal distortion, however great, removes the degeneracy of this Eg state. Thus only a single ligand field peak would be expected to appear in the optical spectrum. However, the spectrum of trigonal complexes shows the same double structure observed in octahedral and tetragonal complexes so a tetragonal com— ponent must also be present. For trigonal complexes the energy of the Alg + Blg transition is taken to be equal to 10 Dq. Visible and ultraviolet spectra have been obtained for many of the titanium(III) complexes reported. All show the asymmetric or double structure caused by tetra— gonal distortion. Several ligands, along with their Dq values, fit into the spectrochemical series in the follow- ing order: CH3CN (2390) > H20 (2010) > (NH2)200 (175A) > (CH3)2CO (1710) > CH3OH (1680) > CAH802 (1590) > CAHBO (1570) > C1 (1370). it is assumed that Jorgensen's rule of average environment 37 is obeyed. The absorption spectrum of a solution of the trisacetonitrile complex of titanium(III) chloride is the same as the reflectance spectrum of the solid. However, the two spectra are clearly different for the tristetrahydrofuran or bisdioxane complexes which would indicate that the nature of the absorbing species in solution is different from that of the solid.8 Charge transfer bands, generally more intense than ligand field bands, appear in the ultraviolet region of the spectrum. They arise from transfer of the d electron of titanium to antibonding n molecular orbitals on the ligands or transfer of n-electrons from the ligands to vacant d orbitals of titanium. The exact nature of the transitions is unknown but tentative assignments have been made in some cases.6’8’9 The infrared spectra of titanium(III) complexes generally Show band shifts and splittings associated with coordination. For example, the 2253 cm..1 band of CH3CN is CN.8 The shifted 35 cm_1 higher in the compound TiCl °3CH 3 3 bands assigned to symmetric and asymmetric C—O-C stretching vibrations of the cycljx:ethers tetrahydrofuran and dioxane are shifted to lower wave numbers and split upon coordina- 8,9 tion through the oxygen. Information about the symmetry of the complex may be obtained from the titanium— halogen bands in the far infrared spectrum. A cis (C3V) lO complex should have two infrared active metal-halogen vibrations while a trans (C2v) complex should have three. On this basis Clark has shown that TiCl °3CuH80 is trans 3 and TiCl '3CH3CN is cis.38 However, the far infrared 3 bands are often too broad to provide any useful informa- tion. Data concerning the variation of the magnetic moment with temperature can provide useful information about axial distortion, spin-orbit coupling, and electron delocaliza- tion in titanium(III) complexes. Extensive reviews of this area are available33’39’uO and only a brief summary will be given here. Paramagnetic susceptibilities, from which the magnetic moments are calculated, vary inversely with temperature according to the Curie law or the Curie-Weiss law _ C xm - T+0 In these equations Xm is the molar susceptibility, corrected for the diamagnetic contribution of titanium and the ligands, T is the absolute temperature, and C is the Curie constant, characteristic for each substance. 0, the Weiss constant which takes intermolecular interactions 11 into account, is the temperature at which a plot of l/xm against T intercepts the T axis. The magnetic moment is calculated from the molar susceptibility by u = 2.8A (me)% for compounds which obey the Curie law.uo’ul The quantity (T+0) must be used in place of T if the Curie-Weiss law is obeyed. However, in many cases the magnetic susceptibility data do not fit either law. The usual practice in these cases is to use the Curie law to determine the magnetic moment at a specific temperature and label this moment the effective magnetic moment “eff at the specified tempera- ture.“2 In the simplest case where there is no orbital angular momentum the magnetic behavior of the complex arises only from electron spin angular momentum. As a result, the spin-only magnetic moment is 1.73 Bohr magnetons for titanium(III). In molecules with a high degree of ligand field symmetry orbital angular momentum is present and will couple with the spin angular momentum. Spin- orbit coupling has a considerable effect upon the variation of the magnetic moment with temperature. Kotani has shown that the magnetic moment of a purely octahedral (11 complex A3 should approach zero as the temperature approaches zero. However, there are no titanium(III) complexes which behave 12 in this manner. Both axial distortion of the ligand field symmetry and orbital delocalization of the t2g electron onto the ligands cause the magnetic moment to tend toward the spin-only value. Figgis has presented a method for the treatment of the magnetic susceptibility data of titanium(III) systems from which estimates can be made of the following factors which quench the orbital angular momentum and cause the magnetic moment to approach the spin—only value: 1. orbital delocalization (l-k) which is the fraction of time the t2g electron spends upon the ligand atoms 2. spin orbit coupling A' which is reduced from the free ion value A by complexationu5 3. energy separation 61 between the orbitals derived from the splitting of the 2T28 state by axial distortion. Figgis has calculated theoretical, effective magnetic moments at several temperatures for a range of values for k (the amount of time the electron spends on the titanium atom) and v, where A positive value for v, and thus for 6 indicates that l, the orbital singlet from the splitting of the 2T28 state l3 lies lowest and is the ground state. A plot is made of the experimental magnetic moments against kT/X where k is the Boltzmann constant, T is the absolute temperature, and A is the free ion spin orbit coupling constant. The resultant curve is compared to the theoretical curves of Figgis and the best values of k and v are determined by the best fit of the curves. The value of A' is determined by substituting A' for A in the abscissa of the plot of the experimental data. The value of A', a fraction of A, is then adjusted until the best match of the curves is made. Although ligand fields of low symmetry cause the magnetic moment to tend toward the spin-only value, the effect is not great even for compounds whiCh have axial ligand field components ten times the spin-orbit coupling constant. The effect of orbital delocalization is much greater. In the preceeding discussion it has been assumed that the individual titanium atoms act independently of each other. Generally this is true but in at least two cases exchange interactions cause the compound to exhibit antiferromagnetism. Figure 2 shows the variation of xm with temperature for paramagnetic and antiferromagnetic substances. Above the Neel point the behavior of both types of compounds is similar, but below the Neel tempera- ture xm of the antiferromagnetic compound rapidly decreases with temperature. Even above the Neel 1A >5 Nee1-———) point T———-—> Figure 2. Magnetic behavior of typical (1) paramagnetic and (2) antiferromagnetic substances as a function of temperature. 15 temperature the magnetic moment of an antiferromagnetic titanium(III) compound is markedly reduced from the spin- only value. Antiferromagnetism may come from intermolecular interactions, extending throughout the Crystal, or intra- molecular interactions, limited to within the molecule. With intermolecular interactions exchange between para- magnetic centers takes place through intervening atoms, known as "super-exchange." Super-exchange is destroyed when the interacting ions are separated by dissolution or dilution in a solid isomorphous diamagnetic host. Intramolecular interactions are unchanged by dissolution or dilution. Earnshaw gt 31.46 have presented a mathemati- cal treatment for the determination of the number of interacting spins S and the value of the exchange integral J for the equation AB = 2JSiS The treatment of the k. magnetic data to find n, the number of spins and conse- quently the number of members of a polymeric chain, and J is similar to the treatment of magnetic data of paramagnetic complexes in the Figgis method. Titanium trichloride A7 exhibits antiferromagnetism but titanium tribromide does A8 not The only complex reported to show antiferromagnetic behavior is TiCl3-CH3CN.25 Electron spin resonance (esr) measurements can provide valuable information concerning the structure, magnetic behavior, and molecular orbital parameters of 16 transition metal complexes. There are many books and reviews which cover the fundamentals and applications of “9 electron spin resonance. Several references which deal specifically with titanium(III) in different crystal field 50 51 symmetries are Kuska, Carrington and Longuet-Higgens, and Gladney and Swalen.52 It would be expected that titanium(III), as a d1 electron system presenting the simplest of crystal field problems, would have been extensively studied. However, a review of the literature has shown that this is not the case. In octahedral lattice sites, such as with cesium titanium alum, the spin—lattice relaxation time is very short because of the presence of a low-lying excited state and resonance is only detected at liquid helium tempera- tures.53-56 Compounds of lower symmetry generally exhibit a resonance at room temperature, although the resonance may be anisotropic and only an average value of the gyromagnetic ratio is seen. Among the titanium(III) complexes of lower symmetry for which a resonance has been observed are those 57,58 59 1,60 containing acetylacetone, halides, or alkoxides. Giggenbach and Brubaker presented the esr spectrum of several mixed complexes containing chloride ion, methanol, 1 methoxide ion, or acetonitrile. The resonance spectrum of trichlortrisacetonitriletitanium(III) has been observed by McDonald gt al.1u and Giggenbach.61 1? T1A8’ which has a natural abundance of 73.9%, has no nuclear spin and consequently does not give rise to hyperfine splitting of the epr spectrum. Tiu7, 7.3% natural abundance, and T1A9, 5.5% natural abundance, have nuclear spins of 5/2 and 3/2 respectively. Hyperfine splittings caused by these two nuclei have been observed by Waters and Makisg and McGarvey.57 The hyperfine patterns are best observed at -A5° and at other temperatures the resonance broadens and only a broad asymmetric signal is seen.1 Thermochromism is the property of a compound to change color when the temperature is changed.u9’62 Generally this occurs in solids when a lowering of the temperature causes the charge transfer bands to sharpen and tail less into the visible region of the spectrum. In this case the color change should be gradual and no other changes in properties should take place. When thermochromism is associated with a phase change in the solid the color change should be abrupt (except for hysteresis) and other changes in the properties of the solid may occur. There are no previous reports of thermochromism being observed in titanium(III) complexes. There are very few reports of titanium(III) complexes of the type TiCl3L2L' where L and L' are different non-halide ligands.. One of the objectives of this research was to pre- pare and characterize compounds of this type which contain acetonitrile, tetrahydrofuran, dioxane, and isopropanol as the non-halide ligands. Another objective was to prepare 18 and study the series of compounds T1013.3L, [TiClu-2L]-, and TiCl3aL. In TiCl3-L chloride ions bridge titanium atoms so that each titanium is coordinated to five chloride ions and one non-halide ligand. The experimental data were examined to find correlations between changes in the properties of the complexes and the nature of the ligands. In some cases the properties of the T101 L' were expected to be inter- 3L2 mediate to those of the parent complexes TiCl onL and TiCl onL'. 3 3 The location of the d-d absorption bands could be predicted from Jorgensen's rule of average environment and the Dq values of the ligands. The symmetries of the TiCl3L2L' complexes were lower than those of the parent complexes. The changes in symmetry were expected to be apparent in the esr spectra and magnetic behavior of the compounds. The effect of an increased halide content was studied in the series TiCl3-3L, (TiClu°2L]-, TiCl3-L. However, TiCI3-L may show additional changes because of the presence of bridging chloride ions. As very few esr spectra of titanium(III) chloride adducts have been reported, the spectra were obtained for all of the compounds in this research to observe the effect of the symmetry and the nature of the ligands upon the g value. EXPERIMENTAL A. Materials Titanium Trichloride.-—Anhydrous titanium trichloride was obtained from K & K Laboratories, Inc. and Research Organic/Inorganic Chemical Company. Titanium(III) and chloride determinations gave an analysis of 99.91% titanium (III) chloride and the compound was used without further purification. It was stored in a Schlenk vessel to facilitate subsequent transfer.63 Solvents.--6uAll solvents were refluxed with an appropriate drying agent under a nitrogen atmosphere except where indicated. Purity of the solvents was checked by examination of their proton magnetic resonance spectrum for evidence of impurities. Acetonitrile was stirred with 200 mesh silica gel several hours and then stirred overnight with calcium hydride. After evolution of hydrogen had ceased, two grams of phos- phorus pentoxide per liter of solvent was added and the acetonitrile was refluxed continuously. It was distilled at a high reflux ratio immediately before use. Chloroform was washed with water and dried with calcium chloride. After the chloroform had refluxed 2A hours with 19 2O phosphorus pentoxide it was distilled and stored in the absence of light before being used. Dioxane was stored over sodium hydroxide. It was refluxed with sodium and distilled after the surface of the sodium had become shiny and the purple color of the disodium benzophenone complex was formed.65 Trimethylamine (anhydrous liquid from Eastman Organic Chemicals) was transferred to a Schlenk vessel and stored at —5°. It was distilled at a reduced pressure into the reaction vessel which was at -78°. Isopropyl alcohol was continuously refluxed with sodium isopropoxide and distilled immediately before use. Tetrahydrofuran, diethyl ether, and benzene were refluxed continuously with lithium aluminum hydride and distilled immediately before use. Tetraethylammonium Chloride.--Tetraethylammonium chloride was obtained from Eastman Organic Chemicals. Twice it was ground, boiled with dry benzene to remove the water, and dried under vacuum. This method gave a product of higher purity than by merely drying the tetraethylammonium chloride in an oven at 100°. Ggggg,--Oil-pumped prepurified nitrogen from Liquid Carbonics was passed through a heated activated copper catalyst (BTS catalyst - BASF R 3-11 from Badische Anilin- & Soda-Fabrik AG) at 150° to remove oxygen and through Aquasorb (indicating moisture absorbant, containing phosphorus pentoxide, 21 from Mallinckrodt Chemical Works) to remove water. The nitrogen in the controlled atmosphere box was continuously recirculated through the Aquasorb and heated BTS columns. The condition of the atmosphere in the controlled atmosphere box was checked by observing the burning time of an exposed light-bulb filament. Hydrogen chloride from Matheson Gas Products was anhydrous and was used without further purification. Analytical Reagents.--Silver nitrate was ground and dried at 180°, reground and stored in a vacuum desiccator. A 16.989 g sample was dissolved in one 1 of distilled water to make a 0.1000 N. solution. A 0.100 N cerium(IV) sulfate solution was prepared from a standard solution from Fisher Scientific and the con- centration checked by titration with ferrous ammonium sulfate. Ferroin (1,10-phenanthroline ferrous sulfate complex) was used as the indicator. Dilute sulfuric acid was prepared by boiling demineral- ized distilled water, adding enough sulfuric acid to make a A M solution, boiling the mixture for another five minutes, and then cooling the solution in a special Schlenk vessel. Nitrogen was rapidly, and continuously bubbled through the solution during its preparation and cooling. B. Analytical Methods Carbon, Hydrogen, and Nitrogen Analysis.--These analyses were performed by the microanalytical laboratory of the Institute of Water Research, Michigan State University, East Lansing, Michigan. Although special handling precautions were taken to avoid decomposition of the samples, they reported the samples very difficult to analyze. Identical samples did not always give similar results. Chloride Analysis.--Generally the solid sample was quickly dissolved in water but in cases where hydrogen chloride may be lost the sample was first cooled to 77°K and dilute sulfuric acid added. After the sample was slowly warmed to room temperature it was treated in the usual manner. It was necessary to oxidize the titanium(III) cataly— tically with air prior to the chloride determination. This was done by the addition of one milliliter of a saturated solution of copper sulfate in 9M sulfuric acid to the dissolved sample and stirring the solution until the color changed from violet to yellow. Chloride was determined by a differential potentiometric titration with a standard silver nitrate solution.66 Silver- silver chloride electrodes and a Corning Model 10 pH meter were used. Titanium(III) Analygis.--Titanium(III) was determined by a cerimetric technique. As the sample was sensitive to oxidation, it was cooled to 77° K in a Schlenk vessel and 23 deoxygenated dilute sulfuric acid was added. After the solution had warmed to room temperature, it was titrated with cerium(IV) sulfate. Ferroin was used as the indicator. A nitrogen atmosphere was maintained over the solution during the dissolution and titration of the sample. The total titanium content was determined spectrophoto- metrically as the titanium(IV) hydrogen peroxide complex in sulfuric acid.67 Beer's law is obeyed at A10 nm for solutions containing up to 75 mg titanium(IV) per liter. The analysis was performed on a portion of the solution prepared for chloride analysis, in which the titanium(III) had been catalytically oxidized. Some samples could not be analyzed in this manner because the solution was turbid and the absorbance reading was thus abnormally high. Although the total titanium analysis also measured any titanium(IV) impurity present in the original sample, the two methods gave very close agreement in most cases. C. Experimental Apparatus and Technique All of the titanium(III) samples prepared were sensi- tive to water and oxygen. All reactions and manipulations were carried out under a nitrogen atmosphere. Schlenk tube techniques, similar to those recently presented by Herzog gt gl.,63 were used throughout. One difference is that inner ground glass joints were generally used on Schlenk vessels to facilitate transfer and so that compounds would not come into contact with the silicone lubricant on the joint. 2A An important factor in the successful preparation of the titanium(III) complexes was the use of a vacuum mani— fold in connection with the nitrogen line. Whenever a glassware connection was made or broken, nitrogen was forced into the vessels through the side—arm stopcocks. This positive flow of nitrogen, about four liters per minute, prevented air from entering the vessel. When the vessel was closed, it was immediately evacuated and the atmosphere replaced with fresh nitrogen. Clean glassware, dried at 180° for several hours, was evacuated and filled with nitrogen several times while it cooled. As the nitrogen of the controlled atmosphere box was not nearly as pure as that in the Schlenk vessels, which could be changed quickly, the box was little used except for the storage of compounds in Schlenk vessels and for the manipulations not amenable to Schlenk tube techniques. These included preparation of mulls for optical spectroscopy and the filling of esr and magnetic susceptibility tubes. D. Preparation of the Compounds Trichlorotrisacetonitriletitanium(III).--Eight grams of titanium trichloride was added to 300 ml of acetonitrile. The mixture was heated and the solvent allowed to reflux several hours. The color of the mixture changed from violet to blue during this time. Half of the solvent was then removed by vacuum evaporation which also served to cool the 25 remaining mixture. The cold solution was filtered and the dark blue crystals were washed with portions of acetonitrile and dried under vacuum. Ti, 17.27; Cl, 38.3A; Analysis. Calc. for TiCl 06H 3 9N3‘ C, 25.98; H, 3.27; N, 15.15. Found: Ti, 17.29; Cl, 38.68; C, 26.3A; H, 3.51; N, 1A.78. Trichlorotristetrahydrofurantitanium(III).—-This com— pound was prepared in a manner similar to that of the above acetonitrile complex. Finely divided turquoise blue crystals were obtained. Analysis. Calc. for T1013012H2u03: Ti, 12.93; Cl, 28.70; C, 38.86; H, 6.53. Found: Ti, 12.98; Cl, 28.70; c, 38.8A; H, 6.5A. Trichlorobisdioxanetitanium(III).--This compound was prepared in a manner similar to that of the above acetonitrile complex. The reaction time was about 16 hours. A pale green powder was collected. Analysis. Calc. for TiCl3C8H16OA: Ti, 1A.50; Cl, 32.19; C, 29.06; H, A.89; Found: Ti, 1A.5A; Cl, 32.22; C, 28.78; H, 5.03. Longer reaction times (60, 100, and 2A0 hours) gave a product similar in appearance, analysis, and behavior. Trichlorotetrakis(2-propanol)t1taniumg I l.-uThis compound was also prepared in a manner similar to that of 26 the above acetonitrile complex. The reaction was complete in two hours and yeilded a blue powder. Analysis. Calc. for TiCl Ti, 12.15; 3012H320A‘ C1, 2Mi96; C, 36.50; H, 8.18. Found: Ti, 12.13; C1, 26.88; C, 36.00; H, 7.95. Trichloroacetonitriletitanium(III).—-A portion of TiCl3-3CH3CN was heated at 100° under a vacuum for several hours until the evolution of a gas ceased. The residue was a brown powder. Analysis. Calc. for TiCl3C2H3N: Ti, 2A.A2; Cl, 5A.A5; C, 12.28; H, 1.55; N, 7.17. Found: Ti, 2A.5A; C1, 5A.l8; C, 11.98; H, 1.65; N, 6.86. The reaction will proceed at 70° but at temperatures above 100° or with prolonged heating a dark violet-brown product was formed. This may contain TiCl3 as titanium and chloride analyses were higher than those above. In some cases a yellow sublimate was collected which appeared to be TiClu-2CH CN. It was thermochromic, turning 3 white as it was cooled to 77°K. Analysis. Calc. for TiCluCuH6N2: Ti, 17.62; C1, 52.17. Found, Ti, 17.72; C1, 53.12. Trichlorotetrahydrofurantitanium(III).--A sample of TiCl3i3CuH80 was heated under vacuum at 75°. A gas was evolved and the sample turned light green and then grey- green. The light green substance, which may be TiCl3-2CuH80, 27 could not be isolated. The gray-green solid turns violet as it is cooled to 77°K. Analysis. Calc. for TiCl3cuH8O: Ti, 21.16, Cl, A6.98; C, 21.20; H, 3.56. Found: Ti, 21.20; 01, A6.80; O, 20.63; H, 3.62. A thermochromic yellow sublimate, apparently TiClA'ZCAH8O’ was also obtained when the samples were briefly exposed to the atmosphere. Analysis. Calc. for TiClAC8H16O: Ti, 1A.35; Cl, A2.A7. Found: Ti, 1A.58; Cl, A2.6A. Trichlorodioxanetitanium(III).—-TiCl 2CAH802 heated 3 to 1A0° under vacuum gave a finely divided cream powder. Analysis. Calc. for TiCl3CAH8O2: Ti, 19.76; Cl, A3.89; C, 19.81; H, 3.33. Found: Ti, 19.98; C1, A3.7A; C, 19.36; H, 3.31. Trichlorotris(2-prgpoxy)dititanium(III).--A sample of TiCl3-AC3H80 was heated under vacuum at 75° for ten hours. Hydrogen chloride and 2-propanol were evolved and the remaining solid turned red-brown. Analysis. Calc. for Ti2Cl3C9H2103: Ti, 25.25; C1, 28.03. Found: Ti, 26.70; 01, 29.33. Trichloroacetonitrilebistetrahydrofurantitanium(III).-- Four grams of TiCl °3CAH8O was suspended in tetrahydrofuran 3 and twenty milliliters of acetonitrile was added. After the solution was stirred and heated nine hours the volume was 28 reduced by vacuum evaporation to forty milliliters. An equal quantity of hexane was added and the solution was stirred vigorously. The blue solid which precipitated was collected on a filter and dried under vacuum. Analysis. Calc. for TiC13ClOH1902N: Ti, 1A.11; C1, 31.33; C, 35.35; H, 5.65; N, A.l3. Found: Ti, lA.25; C1, 31.30; C, 3A.A8; H, 6.02; N, A.01. A different preparative method was tried in which three milliliters of tetrahydrofuran was added to 5.85 g TiCl3-CH3CN suspended in benzene. After the mixture was stirred and heated six hours it was filtered. The gray- green solid, collected by filtration and dried under vacuum, was thermochromic and appeared to be TiCl3-CuH80. Trichloroacetonitrilebisdioxanetitanium(III).--Several grams of TiCl ~3CH CN were suspended in dioxane and the 3 3 mixture was heated and stirred five hours. The very thick blue-green mixture was filtered and the solid was washed with dioxane and dried under vacuum. The blue-green powder is difficult to handle as it becomes charged with static electricity when it is ground. Analysis. Calc. for TiCl OAN: Ti, 12.90; 3ClOH19 01, 28.63; C, 32.31; H, 5.16; N, 3.77. Found: Ti, 12.98; Cl, 28.66; C, 33.Ao; H, 5.15; N, A.78. The preparation worked equally well when TiCl3°CH3CN was used as the starting material or when acetonitrile was added to T1C13~2CuH802 in dioxane. 29 Trichlorotetrahydrofuranbisdioxanetitanium(III).--A quantity of TiCl3-CuH80 was suspended in dioxane and the mixture was heated and stirred. After two hours the solid had dissolved and the solution was brown. The volume of the solution was reduced to forty milliliters under a vacuum and an equal quantity of diethyl ether was added. Immediate precipitation resulted and the solid was washed with dioxane and dried under a vacuum. Analysis. Calc. for TiCl Ti, 11.90; 3012H2A05‘ Cl, 26.A2; C, 35.79; H, 6.01. Found: Ti, 12.00; C1, 26.38; C: 35-13; H, 5.31- The use of TiCl3-3CuH80 as the starting material often did not give a homogeneous precipitate. Trichloro(2-propanol)bisdioxanetitanium(III).--Ten milliliters of isopropyl alcohol was added to six grams of TiCl3o2CuH802 in dioxane and the mixture was stirred and heated overnight. The volume of the solution was reduced under vacuum to A0 m1 and an equal quantity of diethyl ether was added while the mixture was vigorously stirred. The solution was filtered and the pale blue precipitate was dried under vacuum. Analysis. Calc. for TiCl3C11H2A05: Ti, 12.26, Cl, 27.23; C, 33.8A; H, 6.19. Found: Ti, 12.16; Cl, 27.28; C, 33.15; H, 6.08. 3O Tetraethylammonium Tetrachlorobisacetonitriletitanate(III).-- A known quantity of TiCl ~3CH3CN was dissolved in a solution 3 of 10% acetonitrile in chloroform. Upon addition of 2.3 equivalents of tetraethylammonium chloride to the blue solution, the mixture immediately turned yellow. After one-half hour the mixture was brown and after it had been stirred overnight the mixture was green. The solution was filtered and a pale blue-green product was collected and dried under vacuum. Analysis. Calc. for TiCluC Ti, 11.91; 12H26N3‘ C1, 35.27; C, 35.85; H, 6.52; N, 10.A5. Found: Ti, 12.01; 01, 35.15; C, 3A.91; H, 6.1A; N, 10.38. Tetraethylammonium Tetrachlorobistetrahydrofuran- titanate(III).—-A known quantity of TiCl 3.3CAH8O was dissolved in tetrahydrofuran and 2.5 equivalents of tetraethylammonium chloride was added. The mixture was stirred overnight. The solution was filtered and a pale green powder was collected and dried under vacuum. Analysis. Calc. for TiClACl6H 602N: Ti, 10.32; 3 Cl, 30.55; C, A1.A0; H, 7.82; N, 3.02. Found: Ti, 10.36; Cl, 30.52; C, A1.31; H, 8.12; N, 3.25. When the green powder was washed with chloroform it immediately turned orange-brown. After it had been dried under vacuum, the powder was salmon and was thermochromic. This material was also obtained from a solution of 10% tetrahydrofuran in chloroform. The material was non- stoichiometric with a titaniumzchloride ion ratio of 1:A.5. 31 Analysis. Found: Ti, 11.58, Cl, 37.58. Tetraethylammonium Tetrachlorobisdioxanetitanate(III).-- A weighed quantity of TiCl '2CuH8O2 was suspended in dioxane 3 and 2.5 equivalents of tetraethylammonium chloride was added. After the mixture had been stirred overnight it was filtered and the pale green powder which was collected was dried under vacuum. Analysis. Calc. for TiClACl6H36OAN: Ti, 9.67; C1, 28.6A; C, 38.81; H, 7.31; N, 2.83. Found: Ti, 9.61; C1, 28.70; C, 38.59; H, 7.52; N, 2.95. The same material was also prepared from a solution of 10% dioxane in chloroform. E. Spectroscopic Measurements Optical Spectra.——The infrared spectra of the compounds were determined by means of Nujol mulls on sodium chloride plates on a Unicam SP-200 spectrophotometer. Far infrared spectra were recorded on a Perkin-Elmer A57 grating spectro- photometer with the sample enclosed in cesium iodide or polyethylene plates. The visible and ultraviolet spectra were obtained by use of a Unicam SP-800 spectrophotometer and a Cary 1A spectrophotometer. Solid samples were mulled with Nujol and enclosed in quartz plates. As most of the Compounds dissolve only with reaction no solution spectra Of the complexes were obtained. A Beckman DU spectrophoto- meter was used for the spectrophotometric analysis of titanium. 32 Electron Spin Resonance Spectra.-—The X—band esr spectra of the solid complexes were recorded at 297°K and 77°K. Compounds which were thermochromic were also studied at a range of intermediate temperatures. The X-band spectro- photometers used were a Varian E-A esr spectrometer system and the Varian A501-0A spectrometer previously described.“9 First derivative curves were recorded on an X—Y recorder withthe X axis proportional to the magnetic field strength. A sample of pitch in potassium chloride (g = 2.0028) was used as a calibration standard. An E-257 variable tempera- ture unit was used in the variable temperature studies. At lower temperatures the microwave power had to be greatly reduced to avoid saturation of the samples. The g values were calculated from the klystron frequency and the measured magnetic field. F. Magnetic Moment Measurements Magnetic susceptibilities were measured by the Guoy method by use of equipment similar to that previously described.69 A specially designed Dewar vessel was used for temperatures lower than room temperature. A constant flow of helium was used to provide heat transfer from the sample and to prevent water from condensing on the sample tube. 33 TABLE I.——Low Temperature Baths for Magnetic Susceptibilities. Temperature Bath 77°K Liquid Nitrogen 1A8°K Petroleum Ether Slush 175.5°K Methanol Slush 195°K Dry Ice—isopropanol 209.7°K Chloroform Slush 250.3°K Carbon Tetrachloride Slush The magnetic susceptibility was calculated from the experimental data according to the equation lo6x z d+B&F-5) s where x is the gram susceptibility of the sample, a is the correction for the displacement of air by the sample, B is the tube constant, F is the measured force on the sample and the tube, 6 is the force on the empty tube, and WS is the weight of the sample. The correction for displaced air, a, was eliminated by weighing the empty tube filled with nitrogen. The constants B and 5 must be measured for every field strength and temperature used. Although borosilicate glass is diamagnetic, it was found that impurities caused 0 to be positive and on the order of 10-15 milligrams. To evaluate B the tube was standardized with Hg[Co(SCN)u] Which has a susceptibility of 16.AAx106 cgs units of 20°. This standard obeys the Curie-Weiss law with xmm(T+10)_l.7O 3A The magnetic susceptibilities of all compounds were measured at 77, 195, and 297°K. If the sample exhibited normal paramagnetism the magnetic susceptibilities were also measured at several intermediate temperatures so that an accurate plot of the effective moment against gg-could be obtained. As is usual for titanium(III) solvent adducts, no corrections were made for the small temperature inde- pendent paramagnetism of the samples. The molar susceptibility was obtained by multiplying the gram-susceptibility by the molecular weight. The metal ion susceptibility was then found by applying corrections for the diamagnetism of the ligands as computed from Pascal's constants.71 RESULTS AND DISCUSSION A. Preparation and Reactions of the Complexes During the course of this research the complexes of titanium(III) chloride with several coordinating solvents were studied. A new class of compounds, containing two different non-halide ligands, as well as several new compounds of classes previously reported were prepared. In addition, several known compounds were prepared to complete the characterizations. The reactions of all these compounds were studied, especially with regard to non-halide ligand substitution. Titanium trichloride reacted with the coordinating solvents acetonitrile, tetrahydrofuran, and isopropanol to form the adducts TiCl °3CH CN, TiCl -3CuH80, and TiCl ~Ac H80 3 3 3 3 3 reSpectively. When the adduct of one solvent was dissolved in another solvent, all of the non-halide ligands were replaced by the latter solvent. However, TiC13(CuH80)2(CH30N) was formed when TiCl °3CuH80 was dissolved in a solution of 20% 3 acetonitrile in tetrahydrofuran. The solid adduct produced by the reaction of titanium(III) chloride with dioxane was TiCl3°2CuH802 in which only two dioxanes were coordinated with the titanium. Clark gt gt.8 had concluded from the uv spectra that a different species 35 36 exists in solution though the species was not identified. Fowles gt a1.9 reported that a dioxane solution of the very soluble complex TiCl '3CAH8O2 gave a similar spectrum. 3 Fowles gt 21- also reported that with prolonged reaction times (up to sixty hours) TiCl 'CAH802 was formed. However, 3 in this investigation reaction times as long as 2A0 hrs. produced only TiCl ‘ZCAH8O2° As each titanium was coordinated 3 to only two dioxane molecules, compounds of the type TiCl3(CuH802)2L were prepared by the addition of the solvent ligand L to TiCl '2CAH802 suspended in dioxane. This type 3 of complex was also produced when a titanium(III) solvent adduct was dissolved in dioxane. When the solvent adducts TiCl °nL (L = acetonitrile, 3 tetrahydrofuran, dioxane) were heated under vacuum, compounds -L were formed. Fowles and Russ25 had 3 reported that a green product was obtained when TiC13°3CH3CN of the type TiCl was heated one-half hour at 100°. In this research it was found that longer heating times were necessary and that the substance which remained after the evolution of a gas had ceased was brown. However, elemental analysis of this -CH CN. 3 3 '3CAH8O was heated under vacuum at 75° it substance showed it was TiCl When TiCl3 first turned light green and then gray-green. The light green substance could not be isolated as the color of the material was never uniformly light green. Analysis of the gray-green compound showed it was TiCl 3 and Thurn2u reported that a green compound, TiCl 'CuHBO. Seifert 3‘2CuH80, 37 was produced when TiCl '3CAH80 was heated under vacuum at 3 70°. If the temperature was then increased to 90° another mole of tetrahydrofuran was lost and gray TiCl 'CAH8O remained. 3 During the course of the present research it was also found that if the samples had been previously exposed to air a yellow sublimate was also formed. The sublimates were primarily TiClu°2L'butcontained some titanium(III). The TiClu-2L complexes and TiCl3-CuH80 were thermochromic. When compounds of the type TiC13(CuH802)2L were heated under vacuum at 70° they eliminated the ligand L to form TiCl3-20uH802. When the temperature was raised to 100° a mole of dioxane was then lost to produce the final product TiCl3-CuH8O2. TiCl3(CuH8O)2(CH3CN) yielded TiCl when heated under vacuum to 80°. 3'CuH80 Generally the compounds of the type TiCl3-L reacted with coordinating solvents in the same manner as titanium trichloride. However, in dioxane the ligand L was not eliminated when the titanium coordinated with two dioxane molecules. This is a convenient method of preparation of 2)2L (L = acetonitrile, tetrahydrofuran). Attempts were made to produce other the complexes TiC13(CuH80 compounds by similar methods. After the addition of acetonitrile to TiC13~CuH80 or TiC13-CuH8O2 in stoichlometrlc amounts to form TiCl3(CH3CN)2(CuH80) and TiC13(CH3CN)2(CuH802) respectively, the starting materials were recovered unchanged. 38' When a stoichiometric amount of tetrahydrofuran was added to TiCl3-L (L = acetonitrile, dioxane) to produce compounds of the type TiCl3(CuH80)2L, the only product was TiCl3-CuH80. Both hydrogen chloride and isopropyl alcohol were liberated when TiCl3-A03H80 was heated under vacuum at 75°. The red material which remained was (probably) a polymeric E1 chloroalkoxo complex with the simplest formula T12C13(C3H70)3. This material did not appear to react with coordinating solvents which indicated that the alkoxide ligands were not A easily replaced. r The reaction of tetraethylammonium chloride with titanium(III) adducts TiCl onL in tetrahydrofuran or dioxane 3 produced compounds of the type (02H5)uN[TiClu-2L]. However, in acetonitrile the reaction yielded only non-stoichiometric materials with a variable chloride content. In a chloroform- acetonitrileor chloroform-dioxane solvent system the compounds (C2H5)N[TiClu-2L] (L = acetonitrile, dioxane) were produced. When the reaction took place in a chloroform-tetrahydrofuran system, or when (C2H5)N[TiC1u~2CuH80] was washed with chloroform, the tetrahydrofuran was eliminated from the complex and the composition of the solid product approached that of (02H5)uN[TiC1u]. If complexes of the type LiC13L2L' were used in place of TiCl3-nL in the reactions, the ligand L' was eliminated and (02H5)uN[T101u-2L] was formed. 39 B. Infrared Spectra The infrared spectra of the complexes were determined in Nujol mulls and the results are presented in figures 3 through 17. In general the spectra showed the shifts in absorption bands associated with coordination of the solvents. The CN stretching vibration of acetonitrile (2253.5 cm-l) was shifted upward 35 cm.1 upon coordination through the ] nitrogen. The combination band at 2288 cm-1 was shifted upward 25 cm-l. In tetrahydrofuran coordination through the oxygen was indicated by the lowering and splitting of E the symmetric (909 cm-1) and asymmetric (1071 cm-1) C—O-C stretching vibrations of the free ligand by ’60 cm-l. Coordination through oxygen in dioxane lowered and split the free ligand bands occuring at 1125 and 883 cm-1. The spectrum of TiCl '2CAH8O2 showed the presence of both free 3 and coordinated oxygen while the spectrum of TiCl3'CuH8O2 showed only coordinated oxygen. It was apparent from the spectrum of Ti2CI3(C3H7O)3 that the compound was a chloroalkoxo complex and no isopropyl alcohol remained. The spectra of complexes of the type TiC13L2L' showed some shifts of absorbtion bands from their positions in the parent complexes TiCl3onL and TiCl3-nL'. Considerable changes in the spectra were observed when the chloride:non- halide ligand ratio of the complexes was increased. These Changes may have arisen from changes in both the symmetries A0 H. f L l A 4000 2000 1600 1200 800 CH! Figure 3. Infrared spectrum of TiCl3s3CH3CN. ’1 r1 A W l 7 l l h.-- l A I l A _L 4000 2000 1600 1200 800 cm"1 Figure A. Infrared spectrum of TiCl3s3CuH80. Al /\ « L l 4000 2000 1600 cm—1 1200 800 Figure 5. Infrared spectrum of TiCl3-204H802. . A U L 1-1.“. 4000 2000 1600 1200 800 cm"1 Figure 6. Infrared spectrum of TiCl3~AC3H8O. A2 wfl ,. j A L 1 1 L UL 4000 2000 1600 1200 800 cm-1 Figure 7. Infrared spectrum of TiCl3(CuH80)2(CH3CN). “N 0 ‘ L A A l 1 A 4 A L 4000 2000 1600 1200 800 cm‘ 1 Figure 8. Infrared spectrum of TiCl3(CuH802)2(CH3CN)- A3 A A 7 A 1 l j j l l 1 4000 2000 1600 1200 800 cm’1 Figure 9. Infrared spectrum of TiCl3(CAH802)2(CAH80)' fl M (HM .01.”..- 4000 2000 1600 1200 800 cm"1 Figure 10. Infrared spectrum of TiC13(CuH802)2(C3H8O). AA /\ I l l L 4000 2000 1600 cm 1200 800 Figure 11. Infrared spectrum of T1013~CH30N. ”n /\l - A A 4 A L _L A U. 4000 2000 1600 1200 800 cm"1 Figure 12. Infrared spectrum of TiC13~CuH80. A5 A L n L a l l L 4000 2000 1600 1200 800 cm Figure 13. Infrared spectrunlof TiCl3~CuH802. ”\W A A A A n A 1 1 L 4 L l 4 4000 2000 1600 1200 800 cm"1 Figure 1A. Infrared spectrum of T12013(C3H70)3. A6 J 1 l L lie L n L 4000 2000 1600 1200 800 cm Figure 15. Infrared spectrum of (C2H5)4N[T101u-2CH CN] 3 D N b .1 A A A I A A _L A 4000 2000 1600 1200 800 cm'1 Figure 16. Infrared spectrum of (C2H5)uN[TiClu~2CuH8O] A7 /\ n l A 1200 800 P l l I l 4000 2000 1600 cm—1 Figure 17. Infrared spectrum of (C2H5)4N[T101u-2CAH802] A8 and the ligand field strengths of the complexes. The far infrared spectra absorption bands were too broad to provide useful information about the symmetries or the ligand bond strengths. C. Electronic Absorption Spectra The electronic absorption spectra of the complexes were . determined in Nujol mulls and the results are presented in .1 Table 11. Titanium was six-coordinate in all of the complexes which were studied. Although the symmetries of the complexes varied widely, all the spectra except that of (02H5)uN[TiCln] j had two d-d absorption bands. Since no amount of trigonal distortion of the basic octahedral point group symmetry would have removed the degeneracy of the upper 2E d orbital level, the complexes contained some elements of tetragonal or lower symmetry. The spectrochemical series for the ligands studied in this research is CH3CN>C3H80> CAH802>CAH8O' The locations of the absorption bands of the complexes of the type TiCl L L' were about 100 cm"1 3 2 from Jorgensen's rule of average environment and the Dq values higher than the values predicted of the ligands. In the [TiClu-2LJ- complexes the locations of the absorption bands were almost identical, although ' ~ Jorgensen 3 rule would have predicted VCH3CN>vCAH8O2~VCAH8O. A9 TABLE II.-—Electronic Absorption Spectra of Compounds.a’b -1 c Compound 2 Absogption Bands (gm ) 2 B2 " B1 32” Al TiCl3°3CH30N 17,100 1A,700 T1013'3CMH80 1A,700 13,500 TiCl3-2CuH802 1A,800 12,800 TiCl3-AC3H80 16,700 1A,300 TiC13(CuH80)2(CN3CN) 15,600(15,550) 14,000(13,900) TiCl3(CuH802)2(CHZCN) 15,680(15,570) 13,325(13,A33) TiC13(CuH802)2(CuH80) 1A,780(1A,770) 13,100(13,03A) TiCl3(CuH8)2)2(C3H80) 15,A00(15,1A0) l3,220(l3,110) TiCl3-CH3CN 16,000(1A,832) 11,900(12,900) TiC13-CuH80 17,300(1A,o32) 13,000(12,500) TiCl3-CuH802 17,900(1A,065) 1A,600(12,270) T12C13(C3H7O)3 18,950 17,A50 (C2H5)uN[TiC1u-2CH3CN] 15,150(15,800) 1A,150(13,800) (C2H5)uN[T101u-2CAH8O] 1A,900(1A,200) 1A,150(l3,000) (C2H5)uN[TiCIu°2CuH802] 15,000(1u,270) 1A,100(12,730) [(C2H5)uN][TiCln] 12,Aoo aall spectra were determined in Nujol mulls. bvalues in parentheses are those expected from Jorgensen's rule of average environment and the average Dq values of the ligands. cassignments assume tetragonal or lower symmetry. "fl 50 Only one low-lying absorption band was seen for the material with non—integral stoichiometry (C2H5)uN[Ti01n]. The absorption bands in compounds of the type TiCl3tL were considerably higher than would have been expected from Jorgensen's rule. In addition,the position of acetonitrile in the spectrochemical series was shifted so that it became the weakest of the ligands studied. It was expected from Jorgensen's rule that the location of the absorption bands would have shifted to lower frequencies as the ch1oride;non-halide ligand ratio was increased. However, in this investigation it was found that the frequencies of the absorption bands increased with increas- ing chloride content of the titanium(III) complexes. D. Magnetic Moments The magnetic susceptibilities and effective magnetic moments of the complexes at several temperatures are presented in Table 111. Most of the compounds exhibited normal Curie- Weiss paramagnetism. The complexes TiCl -CH CN and 3 3 TiC13-CuH80, in which chloride ions probably bridge two titanium atoms, were anti-ferromagnetic. A similar compound, TiCl3-CuH802, exhibited normal paramagnetism but with a much greater temperature dependence than the other paramagnetic complexes studied. The chloroalkoxo complex Ti2Cl3(C3H7O)3 was diamagnetic. Several ligand field parameters were obtained from graphs of ”eff against.kT/A' (figures 18-29). The experimental 51 TABLE III.--Magnetic Properties of Titanium(III) Complexes. ' 6 Compound Tgmp. Xm x 10 “eff K cgs units B.M. TiC13-3CH3CN 297 12A1 1.72 250.3 1A20 1.69 209.7 16A5 1.67 175.5 1902 1.6A 1A8 2222 1.63 77 A019 1.58 TiCl3°3CuH80 297 12A6 1.72 250.3 1A38 1.70 209.7 16A6 1.67 196 1738 1.66 175.5 1931 1.65 1A8 2220 1.63 77 A021 1.58 TiCl -20 H O 297 1232 1.72 3 u 8 2 250.3 1391 1.68 209.7 161A 1.65 196 1705 1.6A 175.5 1873 1.63 1A8 2122 1.59 77 3670 1.51 TiC13-AC3H8O 297 1270 1.7A 250.3 1A16 1.69 209.7 1616 1.65 196 1692 1.6A 175.5 1859 1.62 1A8 2120 1.59 77 3738 1.52 TiCl (C H 0) (CH CN) 297 1381 1.77 3 u 8 2 3 250.3 1A91 1.73 209.7 1708 1.70 196 1796 1.69 175.5 1988 1.68 1A8 2311 1.66 77 A122 1.60 TiCl (C H 0 ) (CH CN) 297 1230 1.72 3 u 8 2 2 3 250.3 1370 1.66 209.7 1575 1.63 196 1621 1.60 175.5 17A5 1.57 1A8 1957 1.53 77 3302 1.A3 TABLE III.--Continued. 52 Compound Temp. Xm x “eff K cgs units B.M. TiCl (C H 0 ) (C H 0) 297 12A3 1.72 3 u 8 2 2 8 8 250.3 1A01 1.68 209.7 1587 1.6A 196 1686 1.63 175.5 1815 1.60 1A8 2060 1.57 77 3533 1.A8 TiCl (C H 0 ) (C H 0) 297 1255 1.73 3 8 8 2 2 2 8 250.3 1A28 1.70 209.7 1670 1.68 196 17A8 1.66 175.5 1926 1.65 1A8 2231 1.63 77 3928 1.56 T1013.CH3CN 297 789 1.37 196 971 1.2A 77 1928 1.09 196 356 0.75 77 305 O.A3 TiCl -C H O 297 1357 1.80 3 u 8 2 250.3 1A77 1.73 209.7 1625 1.66 196 1624 1.60 175.5 1776 1.59 1A8 1990 1.5A 77 2269 1.19 Ti2CI3(C3H7O)3 diamagnetic (C H ) N[TiCl °2CH CN] 297 1305 1.77 2 5 u u 3 250.3 1A78 1.73 209.7 1661 1.68 196 1739 1.66 175.5 1918 1.65 1A8 2192 1.62 77 3761 1.53 Table III.--Continued. 53 Compound Tgmp. x 10 “eff K cgs units B.M. (C H ) N(TiCl °2C H O] 297 1318 1.78 2 5 u u u 8 250. 1A61 1.72 209. 166A 1.68 196 1758 1.67 175. 1870 1.63 1A8 2165 1.61 77 3668 1.51 (C H ) NETiCl °2C H O J 297 1265 1.7A 2 5 u u u 8 2 250. 139A 1.68 209. 1600 1.6A 196 1638 1.61 175. 1765 1.58 1A8 2008 1.55 77 32A2 1.A2 “eff 5A Figure 18. *4: Graph of “eff vs. §$-ror T1013~3CH3CN. Circles are experimental points. Line calculated from reference AA with k = 0.7, v = A.5, A' = 0.95 A. “eff 55 0.2 0.A 0.6 0.8 1.0 1.2 1.A kT IT Figure 19. Graph of Nerf vs. %$ for TiCl3s3CuH80. Circles are experimental points. Line calculated from reference AA with k = 0.7, v = A.5, A' = 0.95 A. ueff 56 Figure 20. vs. §$-for TiC13~2CuH802. Graph of “eff Circles are eXperimental points. Line calculated from reference AA with k = 0.7, v = 3, A' = 0.95 A. 57 g A! Circles are experimental points. Line calculated from reference AA with k = 0.75, v = 3, A' = 0.95 A. Figure 21. Graph of u vs. for TiCl ~AC H80. eff 3 3 1.8 1.7 ueff 1.6 58 Figure 22. 0.4 0.6 0.8 1.0 1.2 1.A kT A! kT Graph of “eff vs. T7 for TiCl3(CuH80)2(CN3CN). Circles are experimental points. Line calculated from reference AA with k=0.9,v=6,)\'=).. 1.7 1,6 ueff 1.5 1.A 59 ’32. A: Circles are experimental points. Line calculated from reference AA with k=0.7, v=2, 1' =0.95 A. Figure 23. Graph of “eff vs. for TiCl3(CuH802)(CH30N). 60 1.7 1.6 “eff 1.5 l.“ E. it Figure 2”. Graph of ”eff vs. §$ for TiCl3(CuH802)2(CuH8O). l Circles are experimental points. Line calculated from reference nu with k = 0.7, v = 2.5, A' = 0.95 A. 61 ueff Figure 25. Graph of “eff vs. g; for TiCl3(CuH802)2(C3H80). Circles are experimental points. Line calculated from reference HQ with k = 0.7, v = u, A' = 0.90 A. 62 1.8 . CD L- Ci 1.6 r . 0 ID 1:... CH 391.“ r 1.2 _ . 1.0 l L *2— l 0.2 0.1: 0.6 0.8 1.0 1.2 1.14 kT 3‘... Figure 26. Graph of “eff vs. §$ for TiCl3~CuH8O2. Circles are experimental points. Line calculated from reference uu with k = 0.8, v = 0.1, A' = 0.90 A. 63 1.7 eff 3'1.6 1.5 0.2 0.“ 0.6 0.8 1.0 1.2 l.“ . kT O Flgure 27. Graph of “eff vs. IT for (C2H5)uN[TiClu.2CH3CN]. Circles are eXperimental points. Line calculated from reference 44 with k = 0.8, v = 3.5, A' = 0.90 A. 6H C 1.7 - ‘. l O O O O 21.6 - (l) 1.5 , . 0.2 0.4 0.6 0.8 1.0 1.2 1.4 fl Ah Figure 28. Graph of "eff vs. §$ for (C2H5)HN[TiClu-2CMH80]. Circles are experimental points. Line calculated from reference nu with k = 0.8, v = 3.0, M = 0.90 A. “eff 65 1.7 . 1.6 . 1.5 ' 1,H . Figure 29. 0.“ 0.6 0.8 1.0 1.2 l.“ kT A“ kT . Graph of “eff vs. YT for (C2H5)uN[TiClu‘2CuH8O2]. Circles are eXperimental points. Line calculated from reference uu with k = 0.75, v = 2.0, A‘ = 0.95 A. 66 curves were fitted to the proper theoretical curves calculated from the values given by Figgis.uu Values of k, the orbital delocalization factor, A', the spin-orbit coupling constant, and v, the distortion factor, were obtained by this procedure. The energy of separation, 6 of the orbital levels derived 1’ from the 2T2g ground state term was calculated from the equa- tion v=6l/A. The value of 61 was positive in all of the complexes which indicated that the orbital singlet derived from the 2T state was the lowest energy level. The ligand 2% field parameters which were calculated from the magnetic data are given in Table IV. The results were examined to find if any correlations existed between the values of the ligand field parameters 3 and 61 and other ligand field parameters of the complexes. Theoretically a complex which has an undistorted octahedral symmetry should have a magnetic moment which falls to zero as the temperature approaches zero. In complexes having lower symmetry and t2g electron delocalization the moment approaches the spin—only value. Thus it would be expected that as the symmetries of the complexes were increased the magnetic moments should be more temperature dependent. This was the case in the series of compounds with different chloride:non- halide ligand ratio in which both the amount of orbital delocalization (l-k) and the energy of separation 61 decreased with increased chloride content. However, with the complexes of the type TiCl L L' no correlations were 3 2 found between the values of the ligand field parameters k 67 TABLE IVr—Ligand Field Parameters Calculated from Magnetic Data. Compound k v A'(Cm-l) 51(Cm-l) TiCl3°3CH3CN 0.7 u.5 1M6 660 T1013-3cuH80 0.7 u.5 146 660 TiCl3°2CuH802 0.7 3.0 1u6 uuo T1013-uc3H80 0.75 3.0 1M6 uuo TiCl3(CuH80)2(CH3CN) 0.9 6.0 15M 925 TiC13(CuH802)2(CH3CN) 0.7 2.0 1M6 290 TiCl3(CuH802)2(CuH8O) 0.7 2.5 1M6 365 TiCl3(CuH802)2(C3H80) 0.7 u.0 139 550 TiCl3-CuH802 0.8 0.1 139 in (C2H5)uN[TiClu-2CH3CN] 0.8 3.5 139 u85 (C2HS)HN[T101u-2cuH80] 0.8 3.0 139 M15 (C2H5)uN[TiClu-2CMH802] 0.75 2.0 1H6 290 68 and 6 and either the ligand field strength (lODq) or 6 l 2, the energy of separation of the upper levels derived from the 2E state. 2s E. Electron Spin Resonance Spectra The esr spectra of the powdered solid complexes were determined at 297 and 77°K. The results are presented in Table V. Figure 30 shows the three types of spectra generally found. Spectra of the type A occur when g, the Landé spectro- scopic splitting factor, is isotropic or, more commonly, the temperature is too high to allow resolution of anisotropic g values. Compounds with type B spectra generally have an axis of symmetry and gx=gy¢gz. In the titanium(III) com- pounds investigated during the course of this research gi(gx and gy) was greater than g||(gz). Type C spectra are expected when the complex has no axial symmetry and gx#gy#gz. -3CH CN and TiCl The esr spectra of TiCl ~3CuH8O are 38 3 symmetries. 3 3v 2v the titanium atom is six—coordinate in TiCl 3 expected for the respective C and C Although 3-2CuH802 it is not certain which functional group bridges the titanium atoms. If the complex had a dioxane molecule in the sixth position or the two dioxane molecules present were Erggg, a type B spectrum would have been observed. Instead a three-line type C spectrum was observed which suggested that a bridging chloride ion occupied the sixth position and the complex had a gig configuration. TiCl3-UC3H80, which had only two chloride ions coordinated to the titanium atom, also had a three-line type C spectrum and a cis configuration. 69 TYPE A .-— 4 TYPE B TYPE C Figure 30. Representative electron spin resonance spectra. 70 TABLE V.-—The g Values from esr Spectra of the Complexes.a Compound Temp. Type of OK spectrum g1 82 83 TiCl3~3CH3CN 297 A 1.907 77 B 1.921 1.883 TiCl3s3CuH80 297 A 1.88u 77 B 1.89u 1.8u9 TiCl3-2CuH802 297 B 1.972 1.880 77 C 1.982 1.880 1.826 T1013.403H80 297 A 1.901 77 C 1.977 1.899 1.828 T101 (0 H 0) (CH CN) 297 A 1.909 3 u 8 2 3 77 A 1.909 TiCl3(CuH8O2)2(CH3CN) 297 C 1.963 1.890 1.836 77 C 1.987 1.903 1.837 TiCl (c H 0 ) (C H 0) 297 A 1.885 3 u 8 2 2 u 8 77 c 1.939 1.898 1.827 T101 (0 H 0 ) (c H 0) 297 A 1.887 3 u 8 2 2 3 8 77 B 1.911 1.855 T1013~CH30N 297 A 1.911 77 A 1.907 TiCl3.CuH80 297 A 1.898 77 A 1.898 TiCl .c H 0 297 A 1.879 3 u 8 2 77 A 1.881 (C2H5)uN[TiClu~2CH3CN] 297 c 1.9u1 1.869 1.807 77 c 1.9u0 1.869 1.806 (C2H5)uN[TiClu‘2CuH80] 297 c 1.977 1.885 1.796 77 c 1.978 1.886 1.796 (c H ) N[TiCl .20 H 0 ] 297 0 1.983 1.886 1.815 2 5 u u u 8 2 77 c 1.983 1.885 1.813 a. for S GCtI’a Of ’6 e B P - and = o 71 Compounds of the type TiCl3L2L' have no axis of symmetry and were expected to give a three-line type C spectrum. However, this type of spectrum was observed only for T1C13(CuH802)2(CH3CN) and TiCl3(CUH802)2(CHH80)' The unsymmetrical spectrum of TiCl3(CuH80 (C3H80) may have 2)2 been an unresolved three—line spectrum. The spectrum of TiCl3(CuH80)2(CH3CN) remained a single absorption peak even at 77°K. This was caused by the large splitting of the ground state 2T2g (61=990 cm-1 from the magnetic data) which quenched the orbital angular momentum and caused g to become approximately isotropic. The spectra of compounds of the type TiCl -L were expected to show the axial symmetry of these 3 complexes. However, the spectra had only a single absorption peak. The spectrum of TiCl ~CuH802 was very broad due to 3 the small separation 61 of the t2g sublevels. As TiCl3-CHBCN and TiCl3-CuH8O were antiferromagnetic, the separation energy could not be evaluated from the magnetic data. Powdered antiferromagnetic substances generally give broad, inhomogeneous resonance spectra.72 Compounds of the type (C2H5)uN[TiC1u-2L] gave three-line type C spectra even at room temperature. The spectra indicated that the compounds have gig configurations. As the symmetries of the complexes became more distorted from octahedral symmetry, the orbital angular momenta were quenched and the average g values were expected to approach 2.00. Since most of the complexes had average g values close to 1.89 it was difficult to establish the existence of any 72 definite correlations. However, the distortion of the complexes and the approach of the average g values to 2.00 appeared to be in the same order as the ligand field strengths of the complexes with the ligands in the order CH CN>C 3 3H80> CMH802~CHH80>C1' F. Thermochromism Several of the complexes prepared during the course of this investigation were thermochromic. The compounds and the color changes that occur when they are cooled to 77°K are given in Table VI. Thermochromism may be caused by a change in phase or ligand geometry, in which case the color change is abrupt, or a narrowing and shifting of charge transfer bands into the ultraviolet region of the spectrum which causes a gradual change in color. There have been no previous reports of thermochromic titanium complexes. TABLE VI.--Thermochromic Titanium Complexes. Compound Color at 297°K 77°K TiCl3'CuH80 gray-green violet TiClu-ZCH3CN yellow white TiClu-2CuH8O yellow white (C2H5)uN[TiClu-2CH3CN] green blue I a '(C2H5)uN[TiCln]" salmon yellow aa material with non-integral stoichiometry. 73 Several other physical properties of the thermo- chromic complexes were qualitatively examined to observe if the color changes could be correlated with other changes. The optical spectra showed a shift of the lowest-energy charge transfer band toward higher energies when the compounds were cooled. The shift was especially noticeable in the material with non-integral stoichiometry "(02H5)uNETiClx(CAH80)yJ"’ obtained from (02H5)u[TiClA'2CuH8O]’ in which the charge- transfer band at 21,000 cm-1 at room temperature merged into the higher-energy bands. The charge-transfer band of TiCl3vCuH8O was located such that the compound was thermo- chromic. In TiCl °CAH802 and TiCl °CH ON the energies of the 3 3 3 first charge-transfer bands were respectively too high and too low for the compounds to exhibit thermochromism. No abrupt changes were found in the magnetic behavior of the thermo— chromic complexes as they were cooled. Changes in the esr spectra and the average g value of the complexes occured when several of the compounds were cooled. The spectra of TiClu'2CH3CN and TiClu°2CuH80 had a weak absorption peak at g=1.937 due to an unknown titanium(III) impurity. As the compound was cooled the spectrum of TiCl3-3CH3CN appeared and at 77°K the intensity of this signal was very high. A similar change took place with TiClu-2CuH80. The spectrum of "(C2H5)uN[TiClx(CuH80)yJ", which was prepared from (C2H5)uN[TiClu-2CHH80], and 74 (C2H5)uN[TiClu-2CH3CN], which appeared to have been partially oxidized, showed only a very sharp, single absorption peak g=2.00 at room temperature. When the compounds were cooled this peak gradually disappeared and the spectra of the respective (C2H5)uN[TiClu-2L] type compounds appeared. The changes were rapid and reversible as the temperatures were changed. In contrast, pure samples of the respective tetra- ethylammonium titanium complexes did not exhibit thermo- chromism or unusual esr spectra. It is felt that the thermochromism and the changes in the esr spectra of these compounds is somehow related to the presence of both titanium(III) and titanium(IV) in the samples though the connection is not apparent. The esr spectrum of TiCl3-CuH80 did not change except for the customary narrowing when the compound was cooled. BIBLIOGRAPHY 75 10. ll. 12. 13. BIBLIOGRAPHY Giggenbach, W., and Brubaker, C. H., Jr., Inorg. Chem. Bereman, R. D., and Brubaker, C. H., Jr., Inorg. Chem. 8, 2480 (1969). Wentworth, R. A. D., and Brubaker, C. H., Jr., Inorg. Chem., 1, A7 (1964). Rillema D. P., and Brubaker, C. H., Jr., Inorg. Chem., 9. 1655 (1969). Reagan, W. J., and Brubaker, C. H., Jr., Inorg. Chem., Duckworth, M. W., Fowles, G. W. A., and Hoodless, R. A., J. Chem. Soc., 5665 (1963). McDonald, G. D., Thompson, M., and Larsen, E. M., Inorg. Chem., 7, 648 (1968). Clark, R. J. H., Lewis, J., Machin, D. J., and Nyholm, R. S., J. Chem. Soc., 379 (1963). Fowles, G. W. A., Hoodless, R. A., and Walton, R. A., J. Chem. Soc., 5873 (1963). Fowles, G. W. A., Greene, P. T., and Lester, T. E., J. Inorg. Nucl. Chem., 29, 2365 (1967). Schumb, W. C., and Sundster, R. F., J. Amer. Chem. 300-. 22. 596 (1933). Fowles, G. W. A., "Progress in Inorganic Chemistry," Vol. VI, Cotton, F. A., ed., Interscience Publishers, New York, New York, 1964. Antler, M. and Laubengayer, A. W., J. Amer. Chem. Soc., 11. 5250 (1955). 76 1A. 15. l6. l7. l8. 19. 20. 21. 22. 23. 2A. 25. 26. 27. 28. 29. 30. 77 Ehrlich, P., and Siebert, W., Z. Anorg. Allg. Chem., 303, 96 (1960). Fowles, G. W. A., Hoodless, R. A., and Walton, R. A., J. Inorg. Nucl. Chem., 21, 391 (1965). Siddiqui, M. T., Ahmad, N., and Rahman, S. M. F., Z. Anorg. Allg. Chem., 336, 110 (1965). Clark, R. J. H., and Greenfield, M. L., J. Chem. Soc., A. “09 (1967). Edwards, D. A., Fowles, G. W.A., and Walton, R. A., J. Inorg. Nucl. Chem., 21, 1999 (1965). Giggenbach, W. and Brubaker, C. H., Jr., Inorg. Chem., 1. 129 (1968). Schlafer, H. L., and Gotz, R., z. Anorg. Allg. Chem., iéé, l (1964). Fowles, G. W. A., Lester, T. E., and Russ, B. J., J. Chem. Soc., A, 805 (1968). Fowles, G. W. A., and Walton, R. A., J. Chem. Soc., 4953 (1964). Winter, G., Inorg. Nucl. Chem. Lett., 2, 198 (1966). Seifert, J. J., and Thurn, F. M., Z. Anorg. Allg. Chem., 372, 79 (1970). Fowles, G. W. A., and Russ, B. J., J. Chem. Soc., A. 517 (1967). Kern, R. I., J. Inorg, Nucl. Chem., 23, 1105 (1962). Schlafer, H. L., and Gotz, R., Z. Phys. Chem. (Frankfurt am Main), 52, 97 (196A). Hartmann, H., Schlafer, H. L., and Hansen, K. H., Z. Anorg. Allg. Chem., 289, MO (1957). Schlafer, H. L., Lenz, W., and Staab J., Z. Phys. Chem. (Frankfurt am Main), 62, 290 (1968). Gutmann, V., Leitmann, 0., Scherhauger, A., and Czuba, H., Monatsh. Chem., 28, 188 (1967). 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 78 Hathaway, B. J., and Holah, D. G., J. Chem. Soc., 537 (1965). Ballhausen, C. J., and Gray, H. B., Inorg. Chem., l, 111 (1962). Figgis, B. N., "Introduction to Ligand Field Theory," Interscience Publishers, New York, New York, 1966. Jorgensen, C. K., Acta. Chem. Scand., 22, 73 (1957). Leihr,IL, and Ballhausen, C. J., Ann. Phys., 1; 304 (1958). Clark, R. J. H., J. Chem. Soc., 417 (1964). Jorgensen, C. K., "Absorption Spectra and Chemical Bonding in Complexes," Pergamon Press, New York, New York, 1962, p. 109. Clark, R. J. H., Spectrochim. Acta, 2;, 955 (1965). Figgis, B. N., and Lewis, J., "Progress in Inorganic Chemistry," Vol. VI, Cotton, F. A., Interscience Publishers, New York, New York, 1964. Earnshaw, A., "Introduction to Magnetochemistry," Academic Press, New York, 1968. Van Vleck, J. H., "Electric and Magnetic Susceptibilities," Oxford University Press, London, 1932. Cotton, F. A., and Wilkinson, G., "Advanced Inorganic Chemistry," Interscience Publishers, New York, 1966 p. 641. Kotani, M., J. Phys. Soc. Jap., 3, 293 (1949). Figgis, B. N., Trans. Faraday Soc., 51, 198 (1961). Dunn, T. M., J. Chem. Soc., 623 (1959). Earnshaw, A., Figgis, B. N., and Lewis, J., J. Chem. Soc., A, 1656 (1966). Lewis, J., Machin, D. J., Newham, I. E., and Nyholm, R. S., J. Chem. Soc., 2036 (1962). Klemm, W., and Krose, E., Z. Anorg. Allg. Chem., 253, 209 (1947). 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 79 Bereman, R. D., Ph. D. Thesis, Michigan State University, 1969. Kuska, H. A., Ph. D. Thesis, Michigan State University, 1965. Carrington, A., and Longuet-Higgens, H. C., Quart. Rev. (London), 15, 427 (1960). Gladney, H. M. and Swalen, J. D., J. Chem. Phys., 32, 1999 (1965). Bijl, D., Proc. Phys. Soc. (London), A63, 405 (1950). Bleany, B., Proc. Phys. Soc. (London), A63, 407 (1950). Bleany, B., Bogle, G. S., Cooke, A. H., Duffus, R. J., O'Brien, M. C. M., and Stevens, K. W. H., Proc. Phys. Soc. (London), A68, 57 (1955). Jarrett, H. S., J. Chem. Phys. 21, 1298 (1957). McGarvey, B. R., J. Chem. Phys., 28, 388 (1963). Waters, E. L., and Maki, A. H., Phys. Rev., 125, 233 (1962). Gair'yanov, N. S., Danilova, A. V., and Shagidullin, R. R., Opt. Spektrosk., 11, 212 (1962). Giggenbach, W. and Brubaker, C. H., Jr., Inorg. Chem., 1. 129 (1968). Giggenbach, W., unpublished results. Bereman, R. D., and Brubaker, C. H., Jr., J. Inorg. Nucl. Chem., in press. Herzog, S., Dehnert, J., and Luhder, K., "Technique of Inorganic Chemistry," Vol. VII, Jonassen, H. B., and Weissberger, A., eds., Interscience Publishers, New York, New York, 1968. Perrin, D. D., Armarego, W. L. F., and Perrin, D. R., "Purification of Laboratory Chemicals," Pergamon Press, Oxford, 1963. Carter, H. V., McClelland, B. J. and Warhurst, E, Trans. Faraday Soc., fig, 343 (1960). 66. 67. 68. 69. 70. 71. 72. 80 Walton, H. F., "Principles and Methods of Chemical Analysis," Prentice Hall, Inc., Englewood Cliffs, New Jersey, 1964, p. 236. International Union of Pure Applied Chemistry, Commission on Spectroscopy and Other Optical Procedures for Analysis, "Tables of Spectrophotometric Absorption Data of Compounds Used for the Colorimetric Deter- mination of the Elements," London: Butterworths, 1963. Figgis, B. N., and Lewis, J., "Modern Coordination Chemistry," J. Lewis and R. G. Wilkins, Ed. New York: Interscience Publishers, Inc. Vander Vennen, R. E., Ph. D. Thesis, Michigan State University, 1954. Figgis, B. N. and Nyholm, R. S., J. Chem. Soc., 4190 (1958). Foex, G., "Constantes Selectionnes: Diamagnetisme et Paramagnetisme. Relaxation Paramanetiquex," Masson & Cie, Paris, 1957. Bersohn, M. and Baird, J. C., "An Introduction to Electron Paramagnetic Resonance," New York: W. A. Benjamin, Inc., 1966.