* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Download THEORY AND IMPORTANCE OF OXYGEN
Survey
Document related concepts
Transcript
THEORY AND IMPORTANCE OF OXYGEN BRIDGE-BONDING B. JE0WSKA-TRZEBIATOWSKA Institute of Chemistry, The University of Wroclaw, Wroclaw, Poland ABSTRACT An oxygen atom as a donor bridge causes the bonding of two atoms at lower electronegativity. In a case of paramagnetic centres such as transition metal ions the magnetic criterion occurs, i.e. spin quenching follows. The MO model for the single oxygen bridge bonding in Ru4, W4, Mo5, Fe', Re4, Cr', Os4 dimers of d4, d', d', d' and d5 electronic structure were calculated using the SCCC method. This model gives a good explanation of magnetic properties of all the known complexes with one linear or angular bridge. The influence of other ligands on the ground term and on the complex properties has been explained. The structures most convenient to linear bridge formation are those of the d°—d4 electronic ones. At d5 the angular bridge is more energetic- ally convenient. Double oxygen-bridge bonding was studied on the large group of Mo—V dimers. The probable MO system explaining their magnetic properties is given. The spectroscopic characteristic in the infra-red of single and double oxygen bonding confirmed the bonding of the oxygen bridge. The nature of the bonding of oxygen to other atoms, and the role of the oxygen atoms as a bonding bridge have been two of the main problems with which investigators of the structure of chemical compounds and condensed phases have had to deal in recent years. Mutual interaction of atoms through oxygen in solid lattices has been known since the discovery of antiferromagnetism in metal oxide lattices. This effect has been qualitatively explained in the theory of superexchange proposed by Anderson', Kramers2 and Goodenough35. The turning point in the development of views on the role of oxygen was the discovery of molecular antiferromagnetism, that is, spin—spin quenching within one molecule. Melor6, Dunitz and Orgel7 found in 1953 that the diamagnetic ruthenium(iv) compound, the oxychlororuthenate(Iv) built of paramagnetic ions, is a dimer containing a linear Ru—-O-—-Ru core. Later in 1954, the second case of spin-quenching in metal ions by the oxygen bridge in a binuclear complex of rhenium -oxobis(pentachlororhenate) (iv) was discovered by B. Jeowska-Trzebiatowska and S. Wajda8. In 1962 Morrow9 demonstrated by x-ray diffraction that the core is in fact, linear 177°, forming the oxobridged Re—O——Re dimer. In the early 1950s, Pauling's model based upon the valence bond theory was commonly employed to explain the magnetic properties and the electronic structure of complex compounds. This model explained the magnetic properties of the ruthenium complex of a d4—d4 electron structure if the 89 B. JEOWSKA-TRZEBIAT0WSKA oxygen atom was assumed to form two bonds and two it bonds. It failed, however, to explain the properties of the rhenium complex with a d3—d3 electron structure. In the following years a number of metal ion systems have been discovered involving molecular quenching of electron spins by the oxygen bridge. In 1955 Mulay and Selwood'° reported such an interaction on increasing hydrolysis of ferric perchlorate solution which provided evidence that the spin-quenching is not a lattice effect. The studies of other authors yielded similar results1 1• An analogous phenomenon was observed by Sacconi and Cmi in 195412 during hydrolysis of molybdenum(v) complex solutions. A step forward in the study of the quenching effect was the discovery of a temperature singlet—triplet equilibrium between the states S = 1 and S = 0, made by Earnshow and Lewis13 and later by Jeowska-Trzebiatowska and Wojciechowski14 in chromium(iii) rhodochloride(j-oxohis [pentaamminechromium(in) chloride). Many examples of electron spin quenching have been found in oxobridged iron(III) complexes in recent years. First, Schiff's base iron(iii) compounds'5' 16 oxobridged binuclear iron(ni) complex with HEDTA'7' 18 (HEDTA = N-hydroxyethylenediaminetriacetate), phenanthroline oxobridged dimer'3' 16 and, of biological importance, oxobridged haemin dimer19 should be mentioned. The infra-red spectroscopic studies of the it bonds in inorganic compounds have also shown some peculiar properties of the bonding of oxygen to other atoms. Oxygen was found to be doubly or even triply bound, and the oxygen bridges were found to contain, at least partly, it bonding. The role of the oxygen bridge in electron transfer, hydrolysis and polymerization processes stimulated further studies of the oxygen bonding. Such studies have been taken in several world laboratories. In our Institute the mechanism of oxygen bridge formation, the nature of the oxygen bridge and the structure and properties of compounds with oxygen bonding have been studied, using different methods, for many years. The results of these studies induced me to put forward the idea that the oxygen bridges are so important in binding atoms that the term oxygen bonding' may be employed. Our results and consideration of the literature enabled me to make certain generalizations regarding the nature of the oxygen bond2022. The complexes containing the oxygen bonds could be considered as clusters—the compounds of various stabilities depending on the electronic structure, the size of the bridge-bound atoms and the nature of the other ligands. The oxygen bridge plays a similar role as a junction between the atoms as a hydrogen bridge; in contrast to the acceptor hydrogen bridge, however, it is a donor bridge. The ions linked by the oxygen bridge form a core of the complex in dimeric or polymeric clusters. Such a core may be linear or angular; the angle, as a rule, exceeds 90°. The angular bridges appear most frequently in nontransition elements, in transition elements with d° electronic structure, and, in some cases, in transition metal complexes as will be shown later for definite d electronic structures. The double bridge-oxygen bonds are, of course, also angular. Such clusters, for instance, where the oxygen has a coordination number of 3, are formed by chromium(ni) and iron(iii)23. The clusters containing 90 THEORY AND IMPORTANCE OF OXYGEN BRIDGE-BONDING cores with a single or double oxygen bridge arise either as intermediates of the salt hydrolysis prior to the formation of the final hydroxide or make the core of the complex compound. Each of the cations is surrounded by ligands, apart from the bridge oxygens, which complement its local symmetry up to the distorted octahedron or tetrahedron. The nature of the bridge bonding becomes most evident when the core consists of the paramagnetic cations Figure 1. The core containing the d-electron metal ion. Figure 2. Oxo-bridged dimer of the non-transitional ions. Figure 3. Angular oxygen bonding found in dinuclear complexes with one or two bridges. of transition metals. It may then be found experimentally that the oxygen bond makes mutual interaction of the electron spins possible, as in the case of a direct metal—metal bond. The magnetic criterion of the chemical bond appears. Such complexes are, therefore, most helpful in elucidating the nature of the oxygen bond. In dinuclear complexes consisting of the paramagnetic transition metal ions, these ions have, as a rule, a local distorted octahedral symmetry. The present paper will be, first of all, devoted to these types 91 B. JEOWSKA-TRZEBIATOWSKA of cluster structures. The conclusions obtained for the clusters consisting of paramagnetic ions may be generalized to cover compounds consisting of diamagnetic ions. The common occurrence and the role of oxygen-bridge bonds result from the fact that terrestrial media, both air and water, are rich in oxygen. The air Figure 4. The dimer with two oxygen bridges. The oxygen atom can also form a bridge between the three other atoms. Figure 5. Oxygen bridge of coordination number three. is a reservoir of 02 molecules which readily become negative ions with acceptance of electrons and water, a molecule of which is a simple uncharged HOH core which may either be changed into a trinuclear complex H3O or may dissociate, yielding 0H ions. Now, if we know what happens, the next question is why? Why should the oxygen atom be different from other bridge-forming donor atoms like sulphur, nitrogen or the halogens? This 92 THEORY AND TMPORTANCE OF OXYGEN BRIDGE-BONDING preferential position of oxygen results from a number of properties, such as its electronic structure, clectronegativity and the resulting ability of the oxygen ion to delocalize electrons and form a small covalent radius. Sulphur and nitrogen are also capable of bridge formation but could not compete with oxygen in this respect. While the electronegativity of sulphur with respect to the first electron is lower than that of oxygen, it is higher with respect to the second electron and hence its tendency to delocalize electrons is lower than that of oxygen. Table 1. Electronegativity of bridge elements24 Element Valence state Orbital Calculated electronegativity (eV) H s Li C Sp3 N O S' s2P3 s2p5 s2p4 s2p3 s2p5 s2p4 Cl sp5 0 01 F Cl Br I s2p4 s2p5 s21)5 6.92 2.92 s s s 13.70 6.52 8.18 —2.46 p p p p p p p p p p p 9.50 24.58 11.34 7.16 8.88 19.75 8.23 7.49 * Ref. 25. In addition, its larger size is a negative factor. The difference between nitrogen and oxygen is still more significant if we compare the electronegativities of the corresponding electronic structures: N and O, 8.18 and 24.58 eV respectively. This, and the fact that the electronegativity of nitrogen with respect to two electrons is negative, results in an excessive ability of the nitrogen to delocalize electrons. Nitrogen as a bridge should be therefore stabilized by means of a third acceptor, for instance by H which gives more stable NH or NH2 bridges. Only in exceptional cases can the nitrogen bridge be stabilized by metal ions of relatively high electronegativity. The halogen atoms have unfavourable properties with respect to single bridge formation. Comparison of the respective electronic structures 0 to F shows the difference in electronegativity of oxygen and fluorine, the radii of which are approximately equaL In our studies on oxygen bridges we have developed the magnetic method for investigation of the mechanism and determination of the equilibrium constants for the hydrolysis of some mononuclear complexes of rhenium(iv)26, osmium(iv)27, and molybdenum(v)283° in order to demonstrate that spin-quenching is a molecular phenomenon and not a lattice-type effect. 2[ReC16] 2± H20 ± [Re2OCl10] ± 2H ± 2C1 paramagnetic diamagnetic 2[OsCl6] 2- + H20 [Os2OCl10] + 2H + 2Cl paramagnetic diamagnetic 93 B. JE2OWSKA-TRZEBIATOWSKA 2[MoOBr5] 2- ± 4H20 [Mo2O4Br(H20)2] 2 + 6Br + 4W paramagnetic diamagnetic 4[MoOBr5]2 + 4H20 ± [Mo4O4Br,2(OH)4] ' + 8Br ± 4W diamagnetic paramagnetic Correlation of these data with spectroscopic results in solution and results obtained for the complexes in the solid phase enables us to define the structure of the polymerized molecules in solution. These studies have shown that in all cases, except for the osmium complex, the diamagnetic dimeric products of hydrolysis, and in the case of molybdenum the tetramers also, are formed in solution. The osmium dimer is unstable, and in fact exists only in the solid state. The magnetic properties have been studied in the solid state in the temperature range 77°—350°K. The dimeric linear or quasilinear complexes of rhenium8' 3 137 ruthenium38, osmium27, chromium13'14 and molybdenum39—42 obtained by us have been investigated (Table 3). These complexes are diamagnetic or exhibit a small temperature-independent paramagnetism. The chromium complex which shows a singlet— triplet temperature equilibrium is an exception. We have also taken into account the results obtained by other authors on iron'3" 5—19 and tungsten complexes43. In our opinion, the dimer of tungsten44 should be treated like the tungsten(Iv) complex with a d2—d2 electron structure. Now, a theoretical justification of the magnetic anomalies should have been found. We have attempted, as has Martin45, to apply the theory of superexchange in order to explain the diamagnetism of the dinuclear complexes. Any transposition of this qualitative method, however, from a collective spin system in the solid into a molecular system cannot give satisfactory results. In such cases this method may be considered as a prophetic one, but the qualitative picture it provides does not always remain in agreement with the experimental facts. For d3—d3 and d'—d' electronic structures of dimers, this theory would predict a weak paramagnetism which is true for the chromium complex only. In the case of the rhenium(iv) and molybdenum(v) complexes which exhibit only a low temperature independent pararnagnetism, this is inconsistent with the experimental results. The method of superexchange does not fully justify the magnetic properties and yields no information about the spectroscopic properties and the nature of the chemical bond. The electronic structure of the MOM and MM cores is most conveniently described by the molecular orbital method, which successfully explains the structure and properties of molecules forming the species of a definite nature. Studies based on the MO scheme were reported for the first time by Dunitz and Orgel7 in 1953 as well as by Jeowska-Trzebiatowska and Wojciechowski in 196346. The MOM core was treated as an isolated species, and only the interactions between the bridging oxygen and the acceptor ions linked to it were taken into account. The orbital schemes obtained in this way had no general significance because they did not explain all the magnetic anomalies found in dimeric compounds. Therefore, we have, with Natkaniec and Kozlowski47 once again proceeded to develop the orbital model, bearing in mind that in such complex systems as the dinuclear complexes in which 94 THEORY AND IMPORTANCE OF OXYGEN BRIDGE-BONDING each central ion of the core is surrounded, apart from the bridge-oxygen, by five other ligands of a local distorted symmetry, strong mutual interactions exist. Omission of these interactions, as in the previous orbital models, results in too serious approximations. Now we have taken into account three types of essential interaction: 1. Interaction between oxygen and both metal nuclei which is essential for the formation of oxygen bonding. The interaction follows the formation of a stable MOM core and is chiefly responsible for magnetic and spectroscopic properties of a new chemical species formed in this way. 2. Interaction between the metal ions and their local ligand environment excluding the bridging oxygen. The interaction should be considered as similar to that in mononuclear complexes. 3. Interaction between the hgands belonging to the two nuclei, in particular between the ligands placed in the xy planes perpendicular to the bridge axis. It should be emphasized, that the two latter types of interaction as was shown by our present studies, affect also the electronic structure of the core and its related physical properties. 4z IT li-v pz -li-v 7Th' Figure 6. Diagram illustrating molecular interaction. 95 B. JE2OWSKA-TRZEBIATOWSKA Let us consider a type L5—M----O—M—-L5 complex with a linear oxygen bridge of D4h symmetry. This type of structure appears most frequently in the transition metals: rhenium, ruthenium, osmium, chromium and molybdenum. Figure 7. Linear cluster. The MO calculations have been performed by us for the complex molecule as a whole by the self-consistent charge and configuration (SCCC) method48. The d orbitals of the central ion as well as the s and p orbitals of oxygen are transformed as bases of the following irreducible representations. Table 2 shows that the metal d. orbitals of e symmetry interact with the p, p, orbitals of the same symmetry, giving the it-bonding and it-anti-bond molecular orbitals. On the other hand, the combined -d2 orbitals of both metals with symmetries aiq and a2 overlap each other with the oxygen s and-p orbitals respectively, leading to the formation of molecular orbitals. Among the group of molecular orbitals discussed, one should distinguish the bonding and antibonding a2 and e orbitals as bridging ones. These orbitals are characterized by the considerable extent of electron delocalization. This is caused by a large overlap and similar energies of the atomic orbitals of oxygen and of metals of which these MOs consist. The aig orbitals consist of the oxygen and metal orbitals overlapping one another to a large Table 2. Orbital transformation in D4h symmetry Representations b28 b1 eq e Metal d orbitals Oxygen sp orbitals (1/,J2)(d3. + d,) — (1/J2)(d + d) (1/J2)(d + d) (1/J2)(d — d) (1//2) (dy: — d) p P big (1/j2)(d2_2 ± dci_yi) b2 (1/j2) (dXz_),i — d2_,2) ajg (1/j2)fd ±d2) S a2 (1/J2)(d. —d2) -p 96 THEORY AND IMPORTANCE OF OXYGEN BRIDGE-BONDING extent. However, due to the large difference in energies of the oxygen s orbital and metal d orbital, the extent of electron delocalization in these orbitals is insignificant. We have also taken into account the and it interactions of other ligands with metal ions. it orbitals of ligands with b1g and b2 symmetries overlap with the metal orbitals of suitable symmetries, that is, with combinations x Figure 8 (a) and (b). Oxygen—metal p — d,, interaction. of the d2_2 yielding MOs. As a result, the energy ordering of molecular rbitals has been obtained (Figure 10). Three orbitals aq, and e, consisting mainly of the oxygen orbitals, are occupied by the oxygen valence electrons. The a2 and e orbitals act as the main 'binder' in oxygen bond formation. The other orbitals of this ordering consist mainly of the metal ion d orbitals. The metal electrons are placed in these orbitals in the energy ordering. Thus, the oxygen bonding makes possible the pairing of electrons stemming from both metal atoms in the same molecular orbital. 97 P.A.C—27/1_--E B. ThZOWSKA-TRZEBIATOWSKA x z x Figure 9. The overlap between the s, p orbital of oxygen and the d2 orbitals of metal atoms. The orbitals; 1%, e and ht are close to each other in their energies and their ordering is determined by interactions between the ligands of both nuclei. The sequence of these orbitals determines their magnetic properties. The energy interval of the next e orbital is higher by one order. The b2 and biq as well as a2 and aiq orbitals are close to each other in energy. The b29, e0 b1 orbital ordering is shown in the scheme (Figure 10) corresponding to the medium or even weak ic-bonding halogen, thiocyanate and other aig // ___ 2u /7 (nl)d ___ N lu —---- \ _tac LJk; (p.py)O \\\N\N \N \ \ (Pz) \ N \PsO \s-o \Ulg AO of ligonds AO Of metal Figure 10. Molecular orbital diagram for a linear dinuclear transition metal complex with oxygen bridge of D4h symmetry. 98 THEORY AND IMPORTANCE OF OXYGEN BRIDGE-BONDING ligands. For the a-only bonding ligands, this ordering is altered. For instance, in the case of the NH3 ligands appearing in p-oxo-bis[pentaamminechromium(iii)] salt, (NH3)5(CrOCrNH)54, there are in fact no interactions between the ligands of two nuclei coordinated in the plane perpendicular to the line connecting the two nuclei. As a result, the b2g and b1 orbitals become very close in energy, causing a change in the sequence of orbitals to give the ordering in the RHS of Figure 11, which determines the magnetic properties. The orbital scheme obtained by us made it possible for the magnetic properties of all known dinuclear complexes with oxygen bridge of d electron metals with D4h or similar symmetries to be explained (Table 3). [CI5ReOReCI5]4 [(NH3)5CrOCr (NH3 )5]4 e E b1 b1 '2g eg b Figure 11. Diagrams showing MO' sequence dependence on the character of the ligands. The dimers of rhenium, molybdenum, ruthenium and osmium have a diamagnetic singlet ground term 1A1 which results from the completely filled orbitals. The bridge complex of tungsten has a triplet ground state 3A2g resulting from incomplete filling of the eg orbitals. Small differences in energies of the b2g, eg and b1 orbitals and a possibility of electron transition results in an increase of the moment from 1.72 B.M. at 76°K up to 2.40B.M. at 2300K28 per tungsten atom. The dimer of chromium(in) exhibits a singlet—triplet transition. This complex is paramagnetic and its moment is 1.85 B.M. at 293°K while at low temperatures it decreases, almost to zero at — 125°K. This is due to the fact that the energy gap between b2g and b1 is very small and the distance between the corresponding terms is therefore comparable to k T The singlet ground term 'A1 changes into the triplet state 3A2 with simultaneous filling of the bI orbital. I would like to draw attention to the temperature-independent paramagnetism which appears in the dimers possessing a diamagnetic singlet ground state. This temperature-independent paramagnetism depends, of course, on the distance between the interacting levels, that is, the distance between the nearest excited state and the ground state, and on the shape of 99 B. JEOWSKA-TRZEBIATOWSKA Table 3. Magnetic properties of dinuclear complexes of transition metals with oxygen bridge Electronic structure Compound T(°K) Ueff XM )< 106 average values K4[Re2OC110] (b2g)2 (e9)4 (b1)° K4[Re2OC110 H20] K4[Re20(C204)2(OH)6] K4[Re20(C204)4(OH)2] (b2)2 (eq)4 (b1)° (b29)2 (e)4 (b1)° 450 220 diam. (8) diam. (26) diam. (32) diam. (32) 1.73 (43) 2.40 diam. (38) diam. (27) diam. (13,14) 1.8 (13,14) (a) diam(13) (b2g)2 (eq)2 (b1)° K4[W20C110] (b2q)2 (eq)2 (b1j° (b2q)1 (e9)2 (b1)1 K4[Ru2OC110] (b2g)2 (e9)4 (b1)2 K4[Os2OCl1o] (b2g)2 (e9)4 (b1)2 [Cr20(NH3)10]C14 (e6)4 (b2q)2 (b1)° (e9)4 (b2g)1 (b1)' (a) [Fe20(Ph)4(H20)2X4] (b1)2 (a2)2 (b2)2 (a2)2 (a1)2 (b1)° (I) (b1)2 (a2)2 (b2)2 (a2)2 (b) (Fe Selen)20P3,2 (a1)1 (b1)1 (II) (139°) equilibrium 1—11 80—300 80—300 —300 —540 76.5 293 200 — 160 126 293 276 (b)diam(15,16) (a) 1.76(13) (b) 1.87(15,16) (a) 0.77 (13) (b) 0.58 (15,16) 80 their wavefunctions which enter into the matrix element of Zeeman interaction. In the linear bridges, where the d'—d1 and d3—d3 electron structures are involved as molybdenum(vi) and rhenium(iv) complexes, the distance between the filled orbital and the excited orbital is of the order io cm For this . reason these complexes show rather normal temperature-independent paramagnetism. a1 a ———— _______ * ____________ 2u ____________ b2 big ____ bf b1 eg Sg D4h a1 / ______ — — —— a24 + b2tH, b14 f. C2j, Figure 12. Orbital splitting in dinuclear Fe(nI) complexes with the bent bridge. 100 THEORY AND IMPORTANCE OF OXYGEN BRIDGE-BONDING In the d4—d4 ruthenium and osmium dimers the energy separation between the strong n-antibonding bridging e orbital and the completely filled b, iO cm '.Therefore, the temperature-independent orbital is large, range paramagnetism is negligible, viz, less than 200 x 106 3s The d5-configuration appearing in the iron(IIi) dimers exhibits peculiar properties. On the basis of the orbital scheme, two electrons should be placed into the strong antibonding bridging e orbital. The structure of a bent bridge will therefore be more favourable, corresponding to the altered hybridization with the increased p-orbital contribution to the bonds. This results in weaker d—p interactions which cause the splitting of e orbital into two orbitals. Meanwhile, the symmetry is changed into C2,. At the small bend of the Fe—O——Fe bridge'5'9 the splitting is also small and the distance between the a and bT orbitals is small; the singlet—triplet equilibrium and tempera- ture dependence of the moment are observed. The diamagnetic singlet ground state 'A1 of the a, b configuration is changed, on increasing the temperature, into the triplet state 'B, of the b,° configuration. The interpretation of magnetic properties based upon our orbital scheme gives clear evidence that the linear MOM bridges may be found only for d°, d1, d2, d3 and d4 configurations in complexes. In the case of the d5 con- figuration an angular structure of the bridge becomes more convenient. Figure 13. Model of angular cluster. With the increase of electron filling to d6, d7 and d8 configurations, the formation of a bridge becomes less probable since the successive electrons should have been placed in the strong antibonding bridging e orbital and then, in the antibonding orbitals. Such a system would, of course, be very unstable. Now it is necessary to discuss the properties and structure of dimeric clusters with two oxygen bridges, taking as an example the large group of molybdenum(v) complexes synthesized and studied by Rudolf and me. The d'—d' electron structure seemed to be most convenient for the examination of the properties of these more complex systems. These compounds have been obtained as hydrolysis products of the monomeric oxyhalogenomolybdates. Magnetic studies carried out in solution and in the solid phase have shown that dimers or higher polymerized species are formed in which the molybdenum atoms are linked by double oxygen bridges3942. We have isolated about twelve dinuclear di-t-oxoha1ogenomolybdates(v), and a new type of molybdenum complex: tetranuclear molybdenum(v) clusters. I shall deal only with dimers possessing twO oxygen bridges. 101 B. JE2OWSKA-TRZEBIATOWSKA Table 4. Magnetic susceptibilities of Mo(v) dinuclear cornplexes XM x 106 Complex (PyrH)2[Mo2O4C14(H20)2] (Met4N)2[Mo2O4C14(H20)2] (PyrH)2[Mo204C14(H20)2]H20 (NH4)2[Mo2O4Cl2(OH)2(H20)2] (PyrH)4[Mo204C12(OH)4] (PyrH)2[Mo2O4Br4(H20)2] (PyrH)4[Mo204(SCN)6] K2[Mo204(C204)2(H20)2] 3H20 Ba[Mo204(C204)2(H20)2] 31120 207 328 165 157 260 185 448 115 107 The magnetic studies on the molybdenum(v) dimers have shown that their electron spins are almost completely quenched. Only a small, temperature-independent paramagnetism remains3942. The dipyridyl complex [Mo2O4C12(dipy)2], showing a typical dependence of magnetic susceptibility of temperature and a magnetic moment of about 0.9 B.M., is an exception49. Figure 14. Model of cluster with two oxygen bridges. Comparison between the systems with a 'double oxygen bonding' and the system with a 'single oxygen bonding' suggests that the presence of two bridging oxygens and also of strong it bonding ligands as terminal oxygens results in the weakening of the bridge it bonds. There is then a simultaneous increase in the metal—oxygen interactions and a change in the geometry from D4h into C2h which results in a further weakening of the metal—oxygen it bonds. Due to the decrease in the molybdenum—molybdenum distance a direct metal—metal interaction appears. The x-ray data obtained by Cotton5° for K2{[Mo02(C204)H20]20} indicate that this distance is small, about 2.4 A. Due to the presence of two bridging oxygens, an additional oxygen—oxygen interaction also appears. Let us now consider the molecular model for a dimeric molybdenum cluster of C21, symmetry, taking into account all these interactions. The terminal ligands oxygen, water and halogens are also of significance and disturb the electronic structure of the core. In such a system as shown in Figure 15, the d orbitals of the Mo(v) ions and the sp orbitals of the bridge 102 THEORY AND IMPORTANCE OF OXYGEN BRIDGE-BONDING tz 1z CL Figure 15. The structure and stereochemistry of [Mo2O4C14(H2O)2]2. oxygens make up the basis of irreducible representations of the point group C2h. The combinations of the d..2_)2 orbitals of that group form the type bonds along the z and x, y axes, respectively. While the other d orbitals contribute to the molecular it orbitals, the orbitals in the bridge plane and orbitals lie in the planes perpendicular to them. Because of the the short metal—metal distance, a c Mo—Mo interaction through the dX), orbitals is possible here. If the it interaction in the bridge plane is assumed to be weaker, owing to the angular character of the bridges, as shown above, the following energy ordering of the d orbitals should be expected. The magnetic properties of such compounds are determined by the sequence of molecular it orbitals consisting mainly of the metal d orbitals. Table 5. The linear combination of atomic orbitals for the C2h point group in dinuclear complex ag (1/\/2)(dX. + d) (1/2)(d. + d. — b a bg ('/2)( — p,.+ p — (1/j2)(dz + d2) — (1/2)(s + s') d;2) (l/2)(p + p + p' + ji,) (1/,j2)(d2 — d2) (1/.j2)(d,, — di,,) (1/\/2Xp, + i"2) (1/2)(d2 + d. + d. + d2) (l/2)(p — — p' + i';) 1j2(d — d,. + d'- — d) (1/J2)(s — s') (t/2)(.p ± py Pc — (1/j2)(d2_2 +d_y2) (1i2xd — — d ± d;4 (1/'2Xp. — (1/,,J2Xd22_2 — d2_2) 103 B. JEOWSKA-TRZEBIATOWSKA The energy intervals between the ag(xy) and bjxy) orbitals in the above molybdenum(v) complex are sufficiently large, due to the c Mo—Mo interactions, for the compound to be diamagnetic (ground term 1Ag). If, however, instead of oxygen, another ligand of a it acceptor nature or one that does not give any it bonds, is introduced into the molecule, then the separation of the d.) would be very small and the compound lowest orbitals ag(dxy, dr..), would exhibit a temperature-dependent paramagnetism (ground term 3Am). 2 cxOg(dz) comb SaOb O3H2O oCt abu(dz) comb POb Ot 0H20,UCL aa(d2-y2) comb P6Qt, crbg(di-y2) comb P00b 0C1 0CL E lTbu(dxzdyz) comb zObTOt 7TCI fdxzdyz) comb 1TXObtOt rCt comb 7r0 ITCL rag(dxzdyz) comb rO 7TCL lTbu(xy) 1 comb ITfrY)Ob irCt Tfry)Ob ITCI QM0-Mo lrag(xy) J comb Figure 16. Molecular orbital scheme for dinuclear molybdenum complex with two oxygen bridges. The paramagnetic properties of the dipyridyl complexes may be explained by the appearance of the 3A term. Lowering of the magnetic moment with respect to one electron spin per molybdenum atom may result from the equilibrium existing between the paramagnetic and diamagnetic isomers. Vibrational spectra make it possible to draw decisive conclusions on the bridge structure. There is a choice between the linear and angular structures and it is now possible to determine the bond order and bond strength, and hence, the it interactions between the bridge oxygen and the atoms linked to it. The extensive investigations by infrared spectroscopy of different compounds with the oxygen-bridge, carried out by Hanuza and me and obtained from the literature5 054 allow further extension of the oxygen bridge theory. The linear and angular structures of the bridge described as a 'three body system differ in the number of infra-red and Raman active frequencies. The normal modes of vibration given in Table 6 for the linear and angular MOM systems show that for the complexes with angular bridges, all three types of frequencies are both infra-red and Raman active while the linear bridge complexes should not exhibit any symmetric stretching frequencies in the infra-red. The dimers with two oxygen bonds treated as a tetratomic 104 THEORY AND IMPORTANCE OF OXYGEN BRIDGE-BONDING 0 () Dipy Figure 17(a) and (b). Geometrical isomers of [Mo2O4C12(dipy)2]. 105 B. JE2OWSKA-TRZEBIATOWSKA Table 6. Normal modes of vibration of MOM core Assignments Dh symmetry-linear bridge Normal vibration Vs(MOM) M M 0 0 OM 0 0-' 4-0 0 M 0 O—O M 4-0 Vas(MOM) M 4-0 (MOM) Representation Activity Frequency range (cm 1) Ra 200—350 i.r. 750—900 i.r. 50—100 9 Bent bridge C2 symmetry Normal vibration Vs(MOM) VaS(MOM) MOzM ö(MOM) Representation Activity Frequency range (cm ') A1 i.r., Ra 400—600 B2 i.r., Ra 750—900 A1 jr., Ra 100—220 system with a D2h symmetry possess six types of normal modes of vibration. Among them, three are infra-red active: B1, B2, B3. Two strong bands corresponding to the stretching frequencies of the double bridge lie in the ranges 450—670 cm1 and 560—830 cmt (see Table 9). Arguments based upon the molecular orbital scheme suggest that linear bridges are formed, above all, by the d° to d4 electronic structure of metal ions with local distorted octahedron symmetry and a total symmetry of the complex ion D4h. Table 7 contains these types of dinuclear clusters; the values obtained from infra-red studies of various compounds with different outer ions are shown as frequency ranges. The appearance of some infra-red inactive symmetric frequencies over the range 200—250 cm' follows from small distortions of the linear bridge. This is particularly evident for the osmium and ruthenium complexes where the introduction of mixed ligands results in a certain change in symmetry. With the bent bridge, the symmetric frequencies become active in the infra-red according to group theory with simultaneous considerable increase in vibrational energy. The symmetric frequencies are situated, however, in the same range as those of the linear bridges. 106 THEORY AND IMPORTANCE OF OXYGEN BRIDGE-BONDING v(MO)Ag ring deform.B10 v(MM)Ag v(MO)Big v(MO)820 v(MO)830 A Figure 18. Norma! modes of vibration of M202 core. The increase of the symmetry frequencies follows from the decrease of the bridge bond order due to the reduced interactions between the oxygen p orbitals and the p or d5 orbitals of the central ions. This is because the hybridization of oxygen varies with changing angle from 180° to 120° from the diagonal sp into the trigonal sp2 and at 109° into the tetrahedral sp3. As the oxygen bridge bond angle diminishes, the it-interaction in the bridge decreases. Reduction of the angle leads to the formation of a double-bridge bonding, that is to the two oxygen bridges, which appear in the di-t-oxocomplexes. Table 7. Vibrational frequencies of linear oxygen bridge MOM (angle 175—180°) Compound [Re2OC110]4 [Re2O3en2Cl4] [Re2O3py4Cl4] Vas(M0M) Vs(M0M) 850—855 730 700 752 200* K6[Re203(CN)8] [Os2OCl10:14 [0s20Br10]4 [Os2OI10]4 838—850 843—845 835—841 [Os2OCl6Br4]4 [Os2OCl8I2]4 846—848 830—832 [Ru2OC110]4' [Ru2OBr1 ]4_ [Ru2OCI8Br2]4 [Cr20(NH3)10]C14 (NMe)4[Zr2OC110] [Mo203(OX)4]Y2 880—885 C74H42N18Mn'O•2py 200* 200* 200* 205 205 200 204* 213_215* 750 230 230 245* 215 213 860 820 — 858—86 1 852 873 Refs a,b,c,d,e a a a,b,d,e a,b a a,h a a.h,c,d,e abc, a c,d g 6 ii, B. Jezowska-Trzebiatowska. J. Hanuza and M. Baluka. 4cta Phys. Polonica, in press; Specsrochimica Acta, in press. W. P. Griffith, J. Chem. Soc. A. 211 (1969). R. M. Wing and K. P. Callahan, borg. Chess. 8,871 (1969). D. J. Hewkin and W. P. Griffith, J. Chem. Soc. A, 472 (1966). F. A. Cotton and R. M. Wing, borg. Chess. 4. 867 (1965). A. Feltz, Zeit. Anorg. Ailgem. C/tern. 358, 21(1968). P. C. M. Mitchell, J. Chess. Soc. A, 146 (1969). J. A. Elvidge and A. B. P. Lever, Proc. Chern. Soc. 195 (1959). L. H. Vogt, ir, A. Zalkin and D. H. Templeton, Science, 151, 569 (1966). * observed in jr. spectra. 107 B. JE±OWSKA-TRZEBIATOWSKA Table 8. Vibrational frequencies of bent oxygen bridge Compound VSSIM0M) Ge2OFf Sn2OF Sn20(CH3)6 As2OF Sb2OF' P2O As2O' SOS Vs(M0M) 8(MOM) B2 A5 A1 858 820—750 465—490 — — — a a 200—210 c 458 737 415 889—899 859—883 890—980 733—780 760—820 530—538 435—473 682—730 530—560 260—350 558 503 Refs b 200—300 240—245 125—160 d — — — g h Cr2O V2O' 772 810 790 Se2O Si2OCI6 557—670 491—516 170—250 1132 421 100 ' c <200 556 e f I j k B. Je±owska-Trzebiasowslca, J. Hanuza and W. Wojciechowski. Spear. Ada, 23A, 2631 (1967). H. Kriegsmann and S. Pischtschan, Z. Anorg. Allgem. Chess. 315.283(1962). B. Jeiowska-Trzebiatowska and J. Hanuza, Spectr. Acm. in press. E. Steger and G. Leukroth, Z. Assorg. Aligem. Chess. 303. 169 (1960). B. Je±owska-Trzebiasowska and 3. Hanuza, in preparation. R. J. Gillespie and F. A. Robinson, Spear. Acme, 19, 741 (1963). H. Trammreich, D. Bassi, 0. Sala and H. Siebert, Spear. Acme, 13, 192 (1958). W. P. Griffith and T. 0. Wiekins, J. Chem. Soc. A, 1087 (1966). R. M. Wing and K. P. Gallahan, Inorg. Chess. 8, 871 (1969). A. Simon and R. Paetzold, Z. Anorg. Allgern. Chess. 303, 39 (1960); 353, 53 (1967). J. F. Griftiths and D. F. Sturman, Specer. .4cma, 25k, 1415 (1969). Table 9. Vibrational frequencies of double oxygen bridge M Compound [As2O2F8]2 V5(M2o2) V6(M2o2) A19 B39 — — [Sb2O2F8]2' [Mo204(EDTA)]2 759 [Mo2O4C14(H2O)2]2 729 [Mo2O4Br4(H20)2]2 [Mo202(OH)2Br2(OH)4]2 [Mo2O4CI2(OH)4]4" [Mo202(OH)2C14(OH)2] [Mo204ox2(H2O)2]4 K4[Os2O6(N02)4] K4[0s206ox3] [Ru2O6(NH3)4] [W204(NCS)6]4 [1208(OH)2]4 590—605 560—570 736 705 — — 720 739 510 520 745 — — — 654 V5(M2o2) B29 N0Z V4(M2o2) 470—490 291—315 450—470 635 675 490 567 Rels. B5 — 314 296 — — — a a b b, e e e e e 730 742 722 829 480 799 830 503 243 240 490 200 C 483 335 332 d d — 757 498 612 530 613 B. Je±owska-Trzebiatowska and J. Hanuza, Spectr. Acme, in press. R. M. Wing and K. P. Callahan, lnarg. Cheiss. 8, 873 (1969). B. Je±owska-Trzebiatowska, 3. Hanuza and M. Batsika, Acme Phys. Polonica, in press. W. P. Griffith, J. Chess. Soc. A, 211(1969). B. Je±owska'Trzebiasowska and M. Rudolf, to be published. 108 70NM — — b, e c c M X4 Vus(M0M) 0 * V(M_oH) VB3(M202) VB2(M202) VS(MOM) X or Y can be CI, Br, I, ox/2, py, en2, NO2. M(OH)X5 4X M ,/\ 0 VUS(M02) X5MOMX5 VS(M02) M02X4 trans VUS(M03) V503) M03X3 CiS VMO of compound vibrationalmode MOX5 MOX3Y2 Type order Bond fr cm 1 — 550—570 1.0—1.2 800—830 503—530 204—215 2.82—2.99 448—503 3.30—3.87 830-848 775835 720—855 200—205 952 915 889—905 6.06—6.83 799—843 6.96 890—912 890 1.2—1.3 1.3—1.6 2.0—2.1 2.0—2.3 1 j Osmium — cm1 Jr Ruthenium 1.84 2.6—3.25 3.90-4.15 — — 528—584 2.40—2.92 2.8—3.0 720—760 490—670 490 235 830 3.05—3.30 5.2—5.3 6.0—6.71 850—860 220—230 982—910 824—839 820—830 852—885 2.8—3.3 Jr 930—1020 7.20—7.8 cm' Molybdenum 6.09-6.70 807—822 5.12—5.38 770-780 912—993 7.91—8.55 952—960 8.01 cm Rhenium Correlation between the metal—oxygenfrequencies, bond order and force constants [Md/A] 2.2—2.5 Table 10. — 7.0—8.0 J 520—560 860—870 210—220 2.4—2.9 3.2—3.6 980—1020 7.1—7.4 — 950—960 cm Chromium C rn Q rn 0 0 rj rn z (1 0 0 rn B. JEOWSKA-TRZEBIATOWSKA The angle may approach 900 (Table 9). The bond order in such bridges decreases and the bond order has values intermediate between those characteristic of metal—oxygen c bonding and that existing in the linear bridge MOM. Infra-red spectroscopy is the technique which makes possible quantitative studies of the effect of ic-interaction in all oxycompounds. In particular, a correlation between the bond order and the force constants calculated from the frequencies for a number of oxycompounds with various functions of the oxygen is a good illustration for the varying contribution of it-interactions in the bonding. Table 10 contains the observed frequencies and the force constants calculated from them for various types of oxycompounds formed by the transition metals with different ligands which have been discussed above, with respect to their magnetic properties and the theoretical interpretation of their electronic structure on the basis of the MO scheme. A wide range of force constants for oxygen bridges in rhenium complexes results from taking into account the clusters with strong ic-bonding ligands beyond the oxygen bridge. Higher values of force constants are associated with the weak it-bonding ligands such as the halogens. The force constant values obtained are a linear function of the bond order. It follows from these investigations that the contribution of the ic-interaction is largest for the mononuclear mono-oxycomplexes; this decreases, however, with the increasing number of oxygen atoms in the molecule. The it-interaction is, in turn, much lower in the oxygen bridge and extremely low in a double oxygen bridge. It diminishes in the following order in the series of oxycompounds: MOX5 > /o\ MOX6> MOM > M /M > MOH > MOH2 It seems that the correlation of studies by means of several methods, principally magnetic and infra-red spectroscopy, with simultaneous theoretical interpretation by the molecular orbital method gives the clearest idea of the structure and character of oxygen bonding. REFERENCES P. W. Anderson, Phys. Rev. 79, 350 (1950); 79, 705 (1950); 86, 964 (1953). H. Kramers, Physica, 1, 182 (1930); 18, 101 (1952). J. B. Goodenough and A. L. Loeb, Phys. Rev. 98, 391 (1955). J. B. Goodenough, Phys. Rev. 100, 564 (1955). J. B. Goodenough, Magnetism and Chemical Bond, Wiley: New York (1963). 6 D. P. Mellor, J. Proc. Roy Soc. Wales, 47, 145 (1944). J. D. Dunitz and L. E. Orgel, J. Chem. Soc. 2594 (1953). 8 B. Jeowska-Trzebiatowska and S. Wajda, Bull. Acad. Polon. Sci. 2, 249 (1956). J. C. Morrow, Acta Cryst. 15. 851 (1962). 10 L. N. Mulay and P. Seiwood, J. Am. Chem. Soc. 77, 2693 (1955). B. Werbel, V. H. Dibeler and W. C. Vosburgh, J. Am. Chem. Soc. 81, 1033 (1959). 12 L. Sacconi and R. Ci, J. Am. Chein. Soc. 76, 4239 (1954). 13 A. Earnshow and .1. Lewis, J. Chem. Soc. 396 (1961). 14 B. Jeowska-Trzebiatowska and W. Wojciechowski, J. Inorg. NucI. Chem. 25, 1977 (1963). 2 110 THEORY AND IMPORTANCE OF OXYGEN BRIDGE-BONDING C. Scheringer, K. Hinkler and M. V. Stackelberg, J. Anorg. Chem. 306, 35 (1960). J. Lewis, F. E. Mebbs and A. Richards, Nature, London, 207, 855 (1965); M. Gerloch, J. Lewis, F. E. Mebbs and A. Richards, Nature, London, 212, 809 (1966); J. Lewis, F. E. Mebbs and A. Richards, J. Chem. Soc. A, 1014 (1967); M. Gerloch and F. F. Mobbs, J. Chem Soc. A, 1598 (1967); M. Gerloch and F. E. Mobbs,J. Chem. Soc. A, 1900 (1967); M. Gerloch, J. Lewis, F. E. Mobbs and A. Richards, J. Chem. Soc. A, 112 (1968); K. S. Murray, A. von der Bergen, M. J. Connor, N. Rehak and B. 0. West, Inorg. NucI. Chem. Letters, 4, 87 (1968). 16 G. M. Bancroft, A. C. Maddock and R. R. Pandl, J. Chem. Soc. A, 2939 (1968). 17 H. Schugar, C. Walling, R. B. Jones and H. B. Gray, J. Am. Chem. Soc. 89, 3712 (1967). 18 s J. Lippard, H. Schugar and C. Walling, Inorq. Chem. 6, 1825 (1967). 19 S. B. Brown, P. Jones and I. R. Lantzke, Nature, London, 223, 1960 (1969). 20 B. Jelowska-Trzebiatowska, Theory and Structure of Complex Compounds, p 1. Pergamon Press: London; WNT: Warszawa (1964). 21 B. Je±owska-Trzebiatowska, Exper. Suppl., 9, 128 (1964). B. Jeowska-Trzebiatowska, J. Chim. Phys. 765 (1964). 22 B. Jelowska-Trzebiatowska, Coord. Chem. Rev. 3, 255 (1968). 23 B. M. Figgis and C. B. Robertson, Nature, London, 205, 694 (1965). 24 G. Klopman, J. Chem. Phys. 43, 5124 (1965). 25 Hinze and H. H. Jaffé, J. Am. Chem. Soc. 84, 540 (1962). 26 B. 27 28 29 30 31 Jeowska-Trzebiatowska, J. Mroziñski and W. Wojciechowski, Roan. Chem, 39, 1187 (1965). B. Jeowska-Trzebiatowska, J. Hanuza and W. Wojciechowski, J. Inorg. Chem. 28,2701(1966). B. Jeowska-Trzebiatowska and M. Rudolf, Roczn. Chem. 41, 453 (1967). B. Je±owska-Trzebiatowska and M. Rudolf. Roan. Chem. 41. 1879 (1967). B. Je±owska-Trzcbiatowska and M. Rudolf. Roczn. Chem. 42, 1227 (1968). B. Je±owska-Trzebiatowska and S. Wajda, Mémoires présentés a Ia Section de Chimie, XVI IUPAC Congress, Paris (1957). 32 B. Je±owska-Trzebiatowska and S. Wajda, Bull. Acad. Polon. Sc!., Cl. III, 6, 217 (1958). B. Je±owska-Trezbiatowska and S. Wajda, Bull. Acad. Pol. Sc!., Ser. Sc!. Ch!m. 9, 57 (1961). B. Je±owskaTrzebiaeowska, S. Wajda and W. Wojciechowski, Bull. Polon. Sc!. Ser. Sc!. Ch!m. 9, 65 (1961). B. Jelowska-Trzebiatowska, S. Wajda and W. Wojciechowski, Bull. Acad. Polon. Sc!. Ser. 36 Sc!. Ch!m. 9, 767 (1961). B. Je±owska-Trzebiatowska and W. Wojciechowski, Bull. Acad. Polon. Sc!. 9, 685 (1961). S. Wajda and F. Pruchnik, Roczn. Chem., 41, 1169 (1967). 38 B. Je±owska-Trzebiatowska, R. Grobelny and W. Wojciechowski, Bull. Acad. Polon. Sc!. " Ser. Sc!. Ch!m. 12, 827 (1964). B. Je±owska-Trzebiatowska, W. Wojciechowski and M. Rudolf, Chem. Zvest!, 19, 229 (1968). 40 B. Je2owska-Trzebiatowska and M. Rudolf, Bull. Acad. Polon. Sc!., Ser. Sc!. Chim. 17, 419 (1969). 41 B. Je±owska-Trzebiatowska and M. Rudolf, Roczn. Chem. 44 (1969). 42 B. Je±owska-Trzebiatowska and M. Rudolf, Roczn. Chem. 44 (1969). ' ' R. Colton and G. G. Rose, Austral. J. Chem. 21, 883 (1968). E. König, Inory, Chem. 8, 1278 (1969). R. Martin, 'New Pathways !n Inorgan!c Chemistry', p 175. Cambridge University Press: London (1968). 46 B. Je±owska-Trzebiatowska and W. Wojciechowski, Zh. Strukt. Kh!m. 4, 872 (1963). B. Je±owska-Trzebiatowska, L. Natkaniec and H. Kozlowski, Bull. Acad. Polon. Sc!., in press. 48 C. J. 50 52 55 Ballhausen and H. B. Gray, Jnorg. Chem. 1, 111(1962). B. Je±owska-Trzebiatowska and M. Rudolf, unpublished results. F. A. Cotton, S. M. Morehouse and J. S. Wood, Jnorg. Chem. 3, 1603 (1964). B. Je2owska-Trzebiatowska, J. Hanuza and W. Wojciechowski, Spectroch!m. Acta, 23A, 2631 (1967). B. Je±owska-Trzebiatowska and J. Hanuza, Spectroch!m!ca Acta, in press. B. Je±owska-Trzebiatowska, J. Hanuza and M. Baluka, Acta Phys. Polon. 38, 70 (1970). B. Jetowska-Trzebiatowska, J. Hanuza and M. Baluka, Acta Phys. Polon. in press. H. J. Schugar, A. T. Hubbard, F. C. Auson and H. B. Gray, J. Am. Chem. Soc. 31, 71(1969). 111