Chapter 2 Atomic Structure CHAPTER 2: ATOMIC STRUCTURE 2.1 a. = h 6.626 10 34 J s = 2.426 10 11 m mv (9.110 10 –31kg) (0.1) (2.998 108 m s –1 ) b. = h 6.626 1034 J s = 6 10 34 m 3 mv 0.400 kg (10 km/hr 10 m/km 1hr/3600s) c. = h 6.626 10 34 J s 35 m = -1 9.1 10 mv 8.0 lb 0.4536 kg/lb 2.0 m s d. h 6.626 10 34 J s = = mv 13.7 g kg/10 3 g 30.0 mi hr -1 1 hr/3600s 1609.3 m/mi 3.61 10 33 m 2.2 1 hc 1 1 E RH 2 – 2 ; RH 1.097 10 7 m–1 1.097 10 5 cm–1 ; E __ nh 2 nh 4 1 1 12 E RH – RH 20,570cm –1 4.085 10 –19 J 4 16 64 4.862 10 –5 cm = 486.2 nm nh 5 1 21 1 E RH – 23,040cm –1 4.577 10 –19 J ; RH 100 4 25 4.34110 –5 cm = 434.1 nm nh 6 1 8 1 E RH – RH 24,380cm –1 4.841 10 –19 J ; 36 4 36 4.102 10 –5 cm = 410.2 nm 2.3 1 45 1 RH 25,190cm–1 5.002 10 –19 J E RH – 4 196 49 1 __ 3.970 10 –5 cm = 397.0 nm 2.4 hc (6.626 1034 J s)(2.998 108 m/s) 5.178 1019 J 9 (383.65 nm)(m/10 nm) 383.65 nm E 379.90 nm (6.626 10 34 J s)(2.998 108 m/s) E 5.229 10 19 J 9 (379.90 nm)(m/10 nm) hc Copyright © 2014 Pearson Education, Inc. 1 2 Chapter 2 Atomic Structure 1 1 1 E 2 1 1 E 1 and nh E RH 2 2 ; 2 4 RH 4 RH nh nh 2 1 1 5.178 10 19 J 2 9 For 383.65 nm: nh 18 4 2.1787 10 J 1 1 5.229 10 19 J 2 10 For 379.90 nm: nh 18 4 2.1787 10 J 2.5 The least energy would be for electrons falling from the n = 4 to the n = 3 level: 1 1 7 E RH 2 2 2.1787 10 18 J 1.059 10 19 J 3 4 144 The energy of the electromagnetic radiation emitted in this transition is too low for humans to see, in the infrared region of the spectrum. 2.6 E 102823.8530211 cm 1 97492.221701 cm 1 5331.6313201 cm 1 E hc (6.626 1034 J s)(2.998 108 m s)(100cm m)(5331.6313201 cm 1 ) E 1.059 1019 J This is the same difference found via the Balmer equation in Problem 2.5 for a transition from n 4 to n 3. The Balmer equation does work well for hydrogen. 2.7 a. We begin by symbolically determining the ratio of these Rydberg constants: RHe + RHe + RH 2 2 He + (Z He+ )2 e4 (4 0 )2 h2 He+ (Z He+ )2 H (Z H ) 2 4 He+ H RH 2 2 H (Z H )2 e4 (4 0 )2 h2 (since Z 2 for He ) The reduced masses are required (in terms of atomic mass units): 1 1 1 1 1 4 H me m proton 5.4857990946 10 m 1.007276466812 mu u 1 H 1823.88 mu1; H 5.482813 104 mu 1 He 1 He + 1 1 1 1 4 me mHe2 5.4857990946 10 m 4.001506179125 mu u 1823.13 mu1; He + 5.485047 10 4 mu Copyright © 2014 Pearson Education, Inc. Chapter 2 Atomic Structure 3 The ratio of Rydberg constants can now be calculated: RHe+ 4 He+ 4(5.485047 104 mu ) 4.0016 RH H 5.482813 104 mu b. RHe + 4.0016RH (4.0016)(2.1787 1018 J) 8.7184 1018 J c. The energy difference is first converted to Joules: E hc (6.626 1034 J s)(2.998 108 m s)(100cm m)(329179.76197 cm 1 ) E 6.5391 1018 J The Rydberg equation is applied, affording nearly the identical Rydberg constant for He+: 1 1 1 1 E RHe + ( 2 2 ) 6.5391 1018 J RHe + ( ) 1 4 n n l h RHe + 8.7188 1018 J 2.8 a. – h2 2 E ; 8 2 m x 2 A sin rx B cos sx = Ar cos rx – Bs sin sx x 2 = –Ar 2 sin rx – Bs 2 cos sx 2 x – h2 – Ar sin rx – Bs cos sx E A sin rx + B cos sx 8 m 2 2 2 If this is true, then the coefficients of the sine and cosine terms must be independently equal: h 2 Bs 2 h 2 Ar 2 EA ; = EB 8 2 m 8 2 m r 2 s2 8 2 mE 2 ; r s 2mE 2 h h b. A sin rx ; when x 0, A sin 0 0 when x a, A sin ra 0 n ra n ; r a c. E r 2 h 2 n 2 2 h 2 n 2h 2 8 2 m a 2 8 2 m 8ma 2 Copyright © 2014 Pearson Education, Inc. 4 Chapter 2 Atomic Structure d. a a x a dx A sin dx A a n 0 0 2 2 n 2 2 a nx nx sin 2 d a a 0 a a 2 1 nx 1 2nx A sin – 1 n 2 a 4 a 0 aA 2 1 na 1 1 – sin2n – 0 sin0 1 2 a 4 4 n A 2.9 2 a a. 3pz 4dxz b. 3pz 4dxz c. 3pz 4dxz For contour map, see Figure 2.8. Copyright © 2014 Pearson Education, Inc. Chapter 2 Atomic Structure 2.10 a. 4s 5dx2-y2 b. 4s 5dx2-y2 c. 4s 5dx2-y2 z x Dashed circles represent radial nodes. Electron density is inside the smallest node, between the nodes, and outside the largest node. Copyright © 2014 Pearson Education, Inc. 5 6 Chapter 2 Atomic Structure 2.11 a. 5s 4dz2 b. 5s 4dz2 c. 5s 4dz2 z x Dashed circles represent radial nodes. Electron density is inside the smallest node, between the nodes, and outside the largest node. Copyright © 2014 Pearson Education, Inc. Chapter 2 Atomic Structure 2.12 2.13 7 4 f z(x 2 y 2 )orbital a. no radial nodes (The number of radial nodes = n – l –1. Since n = 4 and l = 3 for f orbitals, n – l –1 = 4 – 3 – 1 = 0 radial nodes.) b. 3 angular nodes (The number of angular nodes = l = 3.) c. The angular nodes are solutions for z(x2–y2) = 0. These solutions are z = 0 (xy plane), and the planes where x = y and x = –y, both perpendicular to the xy plane. d. There are 8 lobes, 4 above and 4 below the xy plane. Down the z axis, this orbital looks like a dx2-y2 orbital, but the node at the xy plane splits each lobe in two. For an image of this orbital, please see http://winter.group.shef.ac.uk/orbitron/ or another atomic orbital site on the Web. A 5 f xyz orbital has the same general shape as the 4 f z(x 2 – y 2 ), with the addition of a radial node and rotation so the lobes are between the xy, xz, and yz planes. 2.14 a. 1 radial node (= n – l –1, as in problem 2.12) b. 3 angular nodes (= l) c. The angular nodes are solutions for xyz = 0. These solutions are the three planes where x = 0 (yz plane), y = 0 (xz plane), and z = 0 (xy plane). d. The diagram is similar to that in problem 2.12, with a radial node added and rotated by 45° around the z axis. There are 8 lobes, one in each octant of the coordinate system, both inside and outside the radial node, for a total of 16 lobes. For an image of this orbital, please see http://winter.group.shef.ac.uk/orbitron/ or another atomic orbital site on the Web. a. no radial nodes (= n – l –1, as in problem 2.12) b. 3 angular nodes (= l) c. The angular nodes are solutions for z(5z2 – 3r2) = 0 Solutions are z = 0 (the xy plane) and 5z2 – 3r2 = 0, or 5z2 = 3r2 Because r2 = x2 + y2 + z2 we can write 5z2 = 3(x2 + y2 + z2) 2z2 = 3x2 + 3y2 z2 = 3/2 (x2 + y2) This is the equation for a (double) cone. d. For an image of this orbital, please see http://winter.group.shef.ac.uk/orbitron/ or another atomic orbital site on the Web. Copyright © 2014 Pearson Education, Inc. 8 Chapter 2 Atomic Structure 2.15 a. 2.16 2.17 5d 4f 7g n 5 4 7 l 2 3 4 ml –2,–1,0,1,2 –3,–2,–1,0,1,2,3 –4,–3,–2,–1,0,1,2,3,4 b. n l ml 3 2 –2 3 2 –1 3 2 0 3 2 1 3 2 2 10 possible combinations c. For f orbitals (l = 3) possible values of ml are –3, –2, –1, 0, 1, 2, and 3. d. 2. (No more than 2 electrons can occupy any orbital!) a. For a 5d electron, l = 2 and n = 5. b. At most there can be ten 4d electrons, half of which can have ms = . c. f electrons have quantum number l = 3 and can have ml = –3, –2, –1, 0, 1, 2, or 3. d. For l = 4 (g electrons), ml can be –4, –3, –2, –1, 0, 1, 2, 3, or 4. a. The l quantum number limits the number of electrons. For l = 3, ml can have seven values (–3 to +3) each defining an atomic orbital, and for each value of ml the 3d ms ±2 ±2 ±2 ±2 ±2 1 2 value of ms can be 1 1 or ; there can be two electrons in each orbital. At most, 2 2 therefore, there can be 14 electrons with n = 5 and l = 3. 2.18 b. A 5d electron has l = 2, limiting the possible values of ml to –2, –1, 0, 1, and 2. c. p orbitals have l = 1 and occur for n 2. d. g orbitals have l = 4. There are 9 possible values of ml and therefore 9 orbitals. a. b. One exchange E = e W: The first configuration is favored. This configuration is stabilized by e (which is negative), and the second is destabilized by c (which is positive). and The first configuration is favored. It is 3 exchanges (1–2, 1–3, 2–3) E = 3e 2.19 One pair in same orbital E = c 1 exchange, 1 pair E = e + c X: 2 exchanges (1–2, 3–4) E = 2e stabilized by 2e more than the second configuration, and the second configuration is also destabilized by c. 3 exchanges (1–2, 1–3, 2–3) E = 3e, so this state has lower energy than W (because e is negative). Copyright © 2014 Pearson Education, Inc. Chapter 2 Atomic Structure 2.20 Y: Z: 6 exchanges (1–2, 1–3, 1–4, 2–3, 2–4, 3–4) E = 6e, so this state has lower energy than Y. (because e is negative). 2.21 Three states are possible: (three exchanges, E 3e ); 4 exchanges (1–2, 1–3, 2–3, 4–5); E = 4e 9 (one exchange, E e ); (one exchange and one pair, E e c ). Energy ranking: c Energy 2e 2.22 2.23 2.24 2.25 a. Figure 2.12 and the associated text explain this. Electron-electron repulsion is minimized by placing each electron in a separate orbital when the levels are close enough to allow it. At Cr, the second 4s electron has an energy higher than the lower five 3d electrons, and therefore the configuration is 4s1 3d5. b. Ti is 4s2 3d2, since both 4s electrons have energies below that of the 3d electrons at that Z. For ions, the 3d levels move down in energy, and are below the 4s levels for all transition metal M2+ ions, so Cr2+ is [Ar] 3d4. a. V 1s2 2s2 2p6 3s2 3p6 4s2 3d3 b. Br 1s2 2s2 2p6 3s2 3p6 4s2 3d10 4p5 c. Ru3+ 1s2 2s2 2p6 3s2 3p6 3d10 4s2 4p6 4d5 d. Hg2+ 1s2 2s2 2p6 3s2 3p6 4s2 3d10 4p6 5s2 4d10 5p6 4f14 5d10 e. Sb 1s2 2s2 2p6 3s2 3p6 4s2 3d10 4p6 5s2 4d10 5p3 a. Rb– [Kr] 5s2 b. Pt2– [Xe] 6s2 4f14 5d10 Electrons fill into atomic orbitals to maximize the nuclear attraction they experience. The more attraction an electron experiences, the more it is stabilized, and the more its energy is reduced (the more negative its energy). The order of orbital filling from n = 1 to n = 2 to n = 3 is rationalized by increasing the most probable distances these orbitals offer their electrons. Note that the most probable distances (the distance r where the radial probability is maximized) are roughly the same for all orbitals of the same n. As the most probable distance increases, the nuclear charge an electron experiences is attenuated to a greater extent by shielding by other electrons with shorter most probable distances. The 1s orbital fills first since its electrons have the most unhindered exposure to the charge of the nucleus; these electrons are subsequently Copyright © 2014 Pearson Education, Inc. 10 Chapter 2 Atomic Structure stabilized to the greatest extent. Electrons fill the n = 2 orbitals before the n = 3 orbitals due to the shorter most probable distance (and greater nuclear charge) offered by the former orbitals. Within a shell (orbitals with the same n), orbital energies are governed by the varying penetration abilities of orbitals with different l values. Electrons experience an increased nuclear charge (and are stabilized more) when residing in orbitals that penetrate towards the nucleus, affording electrons a reasonable probability of being found much closer to the nucleus than the most probable distance. Examination of the radial probability functions for n = 3 (Figure 2.7) reveals that a 3s electron has a higher probability of being found closest to the nucleus (inner maximum in graph at say 0.5r) relative to a 3p or 3d electron. A 3d electron has the lowest probability of being found at this distance from the nucleus; this orbital penetrates poorly and has no inner maximum. The order of filling within this shell reflects these relative penetration abilities; the 3s electrons are lower in energy than the 3p electrons, while the 3d electrons are the highest in energy of the electrons in the third shell. 2.26 2.27 2.28 a. When a fluorine atom gains an electron, it achieves the same electron configuration of the noble gas Ne, with all subshells filled. b. Zn2+ has the configuration [Ar] 3d10, with a filled 3d subshell. c. This configuration [Kr] 5s1 4d 5 has each electron outside the Kr core in a separate orbital and has the maximum number of electrons with parallel spins (6). In Mo, this configuration has a lower energy than [Kr] 5s2 4d 4. The greater stability of the s1d 5 configuration is also found in the case of Cr (directly above Mo), but W (below Mo) has an s2d 4 configuration. In Cr and Mo, the ns and (n-1)d subshells are sufficiently close in energy so that electron pairing is avoided. a. Ag+ has the configuration [Kr] 4d10, with a filled 4d subshell. b. In Cm, the s2d1f 7 configuration enables the last eight electrons each to occupy a separate orbital with parallel spin. This minimizes c in comparison with the alternative configuration, in which one f orbital would be occupied by an electron pair. c. The ion Sn2+ has the configuration [Kr] 5s2 4d10, with both the 5s and 4d subshells filled. This is an example of the “inert pair” effect, in which heavier elements of groups 13-16 often form compounds in which the oxidation states of these elements are 2 less than the final digit in the group number (other examples include Tl+ and Bi3+). The reasons for this phenomenon are much more complex than the ions’ electron configurations; see N. N. Greenwood and A. Earnshaw, Chemistry of the Elements, 2nd ed., Butterworth-Heinemann, London, 1997, pp. 226-227 for useful comments. a. Ti2+ has the configuration [Ar] 3d2; Ni2+ has the configuration [Ar] 3d8. In each case it should be noted that all the electrons outside the noble gas cores in these transition metal ions are d electrons, which occupy orbitals that are lower in energy than the 4s orbitals (see Figure 2.12(b) and the associated discussion). b. The preferred configuration of Mn2+ is [Ar] 3d5. As in the examples in part a, the 3d orbitals of Mn2+ are lower in energy than the 4s. In addition, the configuration minimizes electron-electron repulsions (because each d electron is in a separate orbital) and maximizes the stabilizing effect of electrons with parallel spins (maximum e). Copyright © 2014 Pearson Education, Inc. Chapter 2 Atomic Structure 2.29 (1s2) (2s2 2p6) (3s2 3pn) 1 0.85 0.35 15 – (2 × 1 + 8 × 0.85 + 4 × 0.35) = 16 – (2 × 1 + 8 × 0.85 + 5 × 0.35) = 17 – (2 × 1 + 8 × 0.85 + 6 × 0.35) = 18 – (2 × 1 + 8 × 0.85 + 7 × 0.35) = a. Z P S Cl Ar Z* r 4.8 5.45 6.1 6.75 106 pm 102 pm 99 pm 98 pm 11 The size of the atoms decreases slightly as Z increases, even though the number of electrons in the atom increases, because Z* increases and draws the electrons closer. Ar has the strongest attraction between the nucleus and the 3p electron, and the smallest radius. b. Z O2– – F Na+ Mg2+ (1s2) (2s2 2p6) 0.85 0.35 8 – (2 × 0.85 + 7 × 0.35) = 9 – (2 × 0.85 + 7 × 0.35) = 11 – (2 × 0.85 + 7 × 0.35) = 12 – (2 × 0.85 + 7 × 0.35) = Z* r 3.85 4.85 5.85 6.85 126 pm 119 pm 116 pm 86 pm These values increase directly with Z, and parallel the decrease in ionic size. Increasing nuclear charge results in decreasing size for these isoelectronic ions, although – the change between F and Na+ is smaller than might be expected. c. Cu (1s2) (2s2 2p6) (3s2 3p6) (3d10) (4s1) 4s S = 2 + 8 + (8×0.85) + (10×0.85) = 3d S = 2 + 8 + (8×1.00) + (9×0.35) = 25.3 21.15 Z* = 29 – 25.3 = 3.7 Z* = 29 – 21.15 = 7.85 The 3d electron has a much larger effective nuclear charge and is held more tightly; the 4s electron is therefore the first removed on ionization. d. (1s2) (2s2 2p6) (3s2 3p6) (3d10) (4s2 4p6) (4d10) (4fn) Ce S = 2 + 8 + 8 + 10 + 8 + 10 = 46 [Xe] 6s2 4f1 5d1 4f1 Z* = 58 – 46 = 12 S= 2 + 8 + [Xe] 6s2 4f3 8 + 10 + 8 + 10 + (2×0.35) = 46.7 4f3 Z* = 59 – 46.7 = 12.3 Nd S = 2 + 8 + [Xe] 6s2 4f4 8 + 10 + 8 + 10 + (3×0.35) = 47.05 4f4 Z* = 60 – 47.05 = 12.95 Pr The outermost electrons experience an increasing Z*, and are therefore drawn in to slightly closer distances with increasing Z and Z*. Copyright © 2014 Pearson Education, Inc. 12 Chapter 2 Atomic Structure 2.30 Sc: Ti: 3d electron: Electron (1s2)(2s2, 2p6)(3s2, 3p6)(3d1)(4s2) configuration (1s2)(2s2, 2p6)(3s2, 3p6)(3d2)(4s2) Contribution of other d electrons None [only one electron in (3d1)] One other 3d electron: Contribution to S = 1 × 0.35 = 0.35 Contribution of electrons to left of (3dn) 18 × 1.00 = 18.00 18 × 1.00 = 18.00 Total S Z* 18.00 21 – 18.00 = 3.00 18.35 22 – 18.35 = 3.65 4s electron: Electron (1s2)(2s2, 2p6)(3s2, 3p6)(3d1)(4s2) configuration (1s2)(2s2, 2p6)(3s2, 3p6)(3d2)(4s2) Contribution of other (4s2) electron Contribution to S = 0.35 Contribution to S = 0.35 Contribution of (3s2, 3p6)(3dn) electrons 9 × 0.85 = 7.65 10 × 0.85 = 8.50 Contribution of other electrons 10 × 1.00 = 10.00 10 × 1.00 = 10.00 Total S Z* 0.35 + 7.65 + 10.00 = 18.00 21 – 18.00 = 3.00 0.35 + 8.50 + 10.00 = 18.85 22 – 18.85 = 3.15 In scandium, Slater’s rules give the same value for the effective nuclear charge Z* for the 3d and 4s electrons, consistent with the very similar energies of these orbitals. In Ti the effective nuclear charge is slightly less for 4s than for 3d, by 0.50. This difference in the energies of 4s and 3d increases across the row of transition metals; by nickel, as calculated in the example in Section 2.2.4, the 4s orbital has an effective nuclear charge 3.50 less than the 3d orbitals. This is consistent with the experimental observation that transition metal cations have electron configurations in which there are no valence s electrons; the 3d orbitals are lower in energy than the 4s orbital, so d orbitals are the ones that are occupied. A similar phenomenon is observed for the second and third row transition metals. Copyright © 2014 Pearson Education, Inc. Chapter 2 Atomic Structure 2.31 13 N Z* = 7 – (2 × 0.85 + 4 × 0.35) = 3.9 IE = 1.402 MJ/mol r = 75 pm P Z* = 15 – (2 × 1 + 8 × 0.85 + 4 × 0.35) = 4.80 IE = 1.012 MJ/mol r = 106 pm As Z* = 33 – (2 × 1 + 8 × 1 + 18 × 0.85 + 4 × 0.35) = 6.3 IE = 0.947 MJ/mol r = 120 pm The effect of shielding alone is not sufficient to explain the changes in ionization energy. The other major factor is distance between the electron and the nucleus. Wulfsberg (Principles of Descriptive Inorganic Chemistry, Brooks/Cole, 1998) suggests that Z*/r2 correlates better, but As is still out of order: N P As 2.32 radius(pm) 75 106 120 Z*/r2 6.93 × 10–4 4.27 × 10–4 4.375 × 10–4 Zr through Pd have the electron configurations shown here: 40 41 42 43 44 45 46 Zr Nb Mo Tc Ru Rh Pd 5s2 4d2 5s1 4d4 5s1 4d5 5s2 4d5 5s1 4d7 5s1 4d8 4d10 The lower d line crosses the upper s line between elements 40 and 41, the upper d line crosses the upper s line between 43 and 44, and the upper d line crosses the lower s line between 45 and 46. This graph fits the experimental configurations well. 2.33 Li IE 5.39 Z* 1.30 r(pm, covalent) 123 Z*/r2 8.59 r(pm, ionic) 90 16.0 Z*/r2 > Na 5.14 2.20 154 9.28 116 16.3 > K 4.34 2.20 203 5.34 152 9.52 > Rb 4.18 2.20 216 4.72 (all × 10–5 pm–2) 166 7.98 (all × 10–5 pm–2) The Z*/r2 function explains the order, except for lithium, which seems to require more energy for removal of the electron than predicted. Apparently the very small size and small number of electrons on lithium result in the electron being held more tightly than in the other alkali metals. The Z*/r2 function also predicts larger differences between the IE values than are observed. Copyright © 2014 Pearson Education, Inc. 14 Chapter 2 Atomic Structure 2.34 C+ (1s2) (2s2 2p1) B (1s ) (2s 2p ) 2 2 1 Z* = 6 – (2 × 0.85 + 2 × 0.35) = 3.6 Z* = 5 – (2 × 0.85 + 2 × 0.35) = 2.6 The energies change by a factor of approximately three, but Z* changes only by 38%. Based on the data in Table 2.8 and the relative ionic and covalent sizes in problem 2.33, C+ has a radius of about 58 pm, B a radius of about 83 pm. Z*/r2 values are then 10.7 × 10–4 and 3.8 × 10–4, a ratio of 2.8. The ionization energies have a ratio of 2.99. This function of size and effective charge explains this pair quite well. 2.35 The maxima at 4 electrons correspond to the isoelectronic species Li , Be , and B+ ( 1s 2 2s2 ) changing to Li , Be + , and B2+ , respectively ( 1s 2 2s1 ). The energy requirement to remove electrons from species with completely filled subshells is generally relatively high. The minima at 5 electrons correspond to the isoelectronic species Be , B , and C+ ( 1s 2 2s2 2 p1) changing to Be , B+ , and C2+ , respectively ( 1s 2 2s2 ). The removal of an electron from a singly occupied subshell requires relatively less energy. These data support the stability of the 1s 2 2s2 electronic configuration. 2.36 a. He and H cannot form +3 cations; the first element under consideration is therefore Li. The third ionization energy of Be 2+ is expected to be a local maximum on the basis of removing an electron from a 1s 2 configuration. The third ionization energy of B2+ ( 1s 2 2s1 ) will be a local minimum on the basis of removing an electron from a singly occupied subshell. The third ionization energy of C2+ ( 1s 2 2s2 ) will be a local maximum on the basis of the removal of an electron from a filled subshell. The corresponding energies for N 2+ ( 1s 2 2s2 2 p1) and Ne 2+ ( 1s 2 2s 2 2 p 4 ) will be local minima on the graph on the basis of removal of an electron from a singly filled subshell, and the formation of a 2p subshell with three unpaired electrons (and no coulombic repulsion), respectively. Finally, there will be a maximum at Mg 2+ , which has a filled n = 2 subshell and a configuration matching that of neutral Ne. b. 2.37 The graph will share the common features of local maxima at 2 and 4 electrons, and local minima at 3, 5, and 8 electrons, as detailed above. The magnitudes of these third ionization energies are all greater than the corresponding second ionization energies for the same electronic configurations. Because the second ionization energies involve removing an electron from an ion with a single positive charge, the electron configurations of the cations must be considered: He+ Li+ Be+ B+ C+ 1s1 1s2 1s2 2s1 1s2 2s2 1s2 2s2 2p1 N+ O+ F+ Ne+ 1s2 2s2 2p2 1s2 2s2 2p3 1s2 2s2 2p4 1s2 2s2 2p5 Copyright © 2014 Pearson Education, Inc. Chapter 2 Atomic Structure 15 The peaks and valleys now match the peaks and valleys for electron configurations matching those of the atoms in the ionization energy graph in Figure 2.13. For example, because there is a minimum in the ionization energy for Li (electron configuration 1s2 2s1), there should also be a minimum in the graph of second ionization energy for Be+, which also has the configuration 1s2 2s1. The maximum in the ionization energy for Be (configuration 1s2 2s2) would be matched by a maximum in the second ionization graph for B+, which has the same configuration as Be. Overall, the maxima and minima in the two graphs are: Maximum or Minimum Configuration 1st IE 2nd IE (kJ/mol) Maximum Minimum Maximum Minimum 1s2 1s2 2s1 1s2 2s2 1s2 2s2 2p1 1s2 2s2 2p2 1s2 2s2 2p3 1s2 2s2 2p4 1s2 2s2 2p5 He Li Be B C N O F Li+ Be+ B+ C+ N+ O+ F+ Ne+ Maximum Minimum 2.38 2.39 2.40 (7298) (1757) (2427) (2353) (2856) (3388) (3374) (3952) a. Fe (7.87 eV) > Ru (7.37 eV) They have the same Z* (6.25), but Ru is larger, so Z*/r2 is smaller than for Fe. b. P (10.486 eV) > S (10.36 eV) Z* is smaller for P (4.8) than for S (5.45), but S has one electron paired in the 3p level, which increases its energy and makes it easier to remove. c. Br (11.814 eV) > K (4.341 eV) Z* for K is 2.2; for Br it is 7.6, a very large difference. K is also nearly twice as large as Br. Both factors contribute to the difference in IE. d. N (14.534) > C (11.260) Increasing Z* (3.9 for N, 3.25 for C) and decreasing size (75 pm for N, 77 pm for C) both work in the same direction. e. Cd (8.993 > In (5.786) Indium starts a new 5p subshell, so the last electron is easily removed in spite of a larger Z* (2.30 for In, 1.65 for Cd). f. The smaller F has a higher ionization energy (17.422) than Cl (12.967). An electron is more easily removed from the 3p subshell of Cl than from the smaller 2p subshell of F. a. S (2.077 eV) has a smaller EA than Cl (3.617 eV) because Cl has a larger Z* (6.1 vs. 5.45) and a slightly smaller radius. Both increase the attractive power for an electron. b. I (3.059) has a smaller EA than Br (3.365) because it is larger, with the same Z* (7.6). c. B (8.298) has a smaller IE than Be (9.322) because it is starting a new p subshell. d. S (10.360) has a smaller IE than P (10.486) because S is losing one of a pair of p electrons. a. The maximum at Mg comes at a completed subshell (3s2) and the minimum at Al is at the first electron of a new subshell (3p), increasing the energy and making removal of an electron easier. The maximum at P is at a half-filled subshell (3p3) and the minimum Copyright © 2014 Pearson Education, Inc. 16 Chapter 2 Atomic Structure at S is at the fourth 3p electron, which must pair with one of the others; this also raises the energy of the electron and makes its removal easier. b. The reasons for the minima and maxima in the electron affinity graph are the same as in the ionization energy graph. The maxima and minima are shifted by one in the two graphs because the reactants in the process defining electron affinity have a negative charge, one more electron than a neutral atom. For example, minima occur for ionization energy at Al ([Ne]3s23p1 and for electron affinity at Mg (([Ne]3s23p1 Al and – Mg have identical electron configurations. 2.41 The Bohr equation predicts that the energy levels of 1-electron systems should be proportional to Z 2. Z 2 = 4 for He+ and 9 for Li2+, so ratios of the ionization energies to that of H are 4:1 and 9:1. 2.42 In both the transition metals and the lanthanides (problem 2.29.d), the gradual change in Z* pulls the outer electrons closer. The increase in Z* is 0.65 for each unit increase in atomic number, so the increase in attraction is relatively small, and the change in radius must also be small. 2.43 a. Se2– > Br > Rb+ > Sr2+ These ions are isoelectronic, so the sizes are directly dependent on the nuclear charges. For example, Sr2+ has the greatest nuclear charge and is the smallest ion. b. Y3+ > Zr4+ > Nb5+ These ions are also isoelectronic, so the increasing nuclear charge results in decreasing size. c. Co4+ < Co3+ < Co2+ < Co The smaller number of electrons with constant nuclear charge results in a smaller size. a. F– has the largest radius. All three choices in this isoelectronic set have 10 electrons, and F– has the fewest protons (9) to attract these electrons. b. Te2– has the largest volume. In general, atomic and ionic sizes increase going down a column of main group elements as the number of shells increases, with electrons closer to the nucleus shielding outer electrons from the full effect of the nuclear charge. c. Mg, with a filled 3s subshell, has the highest ionization energy. The situation is similar to that in the second period, where Be has a higher ionization energy than Li and B. d. Fe3+ has the greatest surplus of positive charge (26 protons, 23 neutrons), making it the smallest and the most difficult from which to remove an electron. e. Of the anions O–, F–, and Ne–, F– has the configuration of the noble gas Ne, from which it is more difficult to remove an electron than from the other ions; F, therefore, has the highest electron affinity. 2.44 – Copyright © 2014 Pearson Education, Inc. Chapter 2 Atomic Structure 2.45 2.46 2.47 17 a. V is the smallest; the effective nuclear charge on the outer electrons increases across the row of transition metals. b. These species are isoelectronic; each has 18 electrons. S2–, which has the fewest protons, is therefore the largest. c. The ionization energy is smallest for the largest alkali metals (weaker attraction between the nucleus and outermost electrons), so Cs has the lowest value. d. The electron affinity is the greatest for the smallest halogens (stronger attraction between nucleus and outermost electrons), so Cl has the highest value. e. The Cu2+ ion has the greatest surplus of protons and is therefore the smallest and the most difficult from which to remove an electron. a. 7 orbitals b. See images. c. The number of radial nodes increases as n increases: 4f orbitals have no radial nodes, 5f have one radial node, and so forth. a. 9 orbitals b. See images. c. As in problem 2.46.c, with no radial nodes for 5g, one for 6g, etc. Copyright © 2014 Pearson Education, Inc. 18 Chapter 3 Simple Bonding Theory CHAPTER 3: SIMPLE BONDING THEORY 3.1 a. Structures a and b are more likely than c, because the negative formal charge is on the electronegative S. In c, the electronegative N has a positive charge. S CH3 C 1– S CH3 C CH3 S 1– N S b 1+ C CH3 N S 1– CH3 CH3 c – b. a. S N a 3.2 1– The same structures fit (OSCN(CH3)2 . The structure with a 1– formal charge on O is most likely, since O is the most electronegative atom in the ion. 2– 1+ Se The formal charges are large, but match electronegativity. C N C N Negative formal charge of 1– on Se, a low electronegativity atom. 1– Se 1– Se b. the C N Negative formal charge on N, the most electronegative atom. BBest resonance structure of the three. b is better than a, because the formal charge is on more electronegative O. O H a and b are better than c, because one of the formal charges is on the more electronegative O. 1– S H C a c. O 1– S b 1– 1– O S – SNO : a has a 1– formal charge on S. Not very likely, doesn’t match electronegativity (negative formal charge is not on most electronegative atoms). b has 1– formal charge on O, and is a better structure. – 1– 1– S 1+ 1+ 2– N S O 1– S S 1– O N a 1– 1– S c b N N O S a Overall, the S=N–O structure is better based on formal charges, since it has only a negative charge on O, the most electronegative atom in the ion. Copyright © 2014 Pearson Education, Inc. C S b – NSO : a has 2– formal charge on N, 1+ on S. Large formal charges, not very likely. b has 1– formal charges on N and O, 1+ on S, and is a better structure. O C S a 3.3 S 1– O C C b 1– O Chapter 3 Simple Bonding Theory I 3.4 1+ O A N O B N 1+ 1+ O C N II 1– C N C N 1– 1– 1+ C N O III 1+ O N C 1– 1+ N 1– 2– 1+ N O 19 N C N O 1– 1+ N O C N 2+ 1– 1– O C N 2– 1+ N N C 1– 2+ N C N Structure IB is best by the formal charge criterion, with no formal charges, and is expected to be the most stable. None of the structures II or III are as good; they have unlikely charges (by electronegativity arguments) or large charges. 3.5 N 1+ N 1– 1– O N 2– 1+ N O N 1+ N 1+ O The first resonance structure, which places the negative formal charge on the most electronegative atom, provides a slightly better representation than the second structure, which has its negative formal charge on the slightly less electronegative nitrogen. Experimental measurements show that the nitrogen–nitrogen distance (112.6 pm) in N2O is slightly closer to the triple bond distance (109.8 pm) in N2 than to the double bond distances found in other nitrogen compounds, and thermochemical data are also consistent with the first structure providing the best representation. The third resonance structure, with greater overall magnitudes of formal charges, is the poorest representation. O 3.6 H O O 1+ H N O 1– 1+ N O 1– O 3.7 Molecule, Including Usual Formal Charges 1– C 1+ O 1– N O H F Atom Group Number Unshared Electrons C 4 2 O 6 2 N 5 4 O 6 4 H 1 0 F 7 6 A 2 A B 2.544 2 0.83 2.544 3.61 3.61 2 1.17 2.544 3.61 3.066 2 0.92 3.066 3.61 3.61 2 1.08 3.066 3.61 2.300 2 0.71 2.300 4.193 4.193 2 1.29 2.300 4.193 – Number of Bonds Calculated Formal Charge 3 –0.49 3 0.49 2 –0.84 2 –0.16 1 0.29 1 –0.29 Surprisingly, CO is more polar than FH, and NO is intermediate, with C and N the negative – atoms in CO and NO . Copyright © 2014 Pearson Education, Inc. 20 Chapter 3 Simple Bonding Theory 3.8 a. Cl SeCl4 requires 10 electrons around Se. The lone pair of electrons in an equatorial position of a trigonal bipyramid distorts the shape by bending the axial chlorines back. Cl Se Cl Cl b. – I3 requires 10 electrons around the central I and is linear. S c. PSCl3 is nearly tetrahedral. The multiple bonding in the P–S bond compresses Cl the Cl—P—Cl angles to 101.8°, Cl significantly less than the tetrahedral angle. P Cl – d. IF4 has 12 electrons around I and has a square planar shape. e. PH2 has a bent structure, with two lone pairs. – 2– F F f. TeF42– has 12 electrons around Te, with a square planar shape. g. N3 is linear, with two double bonds in its best resonance structure. N SeOCl4 has a distorted trigonal bipyramidal shape with the extra repulsion of the double bond placing oxygen in an equatorial Cl position. Cl h. + H 3.9 PH4 is tetrahedral. Cl Se Cl + H H P H – a. ICl2 has 10 electrons around I and is linear. b. H3PO3 has a distorted tetrahedral shape. c. F – N i. F Te – H – BH4 is tetrahedral. H H B H Copyright © 2014 Pearson Education, Inc. N O Chapter 3 Simple Bonding Theory 21 O d. e. . f. POCl3 is a distorted tetrahedron. The Cl—P—Cl angle is compressed to 103.3° as a result of the P—O double bond. P Cl Cl Cl O – IO4 is tetrahedral, with significant double bonding; all bonds are equivalent. I O O O IO(OH)5 has the oxygens arranged octahedrally, with hydrogens on five of the six oxygens. HO SOCl2 is trigonal pyramidal, with one lone pair and some double bond character in the S–O bond. h. ClOF4 is a square pyramid. The double bonded O and the lone pair occupy opposite positions. i. The F—Xe—F angle is nearly linear (174.7°), with the two oxygens and a lone pair in a trigonal planar configuration. Formal charges favor double bond character in the Xe–O bonds. The O—Xe—O angle is narrowed to 105.7° by lp-bp repulsion. OH O H S Cl Cl OH I HO O g. O – F O Xe O F 3.10 a. SOF6 is nearly octahedral around the S. O F F F IS F F O F b. of POF3 has a distorted tetrahedral shape, with F—P—F angles 101.3°. c. ClO2 is an odd electron molecule, with a bent shape, partial double bond character, and an angle of 117.5°. d. F F P F Cl OI I O N NO2 is another odd electron molecule, bent, with partial double bond I I O character and an angle of 134.25°. This is larger than the angle of ClO2 O because there is only one odd electron on N, rather than the one pair and single electron of ClO2. Copyright © 2014 Pearson Education, Inc. 22 Chapter 3 Simple Bonding Theory S e. S2O42– has SO2 units with an angle of about 30° between their planes, in an eclipsed conformation. O 2– S O O O H f. N2H4 has a trigonal pyramidal shape at each N, and a gauche conformation. There is one lone pair on each N. N N H H g. h. i. ClOF2+ is a distorted trigonal pyramid with one lone pair and double bond character in the Cl—O bond. CS2, like CO2, is linear with double bonds. S C F S + H Cl O F – O – The structure of XeOF5 is based on a pentagonal bipyramid, with a lone pair and the oxygen atom in axial positions. See K. O. Christe et al., Inorg. Chem., 1995, 34, 1868 for evidence in support of this structure. F F F Xe F F 3.11 All the halate ions are trigonal pyramids; as the central atom increases in size, the bonding pairs are farther from the center, and the lone pair forces a smaller angle. The decreasing electronegativity Cl > Br > I of the central atom also allows the electrons to be pulled farther out, reducing the bp-bp repulsion. 3.12 a. AsH3 should have the smallest angle, since it has the largest central atom. This minimizes the bond pair—bond pair repulsions and allows a smaller angle. Arsenic is also the least electronegative central atom, allowing the electrons to be drawn out farther and lowering the repulsions further. Actual angles: AsH3 = 91.8°, PH3 = 93.8°, NH3 = 106.6°. b. Cl is larger than F, and F is more electronegative and should pull the electrons farther from the S, so the F—S—F angle should be smaller in OSF2. This is consistent with the experimental data: the F—S—F angle in OSF2 is 92.3° and the Cl—S—Cl angle in OSCl2 is 96.2°. – c. NO2 has rather variable angles (115° and 132°) in different salts. The sodium salt (115.4°) has a slightly smaller angle than O3 (116.8°). The N–O electronegativity difference should pull electrons away from N, reducing the bp-bp repulsion and the angle. 3.13 – O O O O O – O N N O O – d. BrO3 (104°) has a slightly smaller angle than ClO3 (107°), since it has a larger central atom. In addition, the greater electronegativity of Cl holds the electrons closer and increases bp-bp repulsion. a. N3 is linear, with two double bonds. O3 is bent (see solution to 3.12.c), with one double bond and a lone pair on the central O caused by the extra pair of electrons. – Copyright © 2014 Pearson Education, Inc. O O – Chapter 3 Simple Bonding Theory b. 23 Adding an electron to O3 decreases the angle, as the odd electron spends part of its time on the central O, making two positions for electron repulsion. The decrease in angle is small, however, with angles of 113.0 to 114.6 pm reported for alkali metal ozonides (see W. Klein, K. Armbruster, M. Jansen, Chem. Commun., 1998, 707) in comparison with 116.8° for ozone. O 3.14 Cl O Cl O CH3 H3Si H3C 110.9° 111.8° SiH3 144.1° As the groups attached to oxygen become less electronegative, the oxygen atom is better able to attract shared electrons to itself, increasing the bp-bp repulsions and increasing the bond angle. In the case of O(SiH3)2, the very large increase in bond angle over O(CH3)2 suggests that the size of the SiH3 group also has a significant effect on the bond angle. 3.15 C3O2 has the linear structure O=C=C=C=O, with zero formal charges. + 1+ N5+ with the same electronic structure has formal charges of 1–, 1+, 1+, 1+, 1–, unlikely because three positive charges are adjacent to each other. Changing to N=N=N–NN results in formal charges of 1–, 1+, 0, 1+, 0, a more reasonable result with an approximately trigonal angle in the middle. With triple bonds on each end, the formal charges are 0, 1+, 1–, 1+, 0 and a tetrahedral angle. Some contribution from this would reduce the bond angle. 1– N N N 1+ N 1– 1+ N N N + 1+ N OCNCO+ can have the structure OC–N–CO, with formal charges of 1+, 0, 1–, 0, 1+ and two lone pairs on the central N. N This would result in an even smaller angle in the middle, but has C C positive formal charges on O, the most electronegative atom. O = = O C N–CO has a formal charge of 1+ on the final O. Resonance would reduce that formal charge, making this structure 1– and a trigonal angle more likely. The Seppelt reference also N mentions two lone pairs on N and cites “the markedly C C higher electronegativity of the nitrogen atom with respect to the 1+ O central atom in C3O2, which leads to a higher localization of electron density in the sense of a nonbonding electron pair.” Therefore, the bond angles should be OCCCO > OCNCO+ > N5+. Literature values are 180°, 130.7°, and 108.3 to 112.3° (calculated), respectively. 3.16 a. H H C H H N C H H N H H In ethylene, carbon has p orbitals not involved in sigma bonding. These orbitals interact to form a pi bond between the carbons, resulting in planar geometry. (Sigma and pi Copyright © 2014 Pearson Education, Inc. N N + O 1+ + O 1+ 24 Chapter 3 Simple Bonding Theory bonding are discussed further in Chapter 5.) In hydrazine each nitrogen has a steric number of 4, and there is sigma bonding only; the steric number of 4 requires a threedimensional structure. b. – In ICl2 the iodine has a steric number of 5, with three lone pairs in equatorial positions; – the consequence is a linear structure, with Cl atoms occupying axial positions. In NH2 the two lone pairs require a bent arrangement. 3.17 3.18 3.19 c. Resonance structures of cyanate and fulminate are shown in Figures 3.4 and 3.5. The fulminate ion has no resonance structures that have as low formal charges as structures A and B shown for cyanate. The guideline that resonance structures having low formal charges tend to correspond to relatively stable structures is followed here. Hg(CNO)2, which has higher formal charges in its resonance structures, is the explosive compound. a. PCl5 has 10 electrons around P, using 3d orbitals in addition to the usual 3s and 3p. N is too small to allow this structure. In addition, N would require use of the 3s, 3p, or 3d orbitals, but they are too high in energy to be used effectively. b. Similar arguments apply, with O too small and lacking in accessible orbitals beyond the 2s and 2p. a. The lone pairs in both molecules are equatorial, the position that minimizes 90° interactions between lone pairs and bonding pairs. b. In BrOF3 the less electronegative central atom allows electrons in the bonds to be pulled toward the F and O atoms to a greater extent, reducing repulsions near the central atom and enabling a smaller bond angle. In BrOF3 the Feq–Br–O angle is approximately 4.5° smaller than the comparable angle in ClOF3. F F X O F a. The CH 3 —N— CH 3 angle is expected to be larger than the CH 3 —P— CH 3 angle; bp-bp repulsion will be more intense at the N due to the higher electronegativity of N relative to P. The angles are 108.2° ( CH 3 —N— CH 3 ) and 103.4° ( CH 3 —P— CH 3 ). b. N(CH 3 )3 is expected to exert a greater steric influence on Al(CH 3 )3 relative to P(CH 3 )3 on the basis of a shorter Al—N bond distance (204.5 pm) than Al—P bond distance (253 pm). Therefore, (CH 3 )3NAl(CH 3 )3 has a more acute CH 3 — Al—CH 3 angle (114.4°) than (CH 3 )3PAl(CH 3 )3 (117.1°). c. On the basis of the steric argument applied in part b, (CH 3 )3NAl(CH 3 )3 should have a longer Al—C distance. However, while this distance is slightly longer in Copyright © 2014 Pearson Education, Inc. Chapter 3 Simple Bonding Theory 25 (CH 3 )3NAl(CH 3 )3 relative to (CH 3 )3PAl(CH 3 )3 (1.978 pm vs. 1.973 pm), these lengths are not statistically different when their standard deviations are considered. Data for (CH 3 )3NAl(CH 3 )3 from T. Gelbrich, J. Sieler, U. Dümichen, Z. Kristallogr., 2000, 215, 127. Data for (CH 3 )3PAl(CH 3 )3 from A. Almenningen, L. Fernholt, A. Haaland, J. Weidlein, J. Organomet. Chem., 1978, 145, 109. 2– 3.20 IF32– has three lone pairs and three bonds. Overall, this ion is predicted to be T-shaped, with bond angles slightly less than 90°. 3.21 a. There are three possibilities: b. The third structure, with the lone pair and double bonds in a facial arrangement, is least likely because it would have the greatest degree of electron-electron repulsions involving these regions of high electron concentrations. F I F F The second structure, which has fewer 90° lone pair–double bond repulsions than the first structure, is expected to be the most likely. Experimental data are most consistent with this structure. 3.22 – c. One possibility: XeO2F3 a. Three unique arrangements of the nonbonding pairs and the oxygen atom are possible in XeOF3– when the octahedral electron-group geometry is considered: (1) trans nonbonding pairs; (2) one pair trans and one cis the oxygen atom; and (3) both pairs cis the oxygen atom. F F F Xe 1 O F F Xe 2 F O F F F Xe O 3 A square-planar structure (1) is expected to minimize lp-lp repulsions relative to structures 2 and 3 with unfavorable 90° lp-lp interactions. A low temperature Raman spectroscopic study coupled with quantum-chemical calculations of shock-sensitive salts of XeOF3– confirms this prediction (D. S. Brock, H. P. A. Mercier, G. J. Schrobilgen, J. Am. Chem. Soc., 2010, 132, 10935). Copyright © 2014 Pearson Education, Inc. 26 Chapter 3 Simple Bonding Theory b. 3.23 3.24 3.25 This anion is notable as the first example of a VSEPR arrangement ( AX 3YE 2 ) that features a doubly bond atom (oxygen) positioned approximately 90° to relative the domains of two nonbonding pairs. I(CF3)Cl2 is roughly T-shaped, with the two Cl atoms opposite each other and the CF3 group and two lone pairs in the trigonal plane. The experimental Cl—I—C angles are 88.7° and 82.9°, smaller than the 90° expected if there were no extra repulsion from the lone pairs. Repulsion between the lone pairs and the larger CF3 group put them in the trigonal plane, where there is more room. Cl I CF3 Cl a. CF3 has a greater attraction for electrons than CH3, so the P in PF2(CF3)3 is more positive than the P in PF2(CH3)3. This draws the F atoms in slightly, so the P—F bonds are shorter in PF2(CF3)3 (160.1 pm vs. 168.5 pm). b. Al—O—Al could have an angle near 109°, like water, or could have double bonds in both directions and a nearly linear structure. In fact, the angle is about 140°. The single-bonded picture is more probable; the high O electronegativity of O compared to Al draws the bonding pairs I Al AlI closer, opening up the bond angle. A Lewis structure with zero formal charges on all atoms can be drawn for this molecule with four electrons on each Al. Al c. CAl4 is tetrahedral. Again, a Lewis structure with zero formal charges can be drawn with four electrons on each Al. a. Al Al C Al The Te—X(axial) distances are expected to be longer than the Te—X(equatorial) distances on the basis of the increased lp-bp and bp-bp repulsion that the electron groups in the axial positions experience relative to those in the equatorial positions. The observed bond distances exhibit these features for both TeF4 (Te—F(axial), 189.9 pm; Te—X(equatorial), 184.6 pm) and TeCl4 (Te—Cl(axial), 242.8 pm; Te—Cl(equatorial), 228.9 pm). b. These angles should both be smaller in TeF4 , on the basis of reduced bp-bp repulsion at the Te atom in TeF4 due to the higher electronegativity of F relative to Cl. The angles were determined as 164.3° (F(axial)—Te—F(axial)), 176.7° (Cl(axial)—Te—Cl(axial)), 99.5° (F(equatorial)—Te—F(equatorial)), and 102.5° (Cl(equatorial)—Te— Cl(equatorial). The equatorial nonbonding pair in these complexes has a greater influence in TeF4 than in TeCl4 . 3.26 Octahedral SeCl62– TeCl62– ClF6– Distorted SeF62– IF6– The distorted structures have the smallest outer atoms in comparison with the size of the central atom. In these cases, there apparently is room for a lone pair to occupy a position that can lead to distortion. In the octahedral cases there may be too much crowding to allow a lone pair to distort the shape. Copyright © 2014 Pearson Education, Inc. Chapter 3 Simple Bonding Theory F F 3.27 O 27 Xe N CCH3 Xe F O F In F2OXeNCCH3, the nitrogen–xenon bond is weak; see the reference for details on bond distances and angles. 3.28 a. O is more electronegative than N and can draw the electrons more strongly away from the S. The more positive S in OSCl2 consequently attracts bonding pairs in S–Cl bonds closer to sulfur, increasing bp-bp repulsions and increasing the Cl—S—Cl angle (96.2° in OSCl2, 93.3° in NSCl2–). b. Because the sulfur in OSCl2 attracts the S–Cl bonding pairs more strongly, these bonds – are shorter: 207.6 pm in OSCl2, 242.3 pm in NSCl2 . Cl 3.29 The larger, less electronegative Br atoms are equatorial. Br Br P Cl Cl 3.30 3.31 a. In PCl3(CF3)2, the highly electronegative CF3 groups occupy axial positions. b. The axial positions in SbCl5 experience greater repulsions by bonding pairs, leading to longer Sb–Cl (axial) bonds (223.8 pm) than Sb–Cl (equatorial) bonds (227.7 pm). The pertinent group electronegativity ranking is CF3 > CCl3 > CH 3 . Therefore, ClSO2CF3 is expected to possess the lowest concentration of electron density near the S of the S—C bond, and ClSO2CH 3 the highest concentration of electron density. The bp-bp repulsion that influences the Cl—S—C angles should decrease as ClSO 2CH 3 > ClSO 2CCl3 > ClSO2CF3. Therefore, ClSO2CF3 should exhibit the smallest Cl—S—C angle, and ClSO2CH 3 the largest Cl—S—C angle. The angles measured in the gas phase for these molecules are 101° ( ClSO2CH 3 ; M. Hargittai, I. Hargittai, J. Chem. Phys., 1973, 59, 2513) , 96° ( ClSO 2CCl3 ; N. V. Alekseev, Z. Struki. Khimii, 1967, 8, 532), and 95.4° ( ClSO2CF3; R. Haist, F. Trautner, J. Mohtasham, R. Winter, G. L. Gard, H. Oberhammer, J. Mol. Struc., 2000, 550, 59). 3.32 The FSO 2 X molecule with the smallest O—S—O angle is expected to be that with the greatest concentration of electron density at the S atom from the S—X bond. This molecule should exert the greatest amount of bp-bp repulsion between the S—F and S—X bonds, maximally hindering expansion of the O—S—O angle within this series. The pertinent group electronegativity ranking is F > OCH 3 CH 3; the S— CH 3 bond should possess the greatest electron density at the S Copyright © 2014 Pearson Education, Inc. 28 Chapter 3 Simple Bonding Theory atom. While FSO2CH 3 exhibits a smaller O—S—O angle (123.1°, I. Hargittai, M. Hargittai, J. Mol. Struc., 1973, 15, 399) than found in FSO2 (OCH 3 ) (124.4°, I. Hargittai, R. Seip, K. P. Rajappan Nair, C. O. Britt, J. E. Boggs, B. N. Cyvin, J. Mol. Struc., 1977, 39, 1.), the O—S—O angle of FSO2 F (123.6°, D. R. Lide, D. E. Mann, R. M. Fristrom, J. Chem. Phys., 1957, 26, 734) is smaller than expected on the exclusive basis of group electronegativity arguments. It is noteworthy that O—S—O angles ranging from 122.6 to 130°, with rather large standard deviations (see K. Hagen, V. R. Cross, K. Hedberg, J. Mol. Struc., 1978, 44, 187), have also been reported for FSO2 F . 3.33 a. Because Te is less electronegative than Se, the highly electronegative C5F4 N groups draw electron density away from the Te atom more effectively than from the Se atom, rendering more effective lp-bp repulsion in compressing the C—group 16 atom—C angle in Te(C5F4 N)2 . b. As the electronegativity of the group bound to these atoms increases, lp-bp repulsion is expected to have increasing impact in compressing the C—group 16 atom—C angle. The pertinent group electronegativity ranking is C5F4 N > C6 F5 on the basis of theoretical calculations ( B. Hoge, C. Thösen, T. Hermann, P. Panne, I. Patenburg, J. Flourine Chem., 2004, 125, 831). 3.34 PF4+ has the bond angle expected for a tetrahedron, 109.5°. In PF3O the multiple bond to oxygen results in distortion away from the oxygen, leading to a smaller F–P–F angle. By the LCP approach the F…F distances should be approximately the same in these two structures. They are similar: 238 pm in PF4+ and 236 pm in PF3O. 3.35 As more (less electronegative) CH3 groups are added, there is greater concentration of electrons near P, and greater electron-electron repulsion leads to longer axial P–F bonds. Reported P–F distances: 161 3.36 PF4(CH3) pm PF3(CH3)2 164 pm PF2(CH3)3 168 pm Bond angles and distances: F2C=CF2 F2CO CF4 – F3CO Steric Number 3 3 4 4 C—F (pm) 133.6 131.9 131.9 139.2 FCF angle (°) 109.2 107.6 109.5 101.3 F—F (pm) 218 216 216 215 The differences between these molecules are subtle. The LCP model views the F ligands as hard objects, tightly packed around the central C in these examples. In this approach, the F…F distance remains nearly constant while the central atom moves to minimize repulsions. 3.37 The calculation is similar to the example shown in Section 3.2.4. Copyright © 2014 Pearson Education, Inc. Chapter 3 Simple Bonding Theory Cl O 55.9° C 289 pm x C C Cl Cl Cl Cl 111.8° Cl Cl 144.5 pm 289 pm sin 55.9° = 0.828 = x = 174.5 pm 3.38 29 144.5 x CF3+ is expected to be trigonal planar with a 120° F—C—F angle. F 60o C C x F 120o F 215 pm F 107.5 pm sin 60 0.866 x 124.1 pm 107.5 x The C—F bond length predicted via quantum chemical calculations is 124.4 pm, with a F…F distance of 216 pm (R. J. Gillespie and P. L. A. Popelier, Chemical Bonding and Molecular Geometry, Oxford, New York, 2001, p. 119). 3.39 By the LCP approach, from the structures of HOH and FOF, the hydrogen radius would be 76 pm (half of the H…H distance) and the fluorine radius (half of the F…F distance) would be 110 pm. Because the LCP model describes nonbonded outer atoms as being separated by the sums of their radii, as if they were touching spheres, the H…F distance in HOF would therefore be the sum of the ligand radii, 76 + 110 = 186 pm, in comparison with the actual H…F distance of 183 pm. If the covalent O–H and O–F bonds in HOF are similar to the matching distances in HOH and FOF, the H–O–F angle must be smaller than the other angles because of the H…F distance. O An alternative explanation considers the polarity in HOF. Because of the high + electronegativity of fluorine, the F atom in HOF acquires a partial negative – H charge, which is attracted to the relatively positive H atom. By this approach, F electrostatic attraction between H and F reduces the bond angle in HOF, giving it the smallest angle of the three compounds. 3.40 The electronegativity differences are given in parentheses: a. C–N N is negative (0.522) d. O–Cl O is negative (0.741) b. N–O O is negative (0.544) e. P–Br Br is negative (0.432) c. C–I C is negative (0.185) f. S–Cl Cl is negative (0.280) The overall order of polarity is O–Cl > N–O > C–N > P–Br > S–Cl > C–I. Copyright © 2014 Pearson Education, Inc. 30 Chapter 3 Simple Bonding Theory 3.41 a. b. O VOCl3 has a distorted tetrahedral shape, with Cl—V—Cl angles of 111°, and Cl—V—O angles of 108°. V Cl Cl Cl PCl3 has a trigonal pyramidal shape with Cl—P—Cl angles of 100.4°. F c. SOF4 has a distorted trigonal bipyramidal shape. The axial fluorine atoms are nearly linear with the S atom; the equatorial F—S—F angle is 100°. F S F P Cl Cl Cl O F O d. SO3 is trigonal (triangular), with equal bond angles of 120°. S O e. O ICl3 would be expected to have two axial lone pairs, causing distortion to reduce the Cl (axial)–I–Cl (equatorial) angles to < 90°. However, reaction of I2 with Cl2 yields dimeric I2Cl6, which readily dissociates into ICl and Cl2. Cl I F f. SF6 is a regular octahedron. F F S Cl F F F F g. h. F IF7 is a rare example of pentagonal bipyramidal geometry. F F O F F Xe O CF2Cl2, like methane, is tetrahedral. O P4O6 is described in the problem. Each P has one lone pair, each O has two. I F The structure of XeO2F4 is based on an octahedron, with oxygens in trans positions because of multiple bonding. C Cl Cl j. F F F i. Cl P O O P O F P O Copyright © 2014 Pearson Education, Inc. P O F F Chapter 3 Simple Bonding Theory 3.42 a. PH3 has a smaller bond angle than NH3, about 93°. The larger central atom reduces the repulsion between the bonding pairs. b. H2Se has a structure like water, with a bond angle near 90°. The larger central atom increases the distance between the S–H bonding pairs and reduces their repulsion, resulting in a smaller angle than in water. P H H H H Se H F c. SeF4 has a lone pair at one of the equatorial positions of a trigonal bipyramid, and bond angles of about 110° (equatorial) and 169° (axial). Seesa w shape. F F d. PF5 has a trigonal bipyramidal structure. e. IF5 is square pyramidal, with slight distortion away from the lone pair. f. F XeO3 has a trigonal pyramidal shape, similar to NH3, but with Xe–O double bonds. Cl g. h. i. BF2Cl is trigonal planar, with FBCl larger than FBF. P 31 F Se F F F F I F F F F F O O Xe O B F F SnCl2 has a bond angle of 95° in the vapor phase, smaller than the trigonal angle. As a solid, it forms polymeric chains with bridging chlorines and bond angles near 80°. Sn Cl Cl KrF2 is linear: F—Kr—F . VSEPR predicts three lone pairs on krypton in equatorial positions, with the fluorine atoms in axial positions. O j. IO2F52– has a steric number of 7 on iodine, with oxygen atoms occupying axial positions. F F F F O 3.43 Polar: VOCl3, PCl3, SOF4, ICl3, CF2Cl2 3.44 Polar: PH3, H2Se, SeF4, IF5, XeO3, BF2Cl, SnCl2 Copyright © 2014 Pearson Education, Inc. F I 2– 32 Chapter 3 Simple Bonding Theory 3.45 a. The H–O bond of methanol is more polar than the H–S bond of methyl mercaptan. As a result, hydrogen bonding holds the molecules together and requires more energy for vaporization. The larger molecular weight of methyl mercaptan has a similar effect, but the hydrogen bonding in methanol has a stronger influence. b. CO and N2 have nearly identical molecular weights, but the polarity of CO leads to dipole-dipole attractions that help hold CO molecules together in the solid and liquid states. c. The ortho isomer of hydroxybenzoic acid can form intramolecularhydrogen bonds, while the meta and para isomers tend to form dimers and larger aggregates in their hydrogen bonding. As a result of their better ability to form hydrogen bonds between molecules (intermolecular hydrogen bonds), the meta and para isomers have higher melting points (ortho, 159°; meta, 201.3°; para, 214-5°). C Intramolecular hydrogen bond O H O d. The London (dispersion) forces between atoms increase with the number of electrons, so the noble gases with larger Z have larger interatomic forces and higher boiling points. e. Acetic acid can form hydrogen-bonded dimers in the gas phase, so the total number of particles in the gas is half the number expected by using the ideal gas law. f. 3.46 H O O H3 C H O C C O H Acetone has a negative carbonyl oxygen; chloroform has a CH3 positive hydrogen, due to the electronegative character of the H3C C chlorines. As a result, there is a stronger attraction between the different kinds of molecules than between molecules of the O same kind, and a resulting lower vapor pressure. (This is an unusual case of hydrogen bonding, with no H–N, H–O, or H–F bond involved.) CH3 O Cl H C Cl g. CO has about 76 kJ contribution to its bond energy because of the electronegativity difference between C and O; attraction between the slightly positive and negative ends strengthens the bonding. Although this is not a complete explanation, it covers most of difference between CO and N2. In spite of its high bond energy, N2 is thought by some to have some repulsion in its sigma bonding because of the short bond distance. a. The trend in these angles is counter-intuitive on the basis of electronegativity arguments. Electronegativity decreases as P > As > Sb, and the C—group 15 atom—C angle is expected to decrease as P > As > Sb on the basis of less bp-bp repulsion at the group 15 atom as P > As > Sb. Both As(CF3 )3 and Sb(CF3 )3 are expected to exhibit more acute C—group 15 atom—C angles relative to P(CF3 )3 . b. On the basis of the argument above, the C—Sb—C angle of Sb(CF3 )3 should be reinvestigated. This angle is predicted to be smaller than the newly determined C—As—C angle of As(CF3 )3 . Copyright © 2014 Pearson Education, Inc. Cl Chapter 4 Symmetry and Group Theory 33 CHAPTER 4: SYMMETRY AND GROUP THEORY 4.1 4.2 a. Ethane in the staggered conformation has 2 C3 axes (the C–C line), 3 perpendicular C2 axes bisecting the C–C line, in the plane of the two C’s and the H’s on opposite sides of the two C’s. No h, 3d, i, S6. D3d. b. Ethane in eclipsed conformation has two C3 axes (the C–C line), three perpendicular C2 axes bisecting the C–C line, in the plane of the two C’s and the H’s on the same side of the two C’s. Mirror planes include h and 3d. D3h. c. Chloroethane in the staggered conformation has only one mirror plane, through both C’s, the Cl, and the opposite H on the other C. Cs. d. 1,2-dichloroethane in the trans conformation has a C2 axis perpendicular to the C–C bond and perpendicular to the plane of both Cl’s and both C’s, a h plane through both Cl’s and both C’s, and an inversion center. C2h. a. Ethylene is a planar molecule, with C2 axes through the C’s and perpendicular to the C–C bond both in the plane of the molecule and perpendicular to it. It also has a h plane and two d planes (arbitrarily assigned). D2h. b. Chloroethylene is also a planar molecule, with the only symmetry element the mirror plane of the molecule. Cs. c. 1,1-dichloroethylene has a C2 axis coincident with the C–C bond, and two mirror planes, one the plane of the molecule and one perpendicular to the plane of the molecule through both C’s. C2v. cis-1,2-dichloroethylene has a C2 axis perpendicular to the C–C bond, and in the plane of the molecule, two mirror planes (one the plane of the molecule and one perpendicular to the plane of the molecule and perpendicular to the C–C bond). C2v. trans-1,2-dichloroethylene has a C2 axis perpendicular to the C–C bond and perpendicular to the plane of the molecule, a mirror plane in the plane of the molecule, and an inversion center. C2h. 4.3 a. Acetylene has a C axis through all four atoms, an infinite number of perpendicular C2 axes, a h plane, and an infinite number of d planes through all four atoms. Dh. b. Fluoroacetylene has only the C axis through all four atoms and an infinite number of mirror planes, also through all four atoms. Cv. c. Methylacetylene has a C3 axis through the carbons and three v planes, each including one hydrogen and all three C’s. C3v. d. 3-Chloropropene (assuming a rigid molecule, no rotation around the C–C bond) has no rotation axes and only one mirror plane through Cl and all three C atoms. Cs. e. Phenylacetylene (again assuming no internal rotation) has a C2 axis down the long axis of the molecule and two mirror planes, one the plane of the benzene ring and the other perpendicular to it. C2v Copyright © 2014 Pearson Education, Inc. 34 Chapter 4 Symmetry and Group Theory 4.4 a. Napthalene has three perpendicular C2 axes, and a horizontal mirror plane (regardless of which C2 is taken as the principal axis), making it a D2h molecule. b. 1,8-dichloronaphthalene has only one C2 axis, the C–C bond joining the two rings, and two mirror planes, making it a C2v molecule. c. 1,5-dichloronaphthalene has one C2 axis perpendicular to the plane of the molecule, a horizontal mirror plane, and an inversion center; overall, C2h. d. 1,2-dichloronaphthalene has only the mirror plane of the molecule, and is a Cs molecule. a. 1,1-dichloroferrocene has a C2 axis parallel to the rings, perpendicular to the Cl–Fe–Cl h mirror plane. It also has an inversion center. C2h. b. Dibenzenechromium has collinear C6, C3, and C2 axes perpendicular to the rings, six perpendicular C2 axes and a h plane, making it a D6h molecule. It also has three v and three d planes, S3 and S6 axes, and an inversion center. c. Benzenebiphenylchromium has a mirror plane through the Cr and the biphenyl bridge bond and no other symmetry elements, so it is a Cs molecule. d. H3O+ has the same symmetry as NH3: a C3 axis, and three v planes for a C3v molecule. e. O2F2 has a C2 axis perpendicular to the O–O bond and perpendicular to a line connecting the fluorines. With no other symmetry elements, it is a C2 molecule. f. Formaldehyde has a C2 axis collinear with the C=O bond, a mirror plane including all the atoms, and another perpendicular to the first and including the C and O atoms. C2v. g. S8 has C4 and C2 axes perpendicular to the average plane of the ring, four C2 axes through opposite bonds, and four mirror planes perpendicular to the ring, each including two S atoms. D4d. h. Borazine has a C3 axis perpendicular to the plane of the ring, three perpendicular C2 axes, and a horizontal mirror plane. D3h. i. Tris(oxalato)chromate(III) has a C3 axis and three perpendicular C2 axes, each splitting a C–C bond and passing through the Cr. D3. j. A tennis ball has three perpendicular C2 axes (one through the narrow portions of each segment, the others through the seams) and two mirror planes including the first rotation axis. D2d. a. Cyclohexane in the chair conformation has a C3 axis perpendicular to the average plane of the ring, three perpendicular C2 axes between the carbons, and three v planes, each including the C3 axis and one of the C2 axes. D3d. b. Tetrachloroallene has three perpendicular C2 axes, one collinear with the double bonds and the other two at 45° to the Cl—C—Cl planes. It also has two v planes, one defined by each of the Cl–C–Cl groups. Overall, D2d. (Note that the ends of tetrachlorallene are staggered.) 4.5 4.6 Copyright © 2014 Pearson Education, Inc. Chapter 4 Symmetry and Group Theory 35 c. The sulfate ion is tetrahedral. Td. d. Most snowflakes have hexagonal symmetry (Figure 4.2), and have collinear C6, C3, and C2 axes, six perpendicular C2 axes, and a horizontal mirror plane. Overall, D6h. (For high quality images of snowflakes, including some that have different shapes, see K. G. Libbrecht, Snowflakes, Voyageur Press, Minneapolis, MN, 2008.) e. Diborane has three perpendicular C2 axes and three perpendicular mirror planes. D2h. f. 1,3,5-tribromobenzene has a C3 axis perpendicular to the plane of the ring, three perpendicular C2 axes, and a horizontal mirror plane. D3h. 1,2,3-tribromobenzene has a C2 axis through the middle Br and two perpendicular mirror planes that include this axis. C2v 1,2,4-tribromobenzene has only the plane of the ring as a mirror plane. Cs. 4.7 g. A tetrahedron inscribed in a cube has Td symmetry (see Figure 4.6). h. The left and right ends of B3H8 are staggered with respect to each other. There is a C2 axis through the borons. In addition, there are two planes of symmetry, each containing four H atoms, and two C2 axes between these planes and perpendicular to the original C2. The point group is D2d. i. A mountain swallowtail butterfly has only a mirror that cuts through the head, thorax, and abdomen. Cs j. The Golden Gate Bridge has a C2 axis and two perpendicular mirror planes that include this axis. C2v a. A sheet of typing paper has three perpendicular C2 axes and three perpendicular mirror planes. D2h. b. An Erlenmeyer flask has an infinite-fold rotation axis and an infinite number of v planes, Cv. c. A screw has no symmetry operations other than the identity, for a C1 classification. d. The number 96 (with the correct type font) has a C2 axis perpendicular to the plane of the paper, making it C2h. e. Your choice—the list is too long to attempt to answer it here. f. A pair of eyeglasses has only a vertical mirror plane. Cs. g. A five-pointed star has a C5 axis, five perpendicular C2 axes, one horizontal and five vertical mirror planes. D5h. h. A fork has only a mirror plane. Cs. i. Wilkins Micawber has no symmetry operation other than the identity. C1. Copyright © 2014 Pearson Education, Inc. 36 4.8 4.9 4.10 Chapter 4 Symmetry and Group Theory j. A metal washer has a C axis, an infinite number of perpendicular C2 axes, an infinite number of v mirror planes, and a horizontal mirror plane. Dh. a. D2h f. C3 b. D4h (note the four knobs) g. C2h c. Cs h. C8v d. C2v i. Dh e. C6v j. C3v a. D3h f. Cs (note holes) b. D4h g. C1 c. Cs h. C3v d. C3 i. Dh e. C2v j. C1 Hands (of identical twins): C2 Baseball: D2d Eiffel Tower: C4v 6 × 6: C2v Dominoes: Atomium: C3v 3 × 3: C2 5 × 4: Cs Bicycle wheel: The wheel shown has 32 spokes. The point group assignment depends on how the pairs of spokes (attached to both the front and back of the hub) connect with the rim. If the pairs alternate with respect to their side of attachment, the point group is D8d. Other arrangements are possible, and different ways in which the spokes cross can affect the point group assignment; observing an actual bicycle wheel is recommended. (If the crooked valve is included, there is no symmetry, and the point group is a much less interesting C1.) 4.11 a. Problem 3.41*: a. VOCl3: C3v b. PCl3: C3v c. SOF4: C2v d. SO3: D3h e. ICl3: C2v f. SF6: Oh g. IF7: D5h h. XeO2F4: D4h i. CF2Cl2: C2v j. P4O6: Td * Incorrectly cited as problem 4.30 in first printing of text. Copyright © 2014 Pearson Education, Inc. Chapter 4 Symmetry and Group Theory b. Problem 3.42*: a. PH3: C3v b. H2Se: C2v c. SeF4: C2v d. PF5: D3h e. IF5: C4v f. XeO3: C3v g. BF2Cl: C2v h. SnCl2: C2v i. KrF2: Dh a. CO2: Dh b. SO3: D3h c. CH4: Td d. PCl5: D3h e. SF6: Oh f. IF7: D5h a. CO2: Dh b. COF2: C2v c. NO2 : C2v d. SO3: D3h e. SNF3: C3v f. SO2Cl2: C2v g. XeO3: C3v h. SO42–: Td i. SOF4: C2v j. ClO2F3: C2v k. XeO3F2: D3h l. IOF5: C4v 37 2– j. IO2F5 : D5h 4.12 a. Figure 3.8: g. TaF83–: D4d b. 4.13 4.14 Figure 3.15: – a. px has Cv symmetry. (Ignoring the difference in sign between the two lobes, the point group would be Dh.) b. dxy has D2h symmetry. (Ignoring the signs, the point group would be D4h.) c. dx2–y2 has D2h symmetry. (Ignoring the signs, the point group would be D4h.) d. dz2 has Dh symmetry. e. fxyz has Td symmetry. a. The superimposed octahedron and cube show the matching symmetry elements. C3 The descriptions below are for the elements of a cube; each element also applies to the octahedron. E Every object has an identity operation. 8C3 Diagonals through opposite corners of the cube are C3 axes. 6C2 Lines bisecting opposite edges are C2 axes. 6C4 Lines through the centers of opposite faces are C4 axes. Although there are only three such lines, there are six axes, counting the C43 operations. 3C2 (=C42) The lines through the centers of opposite faces are C4 axes as well as C2 axes. * Incorrectly cited as problem 3.41 in first printing of text. Copyright © 2014 Pearson Education, Inc. C 2, C 4 38 4.15 Chapter 4 Symmetry and Group Theory i The center of the cube is the inversion center. 6S4 The C4 axes are also S4 axes. 8S6 The C3 axes are also S6 axes. 3h These mirror planes are parallel to the faces of the cube. 6d These mirror planes are through two opposite edges. b. Oh c. O a. There are three possible orientations of the two blue faces. If the blue faces are opposite each other, a C3 axis connects the centers of the blue faces. This axis has 3 perpendicular C2 axes, and contains three vertical mirror places (D3d). If the blue faces share one vertex of the octahedron, a C2 axis includes this vertex, and this axis includes two vertical mirror planes (C2v). The third possibility is for the blue faces to share an edge of the octahedron. In this case, a C2 axis bisects this shared edge, and includes two vertical mirror planes (C2v). b. There are three unique orientations of the three blue faces. If one blue face is arranged to form edges with each of the two remaining blue faces, the only symmetry operations are identity and a single mirror plane (Cs). If the three blue faces are arranged such that a single blue face shares an edge with one blue face, but only a vertex with the other blue face, the only symmetry operation is a mirror plane that passes through the center of the blue faces, and the point group is Cs. If the three blue faces each share an edge with the same yellow face, a C3 axis emerges from the center of this yellow face, and this axis includes three vertical mirror planes (C3v). c. 4.16 If there are four different colors, and each pair of opposite faces has the identical color, the only symmetry operations are identity and inversion (Ci). Four point groups are represented by the symbols of the chemical elements. Most symbols have a single mirror in the plane of the symbol (Cs), for example, Cs! Two symbols have D2h symmetry (H, I), and two more ( , S) have C2h. Seven exhibit C2v symmetry (B, C, K, V, Y, W, U). In some cases, the choice of font may affect the point group. For example, the symbol for nitrogen may have C2h in a sans serif font ( ) but otherwise Cs (N). The symbol of oxygen has D∞h symmetry if shown as a circle but D2h if oval. Copyright © 2014 Pearson Education, Inc. Chapter 4 Symmetry and Group Theory 4.17 4.18 a. on-deck circle Dh e. home plate C2v b. batter’s box D2h f. baseball D2d (see Figure 4.1) c. cap Cs g. pitcher C1 d. bat C v a. D2h d. D4h g. D2h b. C2 v e. C5h h. D4h c. C2 v f. C2v i. C2 y N 4.19 N S F SNF3 F2 F F F3 x F1 (top view) Symmetry Operations: F3 F1 F2 N N F1 N F3 F2 F3 after E F1 F2 after C3 after v (xz) Matrix Representations (reducible): 1 0 0 E: 0 1 0 0 0 1 2 cos 3 2 C3: sin 3 0 2 3 2 cos 3 sin 0 1 0 – 2 3 0 2 1 0 3 2 1 – 2 – 0 0 0 1 1 0 0 v(xz): 0 1 0 0 0 1 Characters of Matrix Representations: 3 0 (continued on next page) Copyright © 2014 Pearson Education, Inc. 1 39 40 Chapter 4 Symmetry and Group Theory Block Diagonalized Matrices: Irreducible Representations: E 2 1 2 C3 –1 1 3v 0 1 Coordinates Used (x, y) z Character Table: C3v A1 A2 E E 1 1 2 2 C3 3v 1 1 1 –1 –1 0 Matching Functions z Rz (x, y), (Rx, Ry) x2 + y2, z2 (x2 – y2, xy)(xz, yz) H Cl C 4.20 a. b. c. H C2h molecules have E, C2, i, and h operations. E: C2: i: 1 0 0 0 1 0 0 0 1 1 0 0 0 1 0 0 0 1 1 0 0 0 1 0 0 0 1 C Cl h: 1 0 0 0 1 0 0 0 1 These matrices can be block diagonalized into three 1 × 1 matrices, with the representations shown in the table. Bu Au (E) 1 1 (C2) –1 1 (i) –1 –1 (h) 1 –1 from the x and y coefficients from the z coefficients The total is = 2Bu +Au. d. Multiplying Bu and Au: 1 × 1 + (–1) × 1 + (–1) × (–1) + 1 × (–1) = 0, proving they are orthogonal. Copyright © 2014 Pearson Education, Inc. Chapter 4 Symmetry and Group Theory H 4.21 a. D2h molecules have E, C2(z), C2(y), C2(x), i, (xy), (xz), and (yz) operations. H C H b. 1 0 0 E: 0 1 0 0 0 1 1 0 0 C2(y): 0 1 0 0 0 1 1 0 0 i: 0 1 0 0 0 1 1 0 0 (xy): 0 1 0 0 0 1 1 0 0 (xz): 0 1 0 0 0 1 c. d. e. 1 2 3 E 3 1 1 1 C2(z) –1 –1 –1 1 C2(y) –1 –1 1 –1 H 1 0 0 0 1 0 0 0 1 C2(z): 1 0 0 C2(x): 0 1 0 0 0 1 C C2(x) –1 1 –1 –1 1 0 0 (yz): 0 1 0 0 0 1 i –3 –1 –1 –1 (xy) 1 1 1 –1 (xz) 1 1 –1 1 (yz) 1 –1 1 1 matching B3u matching B2u matching B1u 1 × 2 = 1 × 1 + (–1) × (–1) + (–1) × 1 + 1 × (–1) + (–1) × (–1) + 1 × 1 + 1 × (–1) + (–1) × 1 = 0 1 × 3 = 1 × 1 + (–1) × 1 + (–1) × (–1) + 1 × (–1) + (–1) × (–1) + 1 × (–1) + 1 × 1 + (–1) × 1 = 0 2 × 3 = 1 × 1 + (–1) × 1 + 1 × (–1) + (–1) × (–1) + (–1) × (–1) + 1 × (–1) + (–1) × 1 + 1 × 1 = 0 4.22 a. h = 8 (the total number of symmetry operations) b. A1 × E = 1 × 2 + 2 × 1 × 0 +1 × (–2) + 2 × 1 × 0 + 2 × 1 × 0 = 0 A2 × E = 1 × 2 + 2 × 1 × 0 +1 × (–2) + 2 × (–1) × 0 + 2 × (–1) × 0 = 0 B1 × E = 1 × 2 + 2 × (–1) × 0 +1 × (–2) + 2 × 1 × 0 + 2 × (–1) × 0 = 0 B2 × E = 1 × 2 + 2 × (–1) × 0 +1 × (–2) + 2 × (–1) × 0 + 2 × 1 × 0 = 0 c. E: 4 + 2 × 0 + 4 + 2 × 0 + 2 × 0 = 8 A1 : 1 + 2 × 1 + 1 + 2 × 1 + 2 × 1 = 8 A2 : 1 + 2 × 1 + 1 + 2 × 1 + 2 × 1 = 8 B1 : 1 + 2 × 1 + 1 + 2 × 1 + 2 × 1 = 8 B2 : 1 + 2 × 1 + 1 + 2 × 1 + 2 × 1 = 8 Copyright © 2014 Pearson Education, Inc. 41 42 Chapter 4 Symmetry and Group Theory d. 1 = 2A1 + B1 + B2 + E: A1: 1/8[1 × 6 + 2 × 1 × 0 + 1 × 2 + 2 × 1 × 2 + 2 × 1 × 2] = 2 A2: 1/8[1 × 6 + 2 × 1 × 0 + 1 × 2 + 2 × (–1) × 2 + 2 × (–1) × 2] = 0 B1: 1/8[1 × 6 + 2 × (–1) × 0 + 1 × 2 + 2 × 1 × 2 + 2 × (–1) × 2] = 1 B2: 1/8[1 × 6 + 2 × (–1) × 0 + 1 × 2 + 2 × (–1) × 2 + 2 × 1 × 2] = 1 E: 1/8[2 × 6 + 2 × 0 × 0 + (–2) × 2 + 2 × 0 × 2 + 2 × 0 × 2] = 1 2 = 3 A1 + 2A2 + B1: A1: 1/8[1 × 6 + 2 × 1 × 4 + 1 × 6 + 2 × 1 × 2 + 2 × 1 × 0] = 3 A2: 1/8[1 × 6 + 2 × 1 × 4 + 1 × 6 + 2 × (–1) × 2 + 2 × (–1) × 0] = 2 B1: 1/8[1 × 6 + 2 × (–1) × 4 + 1 × 6 + 2 × 1 × 2 + 2 × (–1) × 0] = 1 B2: 1/8[1 × 6 + 2 × (–1) × 4 + 1 × 6 + 2 × (–1) × 2 + 2 × 1 × 0] = 0 E: 1/8[2 × 6 + 2 × 0 × 4 + (–2) × 6 + 2 × 0 × 2 + 2 × 0 × 0] = 0 4.23 C3v 1 = 3A1 + A2 + E: A1: 1/6[1 × 6 + 2 × 1 × 3 + 3 × 1 × 2] = 3 A2: 1/6[1 × 6 + 2 × 1 × 3 + 3 × (–1) × 2] = 1 E: 1/6[2 × 6 + 2 × (–1) × 3 + 3 × 0 × 2] = 1 2 = A2 + E: A1: 1/6[1 × 5 + 2 × 1 × (–1) + 3 × 1 × (–1)] = 0 A2: 1/6[1 × 5 + 2 × 1 × (–1) + 3 × (–1) × (–1)] = 1 E: 1/6[2 × 5 + 2 × (–1) × (–1) + 3 × 0 × (–1)] = 2 Oh 3 = A1g + Eg + T1u: A1g: 1/48[6 + 0 + 0 + 12 + 6 + 0 + 0 + 0 + 12 + 12] = 1 A2g: 1/48[6 + 0 + 0 – 12 + 6 + 0 + 0 + 0 + 12 – 12] = 0 Eg: 1/48[12 + 0 + 0 + 0 + 12 + 0 + 0 + 0 + 24 + 0] = 1 T1g: 1/48[18 + 0 + 0 + 12 – 6 + 0 + 0 + 0 – 12 – 12] = 0 T2g: 1/48[18 + 0 + 0 – 12 – 6 + 0 + 0 + 0 – 12 + 12] = 0 A1u: 1/48[6 + 0 + 0 + 12 + 6 + 0 + 0 + 0 – 12 – 12] = 0 A2u: 1/48[6 + 0 + 0 – 12 + 6 + 0 + 0 + 0 – 12 + 12] = 0 Eu: 1/48[12 + 0 + 0 + 0 + 12 + 0 + 0 + 0 – 24 + 0] = 0 T1u: 1/48[18 + 0 + 0 + 12 – 6 + 0 + 0 + 0 + 12 + 12] = 1 T2u: 1/48[18 + 0 + 0 – 12 – 6 + 0 + 0 + 0 + 12 – 12] = 0 Copyright © 2014 Pearson Education, Inc. Chapter 4 Symmetry and Group Theory 4.24 The dxy characters match the characters of the B2g representation: dxy y x dxy 43 The dx2-y2 characters match the characters of the B1g representation: dx2–y2 E 1 E 1 C4 –1 C4 –1 C2 1 C2 1 C2 –1 C2 1 C2 1 C2 –1 i 1 i 1 S4 –1 S4 –1 h 1 h 1 v –1 v 1 d 1 d –1 y x dx2–y2 3– 4.25 Chiral: 4.5: O2F2, [Cr(C2O4)3] 4.6: none 4.7: screw, Wilkins Micawber 4.8: recycle symbol 4.9: set of three wind turbine blades, Flying Mercury sculpture, coiled spring 4.26 a. C4v A1 A2 B1 B2 E Point group: C4v E 18 1 1 1 1 2 2C4 2 1 1 –1 –1 0 O C2 –2 1 1 1 1 –2 2v 4 1 –1 1 –1 0 2d 2 1 –1 –1 1 0 b. = 4 A1 + A2 + 2B1 + B2 + 5E c. Translation: A1 + E (match x, y, and z) Rotation: A2 + E (match Rx, Ry, and Rz) Vibration: all that remain: 3 A1 + 2B1 + B2 + 3E d. F F F Xe F z Rz (x, y), (Rx, Ry) The character for each symmetry operation for the Xe–O stretch is +1. This corresponds to the A1 irreducible representation, which matches the function z and is therefore IR-active. Copyright © 2014 Pearson Education, Inc. O F F Xe F F 44 Chapter 4 Symmetry and Group Theory 4.27 For SF6, the axes of the sulfur should point at three of the fluorines. The fluorine axes can be chosen in any way, as long as one from each atom is directed toward the sulfur atom. There are seven atoms with three axes each, for a total of 21. Oh T1u T1g A1g A2u Eg E 21 3 3 1 1 2 8C3 0 0 0 1 1 –1 6C2 –1 –1 –1 1 –1 0 6C4 3 1 1 1 –1 0 3C2 –3 –1 –1 1 1 2 i –3 –3 3 1 1 2 6S4 –1 –1 1 1 –1 0 8S6 0 0 0 1 1 –1 3h 5 1 –1 1 1 2 6d 3 1 –1 1 –1 0 T2u T2g 3 3 0 0 1 1 –1 –1 –1 –1 –3 3 1 –1 0 0 1 –1 –1 1 (x,y,z) (Rx, Ry, Rz) (2z2 – x2 – y2, x2 – y2) (xy, xz, yz) Reduction of gives = 3T1u + T1g + A1g + Eg + T2g + T2u. T1u accounts for translation and also infrared active vibrational modes. T1g is rotation. The remainder are infrared-inactive vibrations. 4.28 a. cis-Fe(CO)4Cl2 has C2v symmetry. O C The vectors for CO stretching have the representation : C2v A1 A2 B1 B2 E 4 1 1 1 1 C2 0 1 1 –1 –1 v(xz) 2 1 –1 1 –1 v(yz) 2 1 –1 –1 1 Cl Cl Fe C O z x y n(A1) = 1/4[4 × 1 + 0 × 1 + 2 × 1 + 2 × 1] = 2 n(A2) = 1/4[4 × 1 + 0 × 1 + 2 × (–1) + 2 × (–1)] = 0 n(B1) = 1/4[4 × 1 + 0 × (–1) + 2 × 1 + 2 × (–1)] = 1 n(B2) = 1/4[4 × 1 + 0 × (–1) + 2 × (–1) + 2 × 1] = 1 = 2 A1 + B1 + B2, all four IR active. Cl b. OC OC trans-Fe(CO)4Cl2 has D4h symmetry. D4h A2u Eu E 4 1 2 2C4 0 1 0 C2 0 1 –2 2C2 2 –1 0 2C2 0 –1 0 Fe CO CO Cl i 0 –1 –2 2S4 0 –1 0 h 4 –1 2 Copyright © 2014 Pearson Education, Inc. 2v 2 1 0 2d 0 1 0 z (x,y) CO CO Chapter 4 Symmetry and Group Theory 45 Omitting the operations that have zeroes in : n(A2u) = 1/16[4 × 1 + 2 × 2 × (–1) + 4 × (–1) + 2 × 2 × 1] = 0 n(Eu) = 1/16[4 × 2 + 2 × 2 × 0 + 4 × 2 + 2 × 2 × 0] = 1 (IR active) Note: In checking for IR-active bands, it is only necessary to check the irreducible representations having the same symmetry as x, y, or z, or a combination of them. c. Fe(CO)5 has D3h symmetry. The vectors for C–O stretching have the following representation : D3h E A2 E 5 2 1 2C3 2 –1 1 3C2 1 0 –1 h 3 2 –1 2S3 0 –1 –1 3v 3 0 1 O C OC OC Fe CO C O (x, y) z n(E) = 1/12 [ (5 × 2) + (2 × 2 × –1) + (3 × 2)] = 1 n(A2) = 1/12 [(5 × 1) + (2 × 2 × 1) + (3 × 1 × –1) + (3 × –1) + (3 × 3 × 1)] = 1 There are two bands, one matching Eand one matching A2. These are the only irreducible representations that match the coordinates x, y, and z. 4.29 In 4.28a, the symmetries of the CO stretching vibrations of cis-Fe(CO)4Cl2 (C2v symmetry) are determined as 2 A1 + B1 + B2. Each of these representations matches Raman-active functions: A1 (x2, y2, z2) ; A2 (xy), B1 (xz); and B2 (yz), so all are Raman-active. In 4.28b, the symmetries of the CO stretching vibrations of trans-Fe(CO)4Cl2 (D4h symmetry) are A1g + B1g + Eu. Only A1g (x2 + y2, z2) and B1g (x2 – y2) match Raman active functions; this complex exhibits two Raman-active CO stretching vibrations. In 4.28c, the symmetries of the CO stretching vibrations of Fe(CO)5 (D3h symmetry) are 2 A1 + E + A2. Only A1 (x2 + y2, z2) and E(x2 – y2, xy) match Raman-active functions; this complex exhibits four Raman-active CO stretching vibrations. 4.30 a. The point group is C2h. b. Using the Si–I bond vectors as a basis generates the representation: C2h Ag Bg Au Bu E 4 1 1 1 1 C2 0 1 –1 1 –1 i 0 1 1 –1 –1 h 0 1 –1 –1 1 z x, y = Ag + B g + A u + B u The Au and Bu vibrations are infrared active. c. The Ag and Bg vibrations are Raman active. Copyright © 2014 Pearson Education, Inc. x2 + y2, z2 xz, yz 46 Chapter 4 Symmetry and Group Theory 4.31 trans isomer (D4h): cis isomer (C2v): The simplest approach is to consider if the number of infrared-active I–O stretches is different for these structures. (Alternatively, one could also determine the number of IR-active I–F stretches, a slightly more complicated task.) trans: D4h A1g A2u E 2 1 1 C2 2 1 1 2C4 2 1 1 C2 C2 0 0 1 1 –1 –1 i 0 1 –1 2S4 0 1 –1 h 0 1 –1 2v 2 1 1 2d 2 1 1 z There is only a single IR-active I–O stretch (the antisymmetric stretch), A2u. cis: C2v A1 B1 C2 0 1 –1 E 2 1 1 (xz) 2 1 1 (yz) 0 1 –1 z x There are two IR-active I–O stretches, the A1 and B1 (symmetric and antisymmetric). Infrared spectra should therefore be able to distinguish between these isomers. (Reversing the x and y axes would give A1 + B2. Because B2 matches y, it would also represent an IR-active vibration.) Because these isomers would give different numbers of IR-active absorptions, infrared spectra should be able to distinguish between them. The reference provides detailed IR data. 4.32 a. As One way to deduce the number of Raman-active vibrations of AsP3 is to first P determine the symmetries of all the degrees of freedom. This complex exhibits P P C3 symmetry, with the C3 axis emerging from the As atom. The (E) is 12; the x, y, and z axes of the four atoms do not shift when the identity operation is carried out. Only the As atom contributes to the character of the C3 transformation matrix; the P atoms shift during rotation about the C3 axis. The general transformation matrix for rotation about the z axis (Section 4.3.3) affords 0 as (C3) for The As atom and one P atom do not shift when a v reflection is carried out, and (v) = 2 (see the v(xz) transformation matrix in Section 4.33 for the nitrogen atom of NH3 as a model of how the two unshifted atoms of AsP3 will contribute to the character of the v transformation matrix). C3v A1 A2 E E 12 1 1 2 C3 0 1 1 –1 v 2 1 –1 0 z Rx (x, y), (Rx, Ry) x2 + y2, z2 (x2 – y2, xy), (xz, yz) Copyright © 2014 Pearson Education, Inc. Chapter 4 Symmetry and Group Theory 47 Reduction of the reducible representation affords 3 A1 + A2 + 4 E. On the basis of the character table, the translational modes of AsP3 have the symmetries A1 + E, and the rotational modes have the symmetries A2 + E. The six vibrational modes of AsP3 subsequently have the symmetries 2 A1 + 2 E. These vibrational modes are Raman-active, and four absorptions are expected (and observed) since the sets of E modes are degenerate. B. M. Cossairt, M.C. Diawara, C. C. Cummins, Science 2009, 323, 602 assigns the bands as: 313 (a1), 345 (e), 428 (a1), 557 (e) cm-1. Alternatively, the set of six bonds may be selected as the basis for a representation focused specifically on stretches of these bonds. This approach generates the following representation: C3v E 6 C3 0 v 2 This representation reduces to 2 A1 + 2 E, the same result as obtained by first considering all degrees of freedom, then subtracting the translational and rotational modes. b. As As2P2 exhibits C2v symmetry, and (like AsP3) will have six vibrational modes P P (3N – 6). Inspection of the C2v character table indicates that all vibrational As modes will be Raman active. Since each irreducible representation has a dimension of 1, the number of Raman absorptions expected is 6 (that is, there will be no degenerate vibrational modes). This prediction can be confirmed via deduction of the symmetries of the vibrational modes. As in part a, we will first determine the symmetries of all the degrees of freedom. The (E) of the transformation matrix is 12. In As2P2, the C2 axis does not pass through any atoms, and all four atoms shift upon rotation; (C2) = 0. Two atoms do not shift upon reflection through each of the v planes. The contribution to the character of the transformation matrix for each of these unshifted atoms is 1, and (v(xz)) = (v(yz)) = 2. C2v A1 A2 B1 B2 E 12 1 1 1 1 C2 0 1 1 –1 –1 v(xz) 2 1 –1 1 –1 v(yz) 2 1 –1 –1 1 z Rz x, Ry y, Rx x2, y2, z2 xy xz yz Reduction of the reducible representation affords 4 A1 + 2 A2 + 3 B1 + 3 B2. On the basis of the above character table, the translational modes of As2P2 have the symmetries A1 + B1 + B2, and the rotational modes have the symmetries A2 + B1 + B2. The six anticipated Raman-active vibrational modes of As2P2 subsequently have the symmetries 3 A1 + A2 + B1 + B2. (continued on next page) Copyright © 2014 Pearson Education, Inc. 48 Chapter 4 Symmetry and Group Theory As in part a, an alternative approach is to select the set of six bonds as the basis for a representation focused specifically on stretching vibrations. This approach generates the following representation: C2v C2 2 E 6 v(xz) 2 v(yz) 2 This representation reduces to 3 A1 + A2 + B1 + B2, the same result as obtained by first considering all degrees of freedom, then subtracting the translational and rotational modes. c. Td A1 A2 E T1 T2 P The issue here is whether or not P4 (Td) exhibits 4 Raman-active vibrations as P does AsP3. We will first determine the symmetries of all the degrees of freedom. The (E) of the transformation matrix is 12. Only one P4 atom is fixed upon rotation about each C3 axis; (C3) = 0. All four atoms are shifted upon application of the C2 and S4 axes; (C2) = (S4) = 0. Two atoms do not shift upon reflection through each of the d planes. The contribution to the character of the transformation matrix for each of these unshifted atoms is 1, and (d) = 2. E 12 1 1 2 3 3 8C3 0 1 1 –1 0 0 3C2 0 1 1 2 –1 –1 6 S4 0 1 –1 0 1 –1 6d 2 1 –1 0 –1 1 P P x2 + y2 + z2 (Rx, Ry, Rz) (x, y, z) (2z2 – x2 – y2, x2 – y2) (xy, xz, yz) Reduction of the reducible representation affords A1 + E + T1 + 2 T2. Since the symmetries of the translational modes and rotational modes are T2 and T1, respectively, the symmetries of the vibrational modes are A1 + E + T2, all of these modes are Ramanactive so three Raman absorptions are expected for P4, and P4 could potentially be distinguished from AsP3 solely on the basis of the number of Raman absorptions. The alternative approach, using the set of six P–P bonds as the basis for a representation focused specifically on bond stretches, generates the following representation: Td E 6 8C3 0 3C2 2 6 S4 0 6d 2 This representation reduces to A1 + E + T2, the same result as obtained by first considering all degrees of freedom, then subtracting the translational and rotational modes. Copyright © 2014 Pearson Education, Inc. Chapter 4 Symmetry and Group Theory 4.33 49 The possible isomers are as follows, with the triphenylphosphine ligand in either the axial (A) or equatorial (B) sites. Ph3P OC CO CO Fe CO Fe OC CO PPh3 CO (A) CO C3v (B) C2v Note that the triphenylphosphine ligand is approximated as a simple L ligand for the sake of the point group determination. Rotation about the Fe–P bond in solution is expected to render the arrangement of the phenyl rings unimportant in approximating the symmetry of these isomers in solution. The impact of the phenyl rings would likely be manifest in the IR ν(CO) spectra of these isomers in the solid-state. For A, we consider each CO bond as a vector to deduce the expected number of carbonyl stretching modes. The irreducible representation is as follows: C3v A1 A2 E E 4 1 1 2 C3 1 1 1 –1 v 2 1 –1 0 z Rx (x, y), (Rx, Ry) x2 + y2, z2 (x2 – y2, xy), (xz, yz) Reduction of the reducible representation affords 2 A1 + E. These stretching modes are IR-active and three ν(CO) absorptions are expected for A. For B, a similar analysis affords the following irreducible representation: C2v A1 A2 B1 B2 E 4 1 1 1 1 C2 0 1 1 –1 –1 v(xz) 2 1 –1 1 –1 v(yz) 2 1 –1 –1 1 z Rz x, Ry y, Rx x2, y2, z2 xy xz yz Reduction of the reducible representation affords 2 A1 + B1 + B2. These stretching modes are IRactive, and four ν(CO) absorptions are expected for A. The reported ν(CO) IR spectrum is consistent with formation of isomer A, with the triphenylphosphine ligand in the axial site. Copyright © 2014 Pearson Education, Inc. 50 Chapter 4 Symmetry and Group Theory 4.34 As in 4.33, we consider the triphenylphosphine ligand as a simple L group for point group determination. The point groups for isomers A, B, and C are as follows: PPh3 CO OC PPh3 Fe PPh3 CO Fe OC CO PPh3 (B) C2v PPh3 CO PPh3 CO (A) Fe OC CO D3h (C) Cs For A, the set of irreducible representations for the three CO stretching vibrational modes is 2 A1 + B1. These modes are all IR-active in the C2v character table, and three ν(CO) IR absorptions are expected for isomer A. For B, the set of irreducible representations for the three CO stretching vibrational modes is A1 + E. Only the E mode is IR-active in the D3h point group, and one ν(CO) IR absorption is expected for isomer B. For C, the set of irreducible representations for the three CO stretching vibrational modes is 2 A These modes are all IR-active in the Cs point group, and three ν(CO) IR absorptions are expected for isomer C. The single ν(CO) IR absorption reported for Fe(CO)3(PPh3)2 supports the presence of the D3h isomer B. The trans isomer B is reported in R. L. Keiter, E. A. Keiter, K. H. Hecker, C. A. Boecker, Organometallics, 1988, 7, 2466, and the authors observe splitting of the absorption associated with the Emode in CHCl3. The forbidden A1 stretching mode was observed as a weak absorption. 4.35 The IR spectrum exhibits two ν(CO) absorptions. The two proposed metal carbonyl fragments will be considered independently for analysis. CO OC Ti CO OC OC Ti OC CO CO C4v D4h The reducible representation for the four vectors associated with the CO bonds in the C4v fragment is as follows: C4v E 4 2C4 0 C2 0 2v 2 2d 0 Reduction of this representation affords A1 + B1 + E. The A1 and E modes are IR-active, and a titanium complex with a square pyramidal titanium tetracarbonyl fragment is supported by the IR spectral data. Copyright © 2014 Pearson Education, Inc. Chapter 4 Symmetry and Group Theory 51 For the D4h fragment, the reducible representation for the set of vectors associated with the CO bonds is: D4h E 4 C2 0 2C4 0 2C2 2 i 0 2C2 0 2v 2 h 2S4 0 4 2d 0 Reduction of the reducible representation affords A1g + B1g + Eu. Only the E modes are IR-active, and a complex with a square planar titanium tetracarbonyl fragment is expected to exhibit a single IR ν(CO) stretching absorption. This D4h possibility can therefore be ruled out on the basis of the spectrum. 4.36 (OPh)3 P A reasonable product is the C4v molecule Mo(CO)5(P(OPh)3), with CO replacing a triphenylphosphite ligand. OC The reducible representation for the five vectors associated with the CO bonds in this molecule is: C4v E 5 C2 1 2C4 1 2v 3 OC Mo CO CO C O 2d 1 Reduction of this representation affords 2 A1 + B1 + E. The A1 and E modes are IR-active, and three IR ν(CO) stretching absorptions are expected. The reported IR spectrum features three strong ν(CO) absorptions, and one “very weak” absorption attributed to the forbidden B1 mode in D. J. Darensbourg, T. L. Brown, Inorg. Chem., 1968, 7, 959. 4.37 I has C2 symmetry, with a C2 axis running right to left, perpendicular to the Cl, N, Cl, N and Cl, P, Cl, P faces. II also has C2 symmetry, with the same C2 axis as I. (Lower left corner occupied by Cl, not C.) III has only an inversion center and Ci symmetry. 4.38 An example for each of the five possible point groups: F Td : C F F 4.39 H C3v: F C F F H C2v: F C F F F Cs : H Br Br F C1 : C H Br Cl C H a. The S–C–C portion is linear, so the molecule has a C3 axis along the line of these three atoms, three v planes through these atoms and an F atom on each end, but no other symmetry elements. C3v b. The molecule has only an inversion center, so it is Ci. The inversion center is equivalent to an S2 axis perpendicular to the average plane of the ring. c. M2Cl6Br4 is Ci. d. This complex has a C3 axis, splitting the three N atoms and the three P atoms (almost as drawn), but no other symmetry elements. C3 Copyright © 2014 Pearson Education, Inc. 52 Chapter 4 Symmetry and Group Theory F e. The most likely isomer has the less electronegative Cl atoms in equatorial positions. Point group: C2v Cl Cl P F 4.40 The structures on the top row are D2d (left) and Cs (right). Those on the bottom row are C2h (left) and C4v (right). 4.41 a. C3v b. D5h c. Square structure: D2d (bottom ligand on lower left Re should be CO instead of L); Corner structure: Cs d. D3d 4.42 F Web of Science and SciFinder Scholar should be helpful, but simply searching for these symmetries using a general Internet search should provide examples of these point groups. Some examples: a. S6 : Mo2(SC6H2Me3)6 (M. H. Chisholm, J. F. Corning, and J. C. Huffman, J. Am. Chem. Soc., 1983, 105, 5924) Mo2(NMe2)6 (M. H. Chisholm, R. A. Cotton, B. A. Grenz, W. W. Reichert, L. W. Shive, and B. R. Stults, J. Am. Chem. Soc., 1976, 98, 4469) [NaFe6(OMe)12(dbm)6]+ (dbm = dibenzoylmethane, C6H5COCCOC6H5) (F. L. Abbati, A. Cornia, A. C. Fabretti, A. Caneschi, and D. Garreschi, Inorg. Chem., 1998, 37, 1430) b. T Pt(CF3)4, C44 c. Ih C20, C80 d. Th [Co(NO2)6]3–, Mo(NMe2)6 In addition to examples that can be found using Web of Science, SciFinder, and other Internet search tools, numerous examples of these and other point groups can be found in I. Hargittai and M. Hargittai, Symmetry Through the Eyes of a Chemist, as listed in the General References section. Copyright © 2014 Pearson Education, Inc. Chapter 5 Molecular Orbitals 53 CHAPTER 5: MOLECULAR ORBITALS 5.1 There are three possible bonding interactions: pz 5.2 dz2 py dyz px dxz a. Li2 has a bond order of 1.0 (two electrons in a bonding orbital; see Figures 5.7 and 5.1). Li2+ has a bond order of only 0.5 (one electron in a bonding orbital). Therefore, Li2 has the shorter bond. b. F2 has a bond order of 1.0 (see Figure 5.7). F2+ has one less antibonding (π*) electron and a higher bond order, 1.5. F2+ would be expected to have the shorter bond. c. Expected bond orders (see Figure 5.1): He2+ Bonding electrons 2 Antibonding electrons 1 HHe+ 2 0 + 1 0 H2 Bond order 1 (2 – 1) = 0.5 2 1 (2 – 0) = 1 2 1 (2 – 1) = 0.5 2 Both He2+ and H2+ have bond orders of 0.5. HHe+ would therefore be expected to have the shortest bond because it has a bond order of 1. 5.3 a. These diatomic molecules should have similar bond orders to the analogous diatomics from the row directly above them in the periodic table: P2 S2 Cl2 b. Cl2 has the weakest bond. The bond orders match those of the analogous oxygen species (Section 5.2.3): S2 + S2 S2 – c. bond order = 3 (like N2) bond order = 2 (like O2) bond order = 1 (like F2) bond order = 2.5 bond order = 2 bond order = 1.5 S2– has the weakest bond. Bond orders: NO+ NO NO– bond order = 3 (isoelectronic with CO, Figure 5.13) bond order = 2.5 (one more (antibonding) electron than CO) bond order = 2 (two more (antibonding) electrons than CO) NO– has the lowest bond order and therefore the weakest bond. Copyright © 2014 Pearson Education, Inc. 54 Chapter 5 Molecular Orbitals 5.4 O22– has a single bond, with four electrons in the π* orbitals canceling those in the π orbitals. – O2 has three electrons in the π* orbitals, and a bond order of 1.5. The Lewis structures have an unpaired electron and an average bond order of 1.5. O2 has two unpaired electrons in its π* orbitals, and a bond order of 2. The simple Lewis structure has all electrons paired, which does not match the paramagnetism observed experimentally. Bond lengths are therefore in the order – O22– > O2 > O2, and bond strengths are the reverse of this order. Copyright © 2014 Pearson Education, Inc. Chapter 5 Molecular Orbitals 5.5 Bond Order (Figures 5.5 and 5.7) C22– N22– O22– O2 Bond Distance (pm) 3 2 1 2 55 Unpaired Electrons 119 122.4 149 (very long) 120.7 0 2 0 2 The bond distance in N22– is very close to the expected bond distance for a diatomic with 12 valence electrons, as shown in Figure 5.8. 5.6 The energy level pattern would be similar to the one shown in Figure 5.5, with the interacting orbitals the 3s and 3p rather than 2s and 2p. All molecular orbitals except the highest would be occupied by electron pairs, and the highest orbital (u*) would be singly occupied, giving a bond order of 0.5. Because the bond in Ar2+ would be weaker than in Cl2, the Ar–Ar distance would be expected to be longer (calculated to be > 300 pm; see the reference). 5.7 a. The energy level diagram for NO is on the right. The odd electron is in a π2p* orbital. 2p 2p b. c. d. O is more electronegative than N, so its orbitals are slightly lower in energy. The bonding orbitals are slightly more concentrated on O. The bond order is 2.5, with one unpaired electron. NO+ NO – NO 5.8 2p 2p 2s 2s Bond order = 3 shortest bond (106 pm) 2s Bond order = 2.5 N NO intermediate (115 pm) Bond order = 2 longest bond (127 pm), two electrons in antibonding orbitals. – 2p 2p 2s O a. The CN energy level diagram is similar to that of NO (Problem 5.7) without the antibonding π* electron. b. The bond order is three, with no unpaired electrons. c. The HOMO is the 2p orbital, which can interact with the 1s of the H+, as in the diagram at right. The bonding orbital has an energy near that of the π orbitals; the antibonding orbital becomes the highest energy orbital. Copyright © 2014 Pearson Education, Inc. 56 5.9 5.10 Chapter 5 Molecular Orbitals a. A diagram is sketched at the right. Since the difference in valence orbital potential energy between the 2s of N (-25.56 eV) and the 2p of F (-18.65 eV) is 6.91 eV, the 2p orbital is expected to be higher in energy 2p* relative to the degenerate 2p set. b. NF is isoelectronic (has the same number of valence electrons) with O2. Therefore, NF is predicted to be paramagnetic with a bond order of 2. The populations of the bonding (8 electrons) and antibonding (4 electrons) molecular orbitals in the diagram suggest a double bond. 2p 2p* 2p 2p 2p 2s 2s* 2s c. The 2s, 2s*, 2p, and 2p* orbitals 2s exhibit Cv symmetry, with the NF NF bond axis the infinite-fold N F rotation axis. The 2p and 2p* orbitals exhibit Cs symmetry. The latter do not possess C2 rotation axes coincident to the infinite-fold rotation axis of the orbitals on the basis of the change in wave function sign upon crossing the nodes on the bond axis. a. OF– has 14 valence electrons, four in the π2p* orbitals (see the diagram in the answer to Problem 5.9). b. The net result is a single bond between two very electronegative atoms, and no unpaired electrons. c. The concentration of electrons in the π* orbital is more on the O, so combination with the positive proton at that end is more likely. In fact, H+ bonds to the oxygen atom, at an angle of 97°, as if the bonding were through a p orbital on O. 5.11 The molecular orbital description of KrF+ would predict that this ion, which has the same number of valence electrons as F2, would have a single bond. KrF2 would also be expected, on the basis of the VSEPR approach, to have single Kr–F bonds, in addition to three lone pairs on Kr. Reported Kr–F distances: KrF+: 176.5-178.3 pm; KrF2: 186.8-188.9 pm. The presence of lone pairs in KrF2 may account for the longer bond distances in this compound. 5.12 a. The KrBr+ energy level diagram is at the right. b. The HOMO is polarized toward Br, since its energy is closer to that of the Br 4p orbital. c. Bond order = 1 d. Kr is more electronegative. Its greater nuclear charge exerts a stronger pull on the shared electrons. HOMO 4p 4p 4s 4s Kr Copyright © 2014 Pearson Education, Inc. KrBr+ Br Chapter 5 Molecular Orbitals 5.13 57 The energy level diagram for SH– is shown below. A bond order of 1 is predicted. The S orbital energies are –22.7 eV (3s) and –11.6 eV (3p); the 1s of H has an energy of –13.6 eV. Because of the difference in their atomic orbital energies, the 1s orbital of hydrogen and the 3s orbital of sulfur interact only weakly; this is shown in the diagram by a slight stabilization of the lowest energy molecular orbital with respect to the 3s orbital of sulfur. This lowest energy orbital is essentially nonbonding. These orbitals are similar in appearance to those of HF in Example 5.3, with more balanced contribution of the hydrogen 1s and sulfur valence orbitals since the valence orbitals of sulfur are closer to the energy of the hydrogen 1s orbital than the valence orbitals of fluorine. 3p 1s 3s H 5.14 a. SH– S The group orbitals on the hydrogen atoms are 2p and 1s The first group orbital interacts with the 2s orbital on carbon: 2s And the second group orbital interacts with a 2p orbital on carbon: Carbon’s remaining 2p orbitals are nonbonding. b. C H C H H H Linear CH2 is a paramagnetic diradical, with one electron in each of the px and py orbitals of carbon. (A bent singlet state, with all electrons paired, is also known, with a calculated bond angle of approximately 130°.) Copyright © 2014 Pearson Education, Inc. 58 Chapter 5 Molecular Orbitals H 5.15 a. BeH2 Be H H Be H Group Orbitals: Be Orbitals with Matching Symmetry: MO Diagram: b. 5.16 2s 2pz The energy level diagrams for CH2 and BeH2 feature the same orbital interactions. One difference is that the different number of valence electrons renders the linear species BeH2 diamagnetic and CH2 paramagnetic. The energy difference between the Be and H valence orbitals is larger than that between the valence orbitals of C and H, and both the 2s and 2p orbitals of Be are higher in energy than the 1s orbital of H. The result is greater bond polarity in BeH2. BeF2 uses s and p orbitals on all three atoms, and is isoelectronic with CO2. The energy level diagram for CO2 in Figure 5.25 can be used as a guide, with the orbitals of Be higher in energy than those of C and the orbitals of F lower in energy than those of O. Calculated molecular orbital shapes are below, for comparison for those of CO2 in Figure 5.25. Copyright © 2014 Pearson Education, Inc. Chapter 5 Molecular Orbitals 4b1u 4ag LUMO (2) 2b2u, 2b3u HOMO (2) 1b2g, 1b3g Bonding (2) 1b2u, 1b3u 3b1u Nonbonding 3ag, Bonding 2b1u 2ag Of the occupied orbitals, there are three bonding (two π and one ) and five nonbonding (two π and three ). (References: W. R. Wadt, W. A. Goddard III, J. Am. Chem. Soc., 1974, 96, 5996; R. Gleiter, R. Hoffmann, J. Am. Chem. Soc., 1968, 90, 5457; C. W. Bauschlicher, Jr., I. Shavitt, J. Am. Chem. Soc., 1978, 100, 739.) Copyright © 2014 Pearson Education, Inc. 59 60 5.17 Chapter 5 Molecular Orbitals a. The group orbitals of the fluorines are: x F b. Xe F z F The matching orbitals on xenon are: F F F pz s, dz2 pz s, dz2 dxz px dyz py F F Xe F F Xe Xe F F Xe F H 5.18 The point group of TaH5 is C4v. 1. H H Ta H H z 2. Axes can be assigned as shown: Ta y x 3. Construction of reducible representation: C4v A1 A2 B1 B2 E E 5 1 1 1 1 2 2C4 1 1 1 –1 –1 0 C2 1 1 1 1 1 –2 2v 3 1 –1 1 –1 0 2d 1 1 –1 –1 1 0 z Rz (x, y), (Rx, Ry) Copyright © 2014 Pearson Education, Inc. z2 x2–y2 xy (xz, yz) F Chapter 5 Molecular Orbitals reduces to 2 A1 + B1 + E 5, 6. Two group orbitals, shown at right, have A1 symmetry. These may interact with the pz, dz2, and s orbitals of Ta. Ta Ta One group orbital has B1 symmetry. It can interact with the dx2–y2 orbital of Ta. A degenerate pair of group orbitals has E symmetry. It may interact with the (px, py) and (dxz, dyz) pairs of Ta. 5.19 61 Ta Ta Ta The energy level diagram for O3 with the simple combinations of s and p orbitals is shown below. Mixing of s and p orbitals is fairly small, showing mostly in the four lowest orbitals. The order of orbitals may vary depending on the calculation method (for example, PM3 and AM1 methods reverse the orders of HOMO and HOMO –1). 12 11 10 9 8 7 6 5 4 3 2 1 O O O O O Copyright © 2014 Pearson Education, Inc. O 62 5.20 Chapter 5 Molecular Orbitals SO3 has molecular orbitals similar to those of BF3 (Section 5.4.6). The irreducible representations below are labeled for the oxygen orbitals. D3h (s) (py) (px) (pz) A1 A2 A2 E E 5.21 π E 3 3 3 3 1 1 1 2 2 2C3 0 0 0 0 1 1 1 –1 –1 (s) = A1 + E (py) = A1 + E (px) = A2 + E (pz) = A2 + E 3C2 1 1 –1 –1 1 –1 –1 0 0 h 3 3 3 –3 1 1 –1 2 –2 2S3 0 0 0 0 1 1 –1 –1 1 3v 1 1 –1 1 1 –1 1 0 0 z (x, y) Sulfur s, px, and py Sulfur s, px, and py Sulfur px and py Sulfur pz As a cyclic (triangular) ion, H3+ has a pair of electrons in a bonding orbital and two vacant orbitals that are slightly antibonding: H H + E' H A 1' H3 + 5.22 – The thiocyanate ion, SCN , has molecular orbitals similar to those of CO2, but with more mixing between the s orbital of C and the s and p orbitals of S. The valence orbital potential energies of S are very close to those of C, and those of N are only slightly lower. There is significant double bonding in thiocyanate on the basis of this excellent orbital energy compatibility. This is consistent with the resonance structures shown at right, with the top structure favored. 1– S C N 1– S C 1+ S N 2– C – N For cyanate, OCN , the s and p orbitals of carbon effectively interact with valence orbitals on the N side, but less on the O side because the oxygen orbital energies are much more negative. The structures described in Section 3.1.3 (a mix of two double bonds and O–C and CN) fit this ion also. – For fulminate, CNO , the large differences between the C and O orbital energies render the contributions of the terminal atom orbitals to the group orbitals relatively uneven. A practical result of this imbalance is less delocalization of electron density within this anion. This is Copyright © 2014 Pearson Education, Inc. Chapter 5 Molecular Orbitals 5.23 63 suggested by the formal charge effects described in Example 3.3. As a result, the bonding in this ion is weak. Fulminate is stable only when complexed with a metal ion. – – The highest occupied orbitals in SCN or OCN are π nonbonding orbitals (see Figure 5.25 for the similar CO2 orbitals). Combination with H+ or with metal ions depends on the energy match of these orbitals with those of the positive ion. The preference for attack can be examined by comparing the energies of the valence orbitals of the terminal atoms that contribute to these nonbonding molecular orbitals. For example, the H 1s orbital energy matches the energy of the N valence orbitals orbital better than either the S or O valence orbitals, and the nitrogen atom is the site of protonation in both anions. The energies of metal ion valence orbitals vary from element to element (and with oxidation state). Some metal ions will be more compatible energetically with the S orbitals while others will be more compatible with the N orbitals. These electronic effects contribute to which site of thiocyanate is appropriate for bonding to metals. The S can also use the empty 3d orbitals to accept electron density via π bonding from some metal ions. – 5.24 The CN molecular orbitals are similar to those of CO (Figure 5.13), but with less difference between the C and N atomic orbital energies than between C and O orbitals. As a result, the HOMO should be more evenly balanced between the two atoms, and bonding at both ends seems – more likely with CN relative CO. The Prussian blue structures (Fe4[Fe(CN)6]3 or KFe(Fe(CN)6) – have iron and CN in arrangements that have both Fe–C and Fe–N bonds. 5.25 a. The resonance structures were considered in Problem 3.3, showing bent structures with primarily double bond character in S=N and single bonding in N–O or S–O. – SNO is more stable on the basis of formal charges. b. The molecular orbitals should be similar to those of O3, with more mixing of s and p orbitals because of the difference between atomic orbital energies of S and O as terminal atoms. The π bonding, nonbonding, and antibonding orbitals are numbers 6, 9, and 10 in the ozone diagram in the Problem 5.19 answer. The relative contributions of the valence orbitals of each atom to the π molecular orbitals – of SNO can be partially rationalized on the basis of electronegativity. In the π bonding orbital, the electron density is highest on the most electronegative O atom, a reasonable expectation for the most stabilized π interaction. In the π antibonding orbital, the electron density is highest on the least electronegative S atom, a feature that contributes to the – destabilization of this orbital. The π molecular orbitals of NSO are not as clearly explained by these electronegativity arguments, possibly due to the ability of S to expand its valence shell to increase its participation in bonding as a central atom. Copyright © 2014 Pearson Education, Inc. 64 Chapter 5 Molecular Orbitals c. π – S—N—O πn – S—N—O S—N—O π – N—S—O πn – N—S—O π* – N—S—O π* – The calculated bond distances for these ions are: ion N–S S–O N–O – 171 pm 120 pm SNO – 146 pm 149 pm NSO – NSO has the shorter N–S bond and the higher energy N–S stretching vibration. 5.26 H2O has C2v symmetry. Figure 5.26 defines the coordinate system to be used. The representation that describes the symmetry of the group orbitals (Section 5.4.3) is A1 B1 . The next step is to track the fate of one of the hydrogen 1s orbitals as the symmetry operations are carried out: Original Orbital E C2 v( xz) v( yz) H1s(a) becomes…. H1s(a) H1s(b) H1s(a) H1s(b) H1s(b) H1s(a) Now we multiply these outcomes by the characters associated with each operation for the A1 and B1 representations, and then add the results to obtain the linear combinations of the H 1s atomic orbitals that define the group orbitals. v( xz) C2 E v( yz) SALCs A1 H1s(a) + H1s(b) + H1s(a) + H1s(b) 2( H1s(a) ) 2(H1s(b) ) B1 H1s(a) – H1s(b) + H1s(a) – H1s(b) 2( H1s(a) ) 2(H1s(b) ) Each group orbital equation must be normalized, so that the sum of the squares of the coefficients 1 within each equation equals 1. The normalization factors, N (ca2 cb2 ) , where ca and cb are the lowest common integer coefficients for the hydrogen 1s orbital wave functions in each SALC, are: A1 : N (1) 2 (1) 2 1 1 2 B1 : N (1) 2 (1) 2 ) 1 This results in the normalized SALC equations for the two group orbitals: 1 1 (H a ) ( H b ) B1 : (H a ) (H b ) . A1 : 2 2 Copyright © 2014 Pearson Education, Inc. 1 . 2 Chapter 5 Molecular Orbitals 65 SALC Coefficients and Evidence of Normalization: Squares of SALC Coefficients Coefficients in Normalized SALCs A1 ca cb ca2 cb2 1 1 2 1 2 1 1 2 1 2 1 2 1 2 1 1 2 2 Sum of the squares for each 1s wavefunction must total 1 for an identical contribution of each atomic orbital to the group orbitals B1 5.27 Sum of Squares =1 for Normalization 1 1 b The irreducible representation associated with the 2s set is A1 E . The atoms will be labeled as shown. The next step is to track the fate of one of the fluorine 2s orbitals as the symmetry operations are carried out: a c Original Orbital E C3 C32 C2(a) C2(b) C2(c) h S3 S32 v(a) v(b) v(c) 2sa becomes… 2sa 2sb 2sc 2sa 2sc 2sb 2sa 2sb 2sc 2sa 2sc 2sb Now we multiply these outcomes by the characters associated with each operation for A1 in the D3h character table to obtain the linear combination. All of the characters are 1 for the totally symmetric A1 irreducible representation. Therefore: A1 : 2sa 2sb 2sc 2sa 2sc 2sb 2sa 2sb 2sc 2sa 2sc 2sb 4(2sa ) 4(2sb ) 4(2sc ) The lowest common integer coefficient is 1, and N (1) 2 (1) 2 (1)2 normalized A1 SALC is 1 1 3 . The 1 (2sa ) (2sb ) (2sc ) . Next we multiply the terms in the 3 table above by the characters of the E irreducible representation. The characters for C2 and v are 0, so multiplication by these characters leads to no contribution to the SALC: E : 2(2s ) 2s 2s 2(2s ) 2s 2s 4(2s ) 2(2s ) 2(2s ) . a b c a b c a b c Reduction to the lowest common integer coefficient affords 2(2sa ) (2sb ) (2sc ) , and N (2) 2 (1) 2 (1)2 1 1 6 . The normalized E SALC is therefore 1 2((2sa ) (2sb ) (2 pc ) . 6 The remaining E SALC can be deduced by examination of the coefficients and the sums of their squares. Copyright © 2014 Pearson Education, Inc. 66 Chapter 5 Molecular Orbitals Coefficients of Normalized SALCs A1 E Squares of SALC Coefficients ca cb cc ca2 cb2 cc2 1 1 1 3 2 3 1 3 1 1 3 2 3 1 3 1 6 1 2 1 3 1 6 1 2 1 1 6 6 1 6 1 2 2 Sum of the squares for each 2s wavefunction must total 1 for an identical contribution of each atomic orbital to the group orbitals E 0 Sum of the Squares = 1 for Normalization 0 1 1 1 1 Since the sum of the squares of the coefficients for 2sa equals 1 without any contribution from the second E SALC, this SALC must have ca 0 . The squares of cb and cc must equal 1 2 to satisfy the normalization and identical contribution from each orbital requirements. Since E matches the symmetry of the x and y axes, and the origin is the center of the group orbital (the boron atom), one of the coefficients must be positive and the other negative. The second E 1 (2sb ) (2sc ) . SALC is therefore 2 b The irreducible representation associated with the 2 p y set is also A1 E . The atoms are labeled as shown. The fate of one of the fluorine 2 p y orbitals as the symmetry operations are carried out is: E 2 p y(a) 2 p y(a) C3 C32 C2(a) C2(b) c C2(c) h S3 S32 a v(a) v(b) v(c) 2 p y(b) 2 p y(c) 2 p y(a) 2 p y(c) 2 p y(b) 2 p y(a) 2 p y(b) 2 p y(c) 2 p y(a) 2 p y(c) 2 p y(b) becomes… Now we multiply these outcomes by the characters associated with each operation for A1 in the D3h character table. All of the characters are 1 for the A1 irreducible representation. Therefore: A1 : 2 p y ( a ) 2 p y (b) 2 p y (c ) 2 p y ( a ) 2 p y ( c ) 2 p y (b ) 2 p y ( a ) 2 p y (b) 2 p y ( c ) 2 p y ( a ) 2 p y ( c ) 2 p y (b) 4(2 p y ( a ) ) 4(2 p y (b) ) 4(2 p y (c ) ). The lowest common integer coefficient is 1, and N (1) 2 (1) 2 (1)2 normalized A1 SALC is therefore 1 1 (2 p y(a) ) (2 p y(b) ) (2 p y(c) ) . 3 Copyright © 2014 Pearson Education, Inc. 1 3 . The Chapter 5 Molecular Orbitals 67 Next we multiply the terms in the table above by the characters of the E irreducible representation. The characters for C2 and v are 0, so multiplication by these characters leads to no contribution to the SALC: E : 2(2 p y(a) ) 2 p y(b) 2 p y(c) 2(2 p y(a) ) 2 p y(b) 2 p y(c) 4(2 p y(a) ) 2(2 p y(b) ) 2(2 p y(c) ) . Reduction to the lowest common integer coefficient affords 2(2 p y(a) ) (2 p y(b) ) (2 p y(c) ) , and N (2) 2 (1) 2 (1)2 1 1 6 . The normalized E SALC is 1 2((2 p y(a) ) (2 p y(b) ) (2 p y(c) ) . 6 The remaining E SALC can be deduced by examination of the coefficients and the sums of their squares. Coefficients of Normalized SALCs A1 E Squares of SALC Coefficients ca cb cc ca2 cb2 cc2 1 1 1 3 2 3 1 3 1 1 3 1 3 1 3 1 2 3 1 6 1 6 1 0 1 2 1 2 1 1 1 1 6 E Sum of the Squares = 1 for Normalization 0 6 1 6 1 2 2 Sum of the squares for each 2py wavefunction must total 1 for an identical contribution of each atomic orbital to the group orbitals Since the sum of the squares of the coefficients for 2 p y(a) equals 1 without any contribution from the second E SALC, this SALC must have ca 0 . The squares of cb and cc must equal 1 2 to satisfy the normalization and identical contribution from each orbital requirements. Since E matches the symmetry of the x and y axes, and the origin is the center of the group orbital (the boron atom), one of the coefficients must be positive and the other negative. The second E 1 (2 p y(b) ) (2 p y(c) ) . SALC is 2 The irreducible representation associated with the 2 px set is A2 E . The b atoms are labeled as shown. The fate of one of the fluorine 2 px orbitals as the symmetry operations are carried out is: c Copyright © 2014 Pearson Education, Inc. a 68 Chapter 5 Molecular Orbitals 2 px(a) becomes… E C3 C32 C2(a) C2(b) C2(c) h S3 S32 v(a) v(b) v(c) 2 px(a) 2 px(b) 2 px(c) 2 px(a) 2 px(c) 2 px(b) 2 px(a) 2 px(b) 2 px(c) 2 px(a) 2 px(b) 2 px(c) Now we multiply these outcomes by the characters associated with each operation for A2 in the D3h character table: A2 : 2 px(a) 2 px(b) 2 px(c) 2 px(a) 2 px(c) 2 px(b) 2 px(a) 2 px(b) 2 px(c) 2 px(a) 2 p y(c) 2 px(b) 4(2 px(a) ) 4(2 px(b) ) 4(2 px(c) ) The lowest common integer coefficient is 1, and N (1) 2 (1) 2 (1)2 normalized A2 SALC is 1 1 3 . The 1 (2 px(a) ) (2 px(b) ) (2 px(c) ) . 3 Next we multiply the terms in the table by the characters of the E irreducible representation. E : 2(2 px(a) ) 2 px(b) 2 px(c) 2(2 p y(a) ) 2 px(b) 2 px(c) 4(2 px(a) ) 2(2 px(b) ) 2(2 px(c) ) Reduction to the lowest common integer coefficient affords 2(2 px(a) ) (2 px(b) ) (2 px(c) ) , and N (2) 2 (1) 2 (1)2 1 1 6 . The normalized E SALC is 1 2((2 px(a) ) (2 px(b) ) (2 px(c) ) . 6 The remaining E SALC can be deduced by examination of the coefficients and the sums of their squares. Coefficients of Normalized SALCs A2 E E Squares of SALC Coefficients Sum of the Squares = 1 for Normalization Requirement ca cb cc ca2 cb2 cc2 1 1 1 3 2 3 1 3 1 1 3 1 3 1 3 1 2 3 1 6 1 6 1 0 1 2 1 2 1 1 1 1 6 0 6 1 6 1 2 2 Sum of the squares for each 2px wavefunction must total 1 for an identical contribution of each atomic orbital to the group orbitals Copyright © 2014 Pearson Education, Inc. Chapter 5 Molecular Orbitals 69 Since the sum of the squares of the coefficients for 2 px(a) equals 1 without any contribution from the second E SALC, this SALC must have ca 0 . The squares of cb and cc must equal 1 2 to satisfy the normalization and identical contribution from each orbital requirements. Because E matches the symmetry of the x and y axes, and the origin is the center of the group orbital (the boron atom), one of the coefficients must be positive and the other negative. The second E 1 (2 px(b) ) (2 px(c) ) . SALC is 2 The irreducible representation associated with the 2 pz set is b A2 E . The atoms are labeled as shown. The fate of one of the fluorine 2 pz orbitals as the symmetry operations are carried out is: 2 pz(a) becomes… a c E C3 C32 C2(a) C2(b) C2(c) h S3 S32 v(a) v(b) v(c) 2 pz(a) 2 pz(b) 2 pz(c) 2 pz(a) 2 pz(c) 2 pz(b) 2 pz(a) 2 pz(b) 2 pz(c) 2 pz(a) 2 pz(c) 2 pz(b) Now we multiply these outcomes by the characters associated with each operation for A2 in the D3h character table: A2: 2 pz(a) 2 pz(b) 2 pz(c) 2 pz(a) 2 pz(c) 2 pz(b) 2 pz(a) 2 pz(b) 2 pz(c) 2 px(a) 2 pz(c) 2 pz(b) 4(2 pz(a) ) 4(2 pz(b) ) 4(2 pz(c) ) The lowest common integer coefficient is 1, and N (1) 2 (1) 2 (1)2 normalized A2 SALC is 1 1 3 . The 1 (2 pz(a) ) (2 pz(b) ) (2 pz(c) ) . 3 Next we multiply the terms in the table by the characters of the E irreducible representation: E : 2(2 pz(a) ) 2 pz(b) 2 pz(c) 2(2 pz(a) ) 2 pz(b) 2 pz(c) 4(2 pz(a) ) 2(2 pz(b) ) 2(2 pz(c) ). Reduction to the lowest common integer coefficient affords 2(2 pz(a) ) (2 pz(b) ) (2 pz(c) ) , and N (2) 2 (1) 2 (1)2 1 1 6 . The normalized E SALC is 1 2((2 pz(a) ) (2 pz(b) ) (2 pz(c) ) . 6 The remaining E SALC can be deduced by examination of the coefficients and the sums of their squares. Copyright © 2014 Pearson Education, Inc. 70 Chapter 5 Molecular Orbitals Coefficients of Normalized SALCs A2 E ca cb cc ca2 cb2 cc2 1 1 1 3 2 3 1 3 1 1 3 1 3 1 3 1 2 3 1 6 1 6 1 0 1 2 1 2 1 1 1 1 6 E Squares of SALC Coefficients Sum of the Squares = 1 for Normalization Requirement 6 1 0 6 1 2 2 Sum of the squares for each 2pz wavefunction must total 1 for an identical contribution of each atomic orbital to the group orbitals Since the sum of the squares of the coefficients for 2 pz(a) equals 1 without any contribution from the second E SALC, this SALC must have ca 0 . The squares of cb and cc must equal 1 2 to satisfy the normalization and identical contribution from each orbital requirements. Since E has the same symmetry as the xz and yz orbitals that have nodes defined by the yz and xz planes, respectively, one of the coefficients must be positive and the other negative. The second E 1 (2 pz(b) ) (2 pz(c) ) . SALC is 2 5.28 The point group is D4h . The reducible representation that describes the symmetries of the group orbitals is: D4h E 2C4 C2 4 0 0 2C2 2 2C2 0 i 2S 4 h 2 v 2 d 0 0 4 2 0 Reduction affords A1g B1g Eu . To deduce the SALCs, we need to track the fate of one of the 3s orbitals through each symmetry operation of the character table. 3s( A) becomes… E C4 C43 C2 C2 (x) C2 ( y) C2 (1) C2 (2) 3s( A) 3s(B) 3s(D) 3s(C) 3s(C) 3s( A) 3s(D) 3s(B) i S4 h v (x) v ( y) d (1) d (2) 3s(C) 3s(B) S43 3s(D) 3s(D) 3s(B) 3s( A) 3s(C) Since all of the characters for A1g are 1, the SALC for A1g is: Copyright © 2014 Pearson Education, Inc. 3s( A) Chapter 5 Molecular Orbitals 71 A1g : 3s( A) 3s( B) 3s(D) 3s(C) 3s(C) 3s( A) 3s(D) 3s(B) 3s(C) 3s( B) 3s(D) 3s( A) 3s(C) 3s( A) 3s( D) 3s( B) 4(3s( A)) 4(3s(B)) 4(3s(C)) 4(3s(D)) This simplifies to 3s( A) 3s(B) 3s(C) 3s(D) , which provides a normalization constant of 1 1 1 and the SALC (3s( A)) (3s(B)) (3s(C)) (3s(D)) . 2 2 Multiplication of the orbitals in the above table by the B1g characters gives: N (1) 2 (1) 2 (1)2 (1)2 B1g : 3s( A) 3s( B) 3s(D) 3s(C) 3s(C) 3s( A) 3s(D) 3s(B) 3s(C) 3s( B) 3s(D) 3s( A) 3s(C) 3s( A) 3s( D) 3s( B) 4(3s( A)) 4(3s(B)) 4(3s(C)) 4(3s(D)) This simplifies to 3s( A) 3s(B) 3s(C) 3s(D) , which provides a normalization constant of N (1) 2 (1) 2 (1)2 (1)2 1 1 1 and (3s( A)) (3s(B)) (3s(C)) (3s(D)) 2 2 as the B1g normalized SALC. Multiplication of the orbitals in the above table by the Eu characters affords: Eu : 2(3s( A)) 2(3s(C)) 2(3s(C)) 2(3s( A)) 4(3s( A)) 4(3s(C)) This simplifies to 3s ( A) 3s (C ), which provides a normalization constant of N (1) 2 (1) 2 1 1 2 and 1 (3s( A)) (3s(C)) as one of the normalized Eu 2 SALCs. The equation for the other Eu SALC can be deduced by consideration of the normalized SALC coefficients and their squares. Coefficients of Normalized SALCs A1g B1g Eu Eu Squares of SALC Coefficients cA cB cC cD c A2 cB2 cC2 cD2 1 2 1 2 1 1 2 1 2 1 2 1 2 1 1 2 1 2 1 4 1 4 1 2 1 4 1 4 1 4 1 4 1 2 1 4 1 4 2 0 0 1 0 2 0 1 2 2 Sum of the squares for each 3s wavefunction must total 1 for an identical contribution of each atomic orbital to the group orbitals 0 1 1 0 1 1 0 1 2 0 1 2 1 1 1 1 Copyright © 2014 Pearson Education, Inc. Sum of the Squares = 1 for Normalization Requirement 72 Chapter 5 Molecular Orbitals The sum of c A2 and cC2 equals 1 without any contribution from the second Eu SALC; c A cC 0 for this second Eu SALC. A magnitude of 1 is required for both c B and cD of the second Eu SALC on 2 the basis of the normalization requirement and equal contribution of each atomic orbital to the group orbital’s requirement. The alternate signs are required since Eu has the symmetry of the x and y axes; these Eu group orbitals need to match the symmetry of the valence p orbitals of the central atom. The 1 (3s(B)) (3s( D)) . 2 Sketches of these group orbitals are below, using the coordinate system specified in this problem. Note the scaling of the orbitals to reflect the larger contribution of the 3s orbitals in the Eu normalized equation for the second Eu SALC is SALCs compared to that in A1g and B1g SALCs. A A B B D A1g D B1g C C A D B Eu 5.29 Eu C For this question, we will label the 2 pz orbitals simply by their letters (A–F) for clarity. The first task is to track the fate of A through all of the D6 h symmetry operations: A becomes… E C6 C3 B C65 F A i S3 S32 -D -C -E h C2 (1) -A d (1) C2 (2) -C d (2) C2 (3) -E d (3) C2 (1) -B v (1) C2 (2) -D v (2) C2 (3) -F v (3) -A B D F A C E C2 C C32 E S6 S65 -B -F D Multiplication by the characters of each irreducible representation of the D6 h character table (and for each symmetry operation) is a tedious, but effective, method to deduce the SALCs. A1g : A B F C E D A C E B D F D C E B F A B D F A C E 0 A2g : A B F C E D A C E B D F D C E B F A B D F A C E 0 Copyright © 2014 Pearson Education, Inc. Chapter 5 Molecular Orbitals 73 B1g : A B F C E D A C E B D F D C E B F A B D F A C E 0 B2g : A B F C E D A C E B D F D C E B F A B D F A C E 4 A 4B 4C 4D 4E 4F E1g : 2 A B F C E 2D 2D C E B F 2 A 4 A 2B 2C 4D 2E 2F E2g : 2 A B F C E 2D 2D C E B F 2 A 0 A1u : A B F C E D A C E B D F D C E B F A B D F A C E 0 A2u : A B F C E D A C E B D F D C E B F A B D F A C E 4 A 4B 4C 4D 4E 4F B1u : A B F C E D A C E B D F D C E B F A B D F A C E 0 B2u : A B F C E D A C E B D F D C E B F A B D F A C E 0 E1u : 2 A B F C E 2D 2D C E B F 2 A 0 E2u : 2 A B F C E 2D 2D C E B F 2 A 4 A 2B 2C 4D 2E 2F The six group orbitals have the symmetries B2g , A2u , E1g , and E2u , expressed in simplified form below, with normalization constants shown at right. 1 B2g : A B C D E F B2g : N (1) 2 (1) 2 (1)2 (1)2 (1)2 (1)2 A2u : A B C D E F A2u : N (1) 2 (1) 2 (1)2 (1)2 (1)2 (1)2 E1g : 2 A B C 2D E F E1g : N (2) 2 (1) 2 (1)2 (2)2 (1)2 (1)2 E2u : 2 A B C 2D E F E2u : N (2) 2 (1) 2 (1)2 (2)2 (1)2 (1)2 The first four SALC equations are: B2g : 1 (2 pz( A) ) (2 pz( B) ) (2 pz(C ) ) (2 pz( D) ) (2 pz( E ) ) (2 pz( F ) ) 6 A2u : 1 (2 pz( A) ) (2 pz( B) ) (2 pz(C) ) (2 pz( D) ) (2 pz( E) ) (2 pz( F ) ) 6 1 2(2 pz( A) ) (2 pz( B) ) (2 pz(C ) ) 2(2 pz( D) ) (2 pz( E) ) (2 pz( F ) ) 12 1 E2u : 2(2 pz( A) ) (2 pz( B) ) (2 pz(C) ) 2(2 pz( D) ) (2 pz( E) ) (2 pz( F ) ) 12 E1g : Copyright © 2014 Pearson Education, Inc. 1 1 6 1 6 1 1 1 12 1 12 74 Chapter 5 Molecular Orbitals The remaining two SALCs ( E1g and E2u ) can be deduced by examination of the coefficients of the normalized equations. Coefficients of Normalized SALCs cA B2g A2u E1g 1 6 1 6 2 cB cC 1 1 6 1 6 1 6 1 6 1 cE 1 1 6 1 6 1 6 2 6 1 12 1 2 1 12 cF c A2 cB2 cC2 cD2 cE2 cF2 1 1 6 1 6 1 3 1 6 1 6 1 12 1 4 1 12 1 4 1 6 1 6 1 12 1 4 1 12 1 4 1 1 6 1 6 1 12 12 1 1 E1g 0 0 2 2 1 1 2 2 E2u 12 12 12 12 12 12 1 1 1 1 E2u 0 0 2 2 2 2 Sum of the squares for each 2pz wavefunction must total 1 for an identical contribution of each atomic orbital to the group orbitals 12 12 1 2 1 cD Squares of SALC Coefficients 0 1 3 0 1 1 1 6 6 1 1 6 6 1 1 12 12 1 1 4 4 1 1 12 12 1 1 4 4 1 6 1 6 1 3 1 1 1 0 1 3 0 Sum of the Squares = 1 for Normalization Requirement 1 1 1 1 1 1 The sum of the squares of the SALC coefficients for A and D equal 0 without any contribution from the second E1g and E2u SALCs; therefore, c A and cD are zero for these two SALCs. The 1 without contributions from the 2 1 second E1g and E2u SALCs. This suggests that c B2 , cC2 , c D2 , and c E2 equal for these two 4 1 SALCs, and that cC , c D , c E , and c F equal . The choice of signs in the table above are 2 those required for the SALCs to satisfy the symmetry requirements of the functions associated with the E1g and E2u representations, and to obtain the number of nodes expected (see sketches sum of the squares of the coefficients for B, C, E, and F equal below). The last two normalized SALCs are: 1 E1g : (2 pz( B) ) (2 pz(C ) ) (2 pz( E ) ) (2 pz( F ) ) 2 1 E2u : (2 pz( B) ) (2 pz(C ) ) (2 pz( E ) ) (2 pz( F ) ) 2 Copyright © 2014 Pearson Education, Inc. Chapter 5 Molecular Orbitals 75 The six group orbitals are sketched below, ranked by their relative energy. The number of nodes increases from zero to three with orbital energy. These orbitals, and a discussion of their energies, are in Section 13.4.4. B2g E2u E2u E1g E1g Energy A2u 5.30 5.31 a. Cl2+ has one fewer electron than Cl2, so the π* levels have three, rather than four, electrons. As a result, Cl2+ has a bond order of 1.5, and the bond is shorter and stronger than that of of Cl2 (189 pm, compared with 199 pm for Cl2). b. Cl4+ has such an elongated rectangular shape (194 pm by 294 pm) that it must be essentially a Cl2 and a Cl2+ side by side, with only a weak attraction between them through the π* orbitals. The Cl–Cl bond in Cl2 is 199 pm long; apparently, the weak side-to-side bond draws off some of the antibonding electron density, strengthening and shortening the other two shorter Cl–Cl bonds. 1+ a. F 5.32 F F B B F F F F 1– B F 1– 1+ F B F 1– F F 1+ b. The 1a2 orbital near the middle of the figure is the π-bonding orbital. c. The LUMO, 2a2, is the best orbital for accepting a lone pair. d. The 1a2 orbital is formed by adding all the pz orbitals together. The 2a2 orbital is formed by adding the B pz orbital and subtracting the three F pz orbitals. SF4 has C2v symmetry. Treating the four F atoms as simple spherical orbitals, the reducible representation can be found and reduced to F S Copyright © 2014 Pearson Education, Inc. F F F 76 Chapter 5 Molecular Orbitals = 2A1 + B1 + B2. Overall, the bonding orbitals can be dsp2 or d2sp, with the s and pz or dz2 orbital with A1 symmetry, px or dxy with B1 symmetry, and py or dyz with B2 symmetry. The pz or dz2 orbital remaining accounts for the lone pair. (The use of the trigonal bipyramidal hybrids dsp3 or d3sp include the lone pair as one of the five locations.) C2v A1 A2 B1 B2 5.33 C2 0 1 1 –1 –1 E 4 1 1 1 1 (xz) 2 1 –1 1 –1 (yz) 2 1 –1 –1 1 z, z2 x, xz y, yz A square pyramidal molecule has the reducible representation = E + 2A1 + B1. C4v E A1 B1 E 5 2 1 1 C2 1 –2 1 1 2C4 1 0 1 –1 2v 3 0 1 1 2d 1 0 1 –1 (x, y) (xz, yz) z, z2, x2+y2 x2–y2 There appear to be three possibilities for combining orbitals, depending on the details of their relative energies: dsp3 (px and py for E, s and pz for A1, dx2–y2 for B1), d2sp2 (substituting dz2 for pz), and d3sp (substituting dxz and dyz for px and py). Although dxz and dyz appear to work, they actually have their electron concentration between the B atoms, and therefore do not participate in bonding, so d3sp or d2sp2 fit better. 5.34 Square planar compounds have D4h symmetry. D4h Eu A1g B1g E 4 2 1 1 2C4 0 0 1 –1 C2 0 –2 1 1 C2 2 0 1 1 C2 0 0 1 –1 i 0 –2 1 1 2S4 0 0 1 –1 h 4 2 1 1 2v 2 0 1 1 2d 0 0 1 –1 = A1g + B1g + Eu 2 2 2 s, dz dx –y px, py dsp2 hybrids are the usual ones used for square planar compounds, although d2p2 is also possible. Since the dz2 orbital does not extend far in the xy plane, it is less likely to participate in bonding. Copyright © 2014 Pearson Education, Inc. (x,y) z2 2 2 x –y Chapter 5 Molecular Orbitals 5.35 a. 77 PCl5 has D3h symmetry. D3h E A1 A2 E 5 2 1 1 2C3 2 –1 1 1 3C2 1 0 1 –1 h 3 2 1 –1 2S3 0 –1 1 –1 3v 3 0 1 1 (x, y) (x2–y2, xy) z2 z = E + 2A1 + A2, so the hybrids are dsp3 or d3sp. b. This could also be analyzed separately for the axial and the equatorial positions. The pz and dz2 orbitals can bond to the axial chlorines (A1 + A2) and the s, px, and py orbitals or the s, dx2–y2, and dxy orbitals can bond to the equatorial chlorines (E). c. The dz2 orbital extends farther than the p orbitals, making the axial bonds a bit longer. 5.36 Ignoring Orbital Lobe Signs Including Orbital Lobe Signs D3h D3h D3h C2v C3v C3v C3h C1 1a2 2a2 1a2 1e Results should be similar to Figure 5.32. The energies of some of the orbitals in the middle of the diagram are similar, and the order may vary with different calculation methods. In addition, the locations of the nodes in degenerate orbitals (e and e) may vary depending on how the software assigns orientations of atomic orbitals. If nodes cut through atomic nuclei, 1e orbitals may have C2 symmetry, matching the symmetry of the first E group orbital shown in Figure 5.31. The table of orbital contributions for each of the orbitals should show the same orbitals as in Figure 5.32. There may be some differences in contributions with different calculation methods, but they should be minor. Assignments to px, py, and pz will also differ, depending on how the software defines orientations of orbitals. Semi-empirical calculation AM1 gives these as the major contributors to the specified orbitals: B F 3a1 4a1 1a2 1a2 2a2 2s 2s 2s 2s, 2py 2pz 2pz 2px 2pz 2pz Copyright © 2014 Pearson Education, Inc. 78 5.37 Chapter 5 Molecular Orbitals a. The shapes of the orbitals, generated using one of the simplest computational methods, Extended Hückel Theory, are shown below, with the most electronegative element shown at right in the heteronuclear cases. 1π 3 1π * N2 NO+ – CN CO b. In the 1π orbitals (bonding), the lobes are increasingly concentrated on the more electronegative atom as the difference in electronegativity between the two atoms increases. This effect is seen most significantly in CO, where the difference in electronegativity is the greatest. In the 1π* orbitals, the antibonding partners of the 1π orbitals, the effect is reversed, with the largest lobes now concentrated on the less electronegative atoms. The greatest effect is again shown in CO, with the lobes on carbon much Copyright © 2014 Pearson Education, Inc. Chapter 5 Molecular Orbitals 79 larger than those on oxygen. The 3 orbitals also show the influence of electronegativity, this time with the lobe extending to the left from the less electronegative atom being the largest, with CO once more showing the greatest effect. This can be viewed in part as a consequence of the 3 orbital being a better match for the energy of the less electronegative atom’s 2s orbital which, together with the 2pz orbital of the same atom, interacts with the 2pz orbital of the more electronegative atom (the atom shown on the right). c. The results vary greatly depending on the software used. The results using one approach, AM1, are shown below (numerical values are energies in electron volts). * π* LUMO HOMO π * CN – 14.7 0.13 –3.13 –5.10 –9.37 –28.0 CO 5.28 0.94 –13.31 –16.30 –22.00 –41.2 6.03 1.00 –14.32 –16.19 –21.43 –41.39 –4.42 –9.62 –26.13 –28.80 –35.80 –56.89 N2 + NO In this table, the energies decrease as the atomic numbers increase (with CO and N2 giving mixed results). There is considerable mixing of the orbitals, a phenomenon that may raise the energy of the (HOMO) orbital above the energy of the π orbitals– as is the case in each of these examples. 5.38 Among the trends that should be observed is the effect on the shapes of the π and π* orbitals (see orbitals of CO labeled as 1π and 1π* in Figure 5.13) as the difference in electronegativity between the atoms increases (this trend is also observed in Problem 37). For BF and BeNe, the lobes of the π orbitals should become increasingly concentrated on the more electronegative atoms, and the lobes of the π* orbitals should become increasingly concentrated on the less electronegative atoms (a pattern that has begun with CO, if the orbital shapes for CO are compared with those of the isoelectronic N2). An additional effect is that the size of the protruding orbital lobe of the less electronegative atom should increase as the difference in electronegativity between the atoms increases; this can be see in the 3 orbital of CO in Figure 5.13. Additional trends in the other molecular orbitals can also be noted. 5.39 In one bonding orbital, the H s orbitals have the same sign and add to the Be s orbital in the HOMO–1 orbital. Subtracting the Be s orbital results in the antibonding LUMO. The difference between the two H s orbitals added to the Be pz orbital results in the HOMO; subtracting the Be pz results in the LUMO+3 orbital. LUMO+1 and LUMO+2 are the Be px and py orbitals and are nonbonding (and degenerate) in BeH2. For an energy level diagram, see the solution to Exercise 5.8 in Appendix A. 5.40 BeF2 is similar to BeH2, with the addition of π and π* orbitals from the px and py orbitals, extending over all three atoms. The F px orbitals with opposite signs do not combine with the Be orbitals, and neither do the py orbitals; the px and py orbitals form the HOMO and HOMO+1 pair. Copyright © 2014 Pearson Education, Inc. 80 Chapter 5 Molecular Orbitals The answer to Problem 5.16 shows more details. 5.41 The azide orbitals are similar to the CO2 orbitals, with some differences in contributions from the atomic orbitals because the CO2 atomic orbitals do not have the identical energies – as the nitrogen atoms do. The two highest occupied orbitals of CO2, BeF2, and N3 all consist of px or py orbitals of the outer atoms with opposite signs, essentially nonbonding orbitals. The third orbital down has more s orbital contribution from the outer atoms than either of the other two; in those cases, the lower orbital energies of the atoms reduce that contribution. See also the solution to Exercise 5.7 in Appendix A. 5.42 One aspect of ozone’s molecular orbitals that should be noted is its π system. For reference, it is useful to compare the bonding π orbital that extends over all three atoms (the atomic orbitals that are involved are shown as molecular orbital 6 in the solution to Problem 5.19); this orbital is the lowest in energy of the 3-orbital bonding/nonbonding/antibonding set (orbitals 6, 9, and 10 in Problem 5.19) involving the 2p orbitals that are not involved in bonding. Another bonding/nonbonding/antibonding set can be seen in the molecular orbitals derived from 2s orbitals (orbitals 1, 2, and 3 in Problem 5.19). 5.43 a. Linear H H Cyclic H H H H In the linear arrangement, the molecular orbitals shown, from bottom to top, are bonding, nonbonding, and antibonding, with only the bonding orbital occupied. In the cyclic geometry, the lowest energy orbital is bonding, and the other two orbitals are degenerate, each with a node slicing through the center; again, only the lowest energy orbital is occupied. 5.44 b. Cyclic H3+ is slightly more stable than linear H3+, based on the energy of the lowest orbital in an AM1 calculation (–28.4 eV versus –26.7 eV). a. The full group theory treatment (D2h symmetry), shown in Section 8.5.1, uses the two bridging hydrogens as one set for group orbitals and the four terminal hydrogens as another set; these sets are shown in Figure 8.11. The representations for these sets can be Copyright © 2014 Pearson Education, Inc. Chapter 5 Molecular Orbitals 81 reduced as follows: The bridging hydrogens have = Ag + B3u. The boron s orbitals have = Ag + B1u. The px orbitals (in the plane of the bridging hydrogens) have = B2g + B3u. The pz orbitals (perpendicular to the plane of the bridging hydrogens) have = Ag + B1u. The boron Ag and B3u orbitals combine with the bridging hydrogen orbitals, resulting in two bonding and two antibonding orbitals. Electron pairs in each of the bonding orbitals result in two bonds holding the molecule together through hydrogen bridges. b. Examples of diborane molecular orbitals are in Figure 8.14. Copyright © 2014 Pearson Education, Inc. 82 Chapter 6 Acid-Base and Donor-Acceptor Chemistry CHAPTER 6: ACID-BASE AND DONOR-ACCEPTOR CHEMISTRY 6.1 Acid Base a. AlBr3 Br – b. HClO 4 CH 3CN Lewis, Brønsted-Lowry c. Ni2+ NH 3 Lewis d. ClF NH 3 Lewis e. SO 2 ClO3 – Lewis f. HF C3H 7COOH Lewis, Brønsted-Lowry Acid Base Definition XeO3 OH – Lewis 6.2 a. Definition Lewis This is not a Brønsted-Lowry reaction since the product connectivity is [XeO3 (OH)]– . b. Pt XeF4 Lewis c. H 2SeO 4 C2 H 5OH Lewis, Brønsted-Lowry d. Lewis [CH 3Hg(H 2O)]+ SH – This reaction likely occurs via the following steps: [CH 3Hg(H 2O)]+ SH – [CH 3Hg(SH)] + H 2O [CH 3Hg(SH)] + H 2O [CH 3HgS]– + H 3O + While a Brønsted-Lowry reaction occurs in the second step, the species listed in Problem 6.2d are involved in the initial Lewis acid-base reaction. 6.3 e. CH 3COOH (benzyl)3N Lewis, Brønsted-Lowry f. HCl SO 2 Lewis Al3+ is acidic: [Al(H2O)6]3+ + H2O [Al(H2O)5(OH)]2+ + H3O+ The hydronium ions react with the basic bicarbonate to form CO2: – H3O+ (aq) + HCO3 (aq) 2 H2O (l) + CO2 (g) – With pKa values of 5.0 for [Al(H2O)6]3+, 6.4 for H2CO3, and 2.0 for HSO4 , the pH is about 3, low enough to convert the bicarbonate to CO2. Copyright © 2014 Pearson Education, Inc. Chapter 6 Acid-Base and Donor-Acceptor Chemistry 83 – 6.4 An increase in conductivity suggests that ions are formed: BrF3 + KF 6.5 a. The ions are [BrF6] and [BrF4]+: 2 Cs[BrF6] + [BrF4][Sb2F11] b. [BrF6] : Oh (this complex features a stereochemically inactive nonbonding pair on the central Br atom. See A. R. Mahjoub, X. Zhang, K. Seppelt, Chem. Eur. J. 1995, 1, 261.) BrF4 + K+ – 3 BrF5 + 2 CsSbF6 – F [BrF4]+: C2v F c. – + F [BrF4] acts as a Lewis acid, accepting F . + Br F 6.6 2 H2SO4 H3SO4+ + HSO4– and 2 H3PO4 H4PO4+ + H2PO4– form enough ions to allow conductance in the pure acids. 6.7 Gas-phase basicity is defined as G for BH + (g) B(g) H + (g) , while proton affinity is H for the same reaction. Since G H T S , and S is undoubtedly positive for these reactions (where one mole of gaseous reactant is converted to two moles of gaseous products), it makes sense that the gas-phase basicities in Table 6.6 are less positive than the corresponding proton affinities. 6.8 These data suggest the following basicity ranking for these ketones: O O O < < H 3C CH3 H3CH2C CH2CH3 Ph Ph A convenient way to rationalize this basicity ranking is to examine the conjugate acids via resonance arguments. One resonance form features a positive charge on the carbonyl carbon (structure B, right). Since benzophenone can further delocalize this positive charge into its phenyl groups, the conjugate acid of benzophenone is the most stabilized of the three acids, leading to benzophenone being the strongest base. Since an ethyl group is H H slightly more electron-releasing than a methyl group, the O O conjugate acid of diethylketone is slightly more stabilized relative to the conjugate acid of acetone (structure B is more effectively stabilized in the conjugate acid of diethylketone R R relative to A, the conjugate acid of acetone). Acetone is R R B A consequently the weakest base among these ketones. 6.9 These data indicate that triphenylphosphine ( PPh 3) is more basic than triphenylamine ( NPh3) in the gas phase. On the sole basis of electronics, triphenylamine would be expected to be more basic by virtue of the higher electronegativity of N relative to P, leading to the N center being more electron rich than the P center. However, the origin of the observed gas-phase basicity ranking must be the varying abilities of these atoms to accommodate the tetrahedral geometries of the conjugate acids. The larger covalent radius of P relative to N results in longer P—C(phenyl) bonds than N—C(phenyl) bonds, resulting in less steric strain between the phenyl rings in Copyright © 2014 Pearson Education, Inc. 84 Chapter 6 Acid-Base and Donor-Acceptor Chemistry [HPPh3 ] relative to [HNPh3 ] . The conjugate acid [HPPh3 ] has less steric hindrance than [HNPh3 ] , and PPh 3 is more basic. 810 The data are graphed below. 750 Dimethylether 730 Ethanol 670 690 710 Methanol Water Gas-Phase Basicity (kJ mol-1) 770 790 Diethylether 650 6.10 ‐2.7 ‐2.5 ‐2.3 ‐2.1 ‐1.9 ‐1.7 ‐1.5 pKa of Conjugate Acid in Water a. The gas-phase and aqueous basicity data correlate poorly. The strongest base in aqueous solution ( H 2O ) is the weakest base in the gas phase. While basicity in water increases as Me 2O < Et 2O < MeOH < EtOH < H 2O , the basicity in the gas-phase increases as H 2O < MeOH < EtOH < Me 2O < Et 2O . b. The ethers are the strongest gas-phase bases within this series on the basis of the electronic impact of two electron-releasing alkyl groups bound to the oxygen atom compared to one group (ROH) or no alkyl groups ( H 2O ). The oxygen atoms of these ethers are relatively electron rich as a result. However, the ethers are the weakest bases in water of this series; their conjugate acids have fewer sites for hydrogen bonding with H 2O relative to the conjugate acids of alcohols and water. The poorer ability of the conjugate acids of the ethers to be solvated by water renders these ethers very poor bases in aqueous solution. c. As shown in the graph and the basicity rankings above, the “ethyl” molecule is more basic than the “methyl” molecule in both the gas-phase and in aqueous solution. This is undoubtedly an inductive effect; the more electron-releasing ethyl group renders Et 2O and EtOH more basic than Me 2O and MeOH , respectively, in both phases. d. H 2O is the strongest base in water of this series on the basis of the excellent ability of its conjugate acid ( H 3O + ) to be solvated by water via hydrogen bonding. It is the weakest gas-phase base within this series as a consequence of the relatively poor inductive effect of H compared to methyl and ethyl in increasing electron density at the oxygen atom. Copyright © 2014 Pearson Education, Inc. Chapter 6 Acid-Base and Donor-Acceptor Chemistry 6.11 85 This BF3 affinity trend is strongly correlated to the inductive ability of the groups bound to sulfur. These data suggest that the electron-releasing ability of substituents in sulfoxides ( R 2SO ) increases as R Ph Me n Bu cyclo-(CH 2 ) . A resonance argument can be employed to further rationalize the relatively low BF3 affinities of Ph 2SO and PhSOMe . The phenyl group permits delocalization of the formal positive charge at the oxygen in structure C (below), rendering this oxygen atom less Lewis basic than when alkyl groups—which cannot enable this attenuation of the positive formal charge—are present instead of phenyl groups. BF3 BF3 6.12 BF3 O O O S S S A B C Other resonance forms with further delocalization of positive charge into phenyl rings a. According the the authors, a reference acid should be a strong enough Lewis acid to react with most common bases; form 1:1 acid-base adducts; have spectroscopic characteristics that can be monitored to observe variations in the strength of Lewis bases when reactions are conducted; and should not undergo side reactions while acting as a Lewis acid. b. Lewis basicity towards the zinc(II) reference is governed significantly by the steric hindrance created at the nitrogen upon complexation. The bases quinuclidine and pyridine, both of which feature insignificant geometric changes at nitrogen upon binding, were found more Lewis basic than all primary, secondary, and tertiary amines examined with the zinc(II) reference. This is not the case when assessing Lewis basicity via BF3 affinities. For example, the BF3 affinity for pyridine (128.08 kJ mol ) is less than for some tertiary amines (for example, the BF3 affinity of Me3N is 139.53 kJ mol ). 6.13 c. The importance of steric hindrance is reflected in the trends observed. When acyclic amines are considered, the less hindered primary amines were generally the strongest Lewis bases. A clear trend was not observed with secondary and tertiary amines, but secondary amines were found stronger bases than tertiary amines when bases with the same alkyl group were examined. Among alicyclic amines, the trend is opposite in that quinuclidine (tertiary) was found stronger than the secondary amine piperidine. The authors state the relative order of the Lewis basicity for acyclic amines as primary > secondary > tertiary, but an inverted order for alicyclic amines ( tertiary > secondary ≈ primary (acyclic) ). a. The frustrated Lewis pair of the sterically encumbered P(t-C4 H 9 )3 , in combination with the highly Lewis acidic B(C6 F5 )3 , binds N 2O to give a PNNOB linkage. b. This complex has been characterized by single crystal X-ray crystallography (figure at right). Copyright © 2014 Pearson Education, Inc. tBu tBu tBu P N N O B C6F5 C6F5 C6 F5 86 Chapter 6 Acid-Base and Donor-Acceptor Chemistry 6.14 a. The reaction coordinate diagram is below. The van der Waals complex is hypothesized to be stabilized via significant pi-stacking between the aromatic rings of the borane and the secondary amine. In the transition state, the B—H bond has fully formed, and the proton is beginning to form a C—H bond with the previously aromatic ring of the amine. The formal positive charge of the hydrogenated intermediate (not shown on the ispo carbon of the amine) is stabilized by resonance. F F C 6F 5 F B C 6F 5 F F C6F5 H H F N H F B C 6F 5 H F tBu H H F F tBu H F B TS N F C6 F5 F 15.7 kcal/mol C 6F 5 van der Waals complex (12.3 kcal/mol) H F F H N H tBu H Hydrogenated Intermediate (4.8 kcal/mol) N H tBu Ph C 6F 5 B C6 F5 C 6F 5 H2 Frustrated Lewis Pair and Hydrogen (0.0 kcal/mol) [Ph(tBu)NH2][HB(C6F5)3] (-10.0 kcal/mol) [(C5H11)(tBu)NH2][HB(C6F5)3] (Relative energy unspecified) b. As proposed in the reference, the phenyl ring of the amine in the van der Waals complex rotates towards the boron atom. The hydrogen molecule is then “split” in this cavity formed by the borane and the phenyl ring of the secondary amine. The hydride bonds to the boron, and the proton binds to the para carbon of the amine phenyl ring. c. The activation barrier for hydrogenation of the phenyl ring of the secondary amine is higher than that for formation of [Ph(tBu)NH 2 ][HB(C6 F5 )3 ] , which is considered the resting state for this reaction. Therefore the rate of the hydrogenation reaction is exceedingly low in pentane at ambient temperature, but viable in refluxing toluene. The salt [i Pr2 NHPh][HB(C6 F5 )3 ] forms when the more basic i Pr2 NPh is employed. It is d. hypothesized that i Pr2 NPh is sufficiently basic to not allow formation of the areneborane van der Waals complex necessary to permit hydrogenation of the aromatic ring of the amine. 6.15 The N—O distance in the complex 2b is 129.6 pm, longer than the corresponding bond in free NO (115.1 pm). Bond lengthening is expected since NO likely uses its * orbital (its LUMO) to accept electron density from the phosphine. Population of an orbital that is antibonding with respect to the N—O bond will result in a longer bond distance. It is also noteworthy that the LUMO of NO features greater orbital contribution from the nitrogen atom relative to the oxygen atom, and the nitrogen atom binds to the phosphorous atom in 2b. a. Complex 2b has strong radical character at oxygen, and reacts with 1,4-cyclohexadiene via a hydrogen-atom abstraction pathway to afford benzene. Copyright © 2014 Pearson Education, Inc. Chapter 6 Acid-Base and Donor-Acceptor Chemistry (C6F5)2 B Resonance stabilzed radical (C6F5)2 B H H O N H H N P (Mes)2 P (Mes)2 (C6F5)2 B (C6F5)2 B H H O N N H 87 OH + OH + H H H P (Mes)2 P (Mes)2 Complex 2b also reacts with the stronger C—H bond of toluene, with hydrogen-atom abstraction followed by radical coupling to form an O—C bond. (C6F5)2 B O N (C6F5)2 B H H2 C P (Mes)2 N (C6F5)2 B O O N H2 C P (Mes)2 P (Mes)2 While the radical formed upon hydrogen-atom abstraction from the CH 3 group of toluene is resonance stabilized, C—O coupling involving a carbon atom in the ring would lead to a nonaromatic product (one such product is shown at right). Coupling at the primary carbon results in the retention of aromaticity in the product (left). N CH2 (C6F5)2 B (C6F5)2 B O vs. P (Mes)2 6.16 H2 C P (Mes)2 (C6F5)2 B b. + OH N N O P (Mes)2 a. The pK a for the water/HSCN pair is much smaller than for the water/HCN pair. The hydrogen bonding between water and HSCN is predicted as much stronger. b. The hydrogen bond donors that have the smallest pK a with nitriles are HClO 4 and HI. c. (NC)3CH has the smallest pK a with organic sulfides among the organic acids. d. Classification on the basis of estimated pK a values from the Slide Rule and the criteria in Section 6.5.1: Amines (medium strong), triphosphines (medium strong), sulfoxides (medium), ketones (medium/medium weak), and nitro compounds (medium weak). Copyright © 2014 Pearson Education, Inc. 88 Chapter 6 Acid-Base and Donor-Acceptor Chemistry 6.17 a. The structure has the Br atoms in a staggered structure, resulting in S6 symmetry. b. It is convenient to visualize this using tetrahedral, sp3 hybridized As. An sp3 orbital on each As points inward toward the benzene ring. If the hybrid lobes have opposite signs of their wave functions, they fit the symmetry of the π orbitals of the benzene ring to form bonding and antibonding molecular orbitals. The bonding interaction is shown. 6.18 As As a. The very high electronegativity of O in comparison with Al pulls the bonding pair very close to O. This increases the repulsion between the bonding pairs and causes the large angle. The steric bulk of the AlCl3 and PCl3 is not considered an important factor in dictating the large angle. b. The dative bond between OPCl3 and AlCl3 likely employs a nonbonding donor orbital of OPCl3 . The donation of electron density from a nonbonding orbital of OPCl3 would not be expected to significantly impact the P—O bond order. 6.19 6.20 a. The methyl groups in (CH3)3N—SO3 donate electrons to the nitrogen, making (CH3)3N a stronger Lewis base and strengthening and shortening the N—S bond. (CH3)3N—SO3 H3N—SO3 N—S 191.2 pm 195.7 pm N—S—O 100.1° 97.6° b. The greater concentration of electrons in the N—S bond of (CH3)3N—SO3 increases electron-electron (bp-bp) repulsions, opening up the N—S—O bond in comparison with H3N—SO3. a. The polarity of the Xe–F bonds concentrates electrons on the fluorine atoms, which act as the centers of Lewis basicity. As shown in the reference, which provides the structure of [Cd(XeF2)](BF4)2, the Xe–F–Cd bond is strongly bent at the fluorine and the geometry around xenon is nearly linear, with an F–Xe–F bond angle of 179.1°. b. The BF4– ion is smaller than AsF6–, and the charge per fluorine is also greater in BF4–, making BF4– the stronger Lewis base. In addition, the higher oxidation state of cadmium(II) in [Cd(XeF2)](BF4)2 enables stronger interaction with the fluorines in BF4– than occurs between the silver(I) ion and AsF6– in [Ag(XeF2)]AsF6. Copyright © 2014 Pearson Education, Inc. Chapter 6 Acid-Base and Donor-Acceptor Chemistry 6.21 89 An energy level diagram for NO- is below. The HOMO is the π* orbital. Bonding with H+ depends on which end of the π* orbital carries more electron density. Calculation shows slightly more electron density on N, making HNO the more likely (bent) molecule. This is consistent with the energy of the nitrogen 2p subshell (-13.18 eV) relative that of the oxygen 2p subshell (-15.85 eV); the nitrogen 2p orbitals should contribute more to these π* orbitals relative to the oxygen 2p orbitals. Problem 6.43 asks for calculation of the molecular orbitals of both HNO and HON to address this question from an alternate perspective. 2p 2p 2p 2p 2p 2p 2s 2s 2s 2s NO– N 6.22 O a. This is similar to the effects described in Section 6.4.2 for I2. Br2 forms charge-transfer complexes with donor solvents such as methanol. b. The 500 nm band (π* *) should shift to shorter wavelength (higher energy) because the difference in energy between the π* and * orbitals is greater in Br2•CH3OH than in Br2. – – – 6.23 AlF3 + F AlF4 . The Na+, AlF4 salt can dissolve in HF. When BF3 is added, it has a stronger – – attraction for the fluoride ions, with the reaction AlF4 + BF3 AlF3 + BF4 . 6.24 Soft metal ions do not combine with oxygen as strongly as hard metal ions, so reactions such as 2 HgO 2 Hg + O2(g) 2 Ag2O 2 CuO 2 Cu + O2(g) 4 Ag + O2(g) CuO + C Cu + CO(g) 2 CuO + C 2 Cu + CO2(g) are more easily carried out. Reduction of some of the softer metals can be carried out with relatively low temperatures (campfires); some think use of rocks containing ores could have led to accidental reduction to the metals and discovery of the smelting process. Harder metals such as iron require much higher temperatures for the reduction process. Copyright © 2014 Pearson Education, Inc. 90 Chapter 6 Acid-Base and Donor-Acceptor Chemistry 6.25 Hg is a much softer metal, and combines with the soft sulfide ion more strongly. Zinc and cadmium are more borderline metals, and combine with all the anions with more nearly equal facility. 6.26 When any of these salts vaporize, the vapor phase consists of molecules. When they are in the solid state, they are ionic, with some covalent properties. The liquid state is between the two, and can be made up of ions, covalent molecules, or something between these two extremes. If the liquid is molecular, vaporization should be easier (molecules in the liquid phase being converted to molecules in the vapor phase). If the liquid is mostly ionic, vaporization is more difficult. Using this criterion, the most ionic liquids should be ZnF2 and CdF2, and the most molecular liquids should be HgF2 and HgCl2. On a more general view, mercury, as the softest metal in the series, forms the more molecular compounds and zinc, as the hardest, forms the more ionic compounds. Cadmium forms the most molecular compound with the borderline bromide. 6.27 a. pyridine + BF3: H = E py E BF3 Cpy CBF3 = –[(1.17)(9.88) + (6.40)(1.62)] = –21.9 kcal/mol or –91.6 kJ/mol. This predicted value is roughly 13% less exothermic than the experimental value of -105 kJ/mol. pyridine + B(CH3)3: H = E py E B(CH 3 )3 Cpy CB(CH 3 )3 = –[(1.17)(6.14) + (6.40)(1.70)] = –18.1 kcal/mol, or –75.7 kJ/mol. This predicted value is more exothermic than the experimental values of -64 and -71.1 kJ/mol, respectively. 6.28 b. Fluorine is electron-withdrawing, and CH3 electron-releasing. Therefore, B in BF3 carries a greater positive charge and interacts more strongly with Lewis bases such as pyridine. c. The harder acid BF3 interacts more strongly with the borderline base pyridine. NH3 + BF3: H = E NH 3 E BF3 CNH 3 CBF3 = –[(1.36)(9.88) + (3.46)(1.62)] = –19.0 kcal/mol, or –79.5 kJ/mol NH3 + B(CH3)3: H = E NH 3 EB(CH 3 )3 CNH 3 CB(CH 3 )3 = [(1.36)(6.14) + (3.46)(1.70)] = –14.2 kcal/mol, or –59.4 kJ/mol The order is pyridine + BF3 > NH3 + BF3 > pyridine + B(CH3)3 > NH3 + B(CH3)3. The change from BF3 to B(CH3)3 is larger than the change from pyridine to NH3. Copyright © 2014 Pearson Education, Inc. Chapter 6 Acid-Base and Donor-Acceptor Chemistry 6.29 91 Absolute hardness parameters: I 15.81 10.7 9.3 7.8 BF3 NH3 C5H5N N(CH3)3 A –3.5 –5.6 –0.6 –4.8 9.7 8.2 5.0 6.3 The difference between the HOMO of pyridine (9.3) and the LUMO of BF3 (–3.5) is smaller than the other possible interaction, so this pair has the largest –H, in spite of the difference in hardness. By comparison with the nitrogen compounds, B(CH3)3 is expected to have an absolute hardness lower than that of BF3 , approximately 7.5—8. 6.30 CsI and LiF fit Basalo’s principle that ions of similar size and equal (but opposite) charge form the least soluble salts. CsF and LiI have ionic sizes that are very different, and they do not fit as well into an ionic lattice. In addition, CsI and LiF are soft—soft and hard—hard combinations, which combine better than the hard—soft LiI and soft—hard CsF. 6.31 The following reaction is unlikely since C is soft and O is hard. H3C CH3 H3C + CH3 O H H H O H The second reaction is is more likely. Adding the carbonyl oxygen makes C harder, and C in CH3 and the H atom are soft. O O H 3C C CH3 O H H 6.32 a. H 3C C + O CH3 H H Solubilities: MgSO4 > CaSO4 > SrSO4 > BaSO4 Electrostatic forces predict the reverse order due to cation sizes (the charge density of Mg 2+ is greater than that of the larger Ba 2+ ), but the larger cations fit better with the large sulfate anion in the crystals. Hydration of the cations is the strongest for Mg2+, weakest for Ba2+, agreeing with the solubility order. b. 6.33 Solubilities: PbCl2 > PbBr2 > PbI2 > PbS As a moderately soft cation, Pb2+ has stronger interactions with the softer anions – – – (hardness order: Cl > Br > I > S2–). In addition, hydration of the anions is largest for the chloride, smallest for sulfide, based on size. The order is Al–OC–W (see Figure 13.17). The oxygen of CO is the “harder” end of this molecule and bonds with the harder Al atom; carbon and W engage in a soft–soft interaction. Copyright © 2014 Pearson Education, Inc. 92 Chapter 6 Acid-Base and Donor-Acceptor Chemistry 6.34 a. TeH2 is the strongest acid, because Te is more electronegative than either Sn or Sb. Therefore, the hydrogen is more positive and acidic. b. NH3 is the strongest base because N is more electronegative than either P or Sb. c. (CH3)3N is the strongest base in the gas phase because the methyl groups contribute electron density to the nitrogen. In solution, the order is scrambled, probably due to solvation (Section 6.3.6). d. 4-Methylpyridine > pyridine > 2-methylpyridine. Again, the methyl group adds electron density to the N. However, with methyl in the 2 position, steric hindrance makes bonding to BMe3 more difficult. – 6.35 In general, oxide ion reacts with water to form hydroxide: O2– + H2O 2 OH . In B2O3, the small, hard B3+ holds on to the oxide ions strongly. As a result, B2O3 + 3 H2O 2B(OH)3 – H+ + H2BO3 , and the solution is very weakly acidic (pKa = 9.25). In Al2O3, the Al3+ ion – is larger and softer. It can form either [Al(OH)4] (acting as an acid) or [Al(H2O)6]3+ (acting as a base), depending on the other species in solution. Sc3+ is still larger and softer, so it combines better with water than with hydroxide ion. The result is the reaction Sc2O3 + 15 H2O – 2[Sc(H2O)6]3+ + 6 OH . 6.36 a. CaH2 + 2 H2O Ca2+ + 2 H2 + 2 OH Calcium has a lower electronegativity than hydrogen, so CaH2 is Ca2+(H–)2 and the hydride ions react with water. b. HBr + H2O H3O+ + Br Bromine is more electronegativite than hydrogen, so the hydrogen is strongly positive and is readily transferred to the lone pair of water. c. H2S + H2O – – – + H3O + HS Sulfur is slightly more electronegative than hydrogen, and the positive hydrogen in H2S can dissociate to a small extent. d. CH4 + H2O no reaction. The C–H bond is almost nonpolar; the interaction between the oxygen atom of water and this C–H bond required for hydrogen ion transfer is exceedingly weak. 6.37 BF3 > B(CH3)3 > B(C2H5)3 > B(C6H2(CH3)3)3 Alkyl groups are electron-donating and increase the electron density on B and reduce the attraction for the lone pair of NH3. The bulky mesityl groups render B(C6H2(CH3)3)3 less Lewis acidic on the basis of the steric hindrance introduced when these boron centers adopt the required tetrahedral geometry upon complexation of NH 3 . 6.38 a. CH3NH2 is a stronger base on the basis of the electron-releasing ability of the methyl group. b. Although 2-methylpyridine is the stronger base with smaller acid molecules, the methyl group interferes with adduct formation with trimethylboron (F-strain) and the pyridine-trimethylboron dative bond is stronger. Copyright © 2014 Pearson Education, Inc. Chapter 6 Acid-Base and Donor-Acceptor Chemistry 6.39 c. Trimethylboron forms a stronger adduct with ammonia because the three phenyl rings of triphenylboron cannot bend back readily to allow the boron to become tetrahedral (B-strain). a. With the acids listed in order of increasing acidity: H3AsO4 H2SO3 H2SO4 pKa (9-7n) 2 2 –5 pKa (8-5n) 3 3 –2 pKa (exp) 9.2 2.2 1.8 b. With the acids listed in order of increasing acidity: pKa (9-7n) pKa (8-5n) pKa (exp) 6.40 HClO 9 8 7.4 HClO2 2 3 2 HClO3 –5 –2 –1 93 HMnO4 –12 –7 –11 HClO4 –12 –7 –10 Dimethylamine acts as weak base in water, with a very small amount of OH– provided by – the reaction (CH3)2NH + H2O (CH3)2NH2+ + OH . Acetic acid is a stronger acid than water, so dimethylamine acts as a stronger base, and the reaction – (CH3)2NH + HOAc (CH3)2NH2+ + OAc goes to completion. 2-Butanone is a neutral solvent; there is no significant acid-base reaction with dimethylamine. 6.41 SbF5 in HF reacts to increase the H+ concentration and decrease H0: SbF5 + HF 6.42 – H+ + SbF6 . These ions then can react with alkenes. a. As the Lewis acids BF3 and BCl3 interact with NH3, the geometry around boron changes from planar to trigonal pyramidal; however, in accord with the LCP model the nonbonded F…F and Cl…Cl distances are nearly constant, suggesting that these atoms remain in contact with each other. Because these nonbonded distances remain essentially constant, the boron–halogen distance must increase as the distortion from trigonal geometry occurs. Because of the strength of the B–F bond, a consequence of the large electronegativity difference between B and F and the small size of the fluorine atom, more energy is required to distort BF3 from planarity than to similarly distort BCl3. (Calculations in support of this argument are presented in the reference.) The consequence of this energy requirement is that BF3 is a weaker Lewis base than BCl3 toward NH3. The article does not address the relative Lewis basicity of BBr3, but a similar argument could apply for this compound. b. This article does not consider the LCP model but focuses on ab initio calculations on the adducts X3B–NH3. These calculations show a higher B–N bond dissociation energy in the BCl3 adduct than in the BF3 adduct. Although many factors are involved, an important issue is that BCl3 has a lower energy LUMO that is able to interact more strongly with the donor orbital of NH3, giving a stronger covalent interaction (and stronger B–N bond) in Cl3B–NH3. Copyright © 2014 Pearson Education, Inc. 94 Chapter 6 Acid-Base and Donor-Acceptor Chemistry 6.43 a. The energy diagram and the orbitals are shown below. 2p 2p 2p 2p 2p 2p 2s 2s 2s 2s N NO– O The HOMO is the π*, with greater concentration of electron density on N. This orbital can overlap with the empty H+ 1s orbital, forming HNO, a bent molecule. b. Calculations predict that HNO is more stable than NOH, with the energy of the HNO HOMO lower than the energy of the NOH HOMO. The HOMO and HOMO-1 of HNO and NOH, respectively, two of the nine molecular orbitals of these species arising from the valence orbitals, are shown below. HNO NOH HOMO HOMO–1 Copyright © 2014 Pearson Education, Inc. Chapter 6 Acid-Base and Donor-Acceptor Chemistry 95 6.44 LUMO HOMO HOMO –1 HOMO –2 Br2 Br2–HOCH3 HOCH3 The interaction of the HOMO of the methanol and the LUMO of the Br2 results in the adduct LUMO and HOMO –2 orbitals, respectively. The geometry shown features the Br2 at approximately a trigonal angle (H—O—Br is 113° and C—O—Br is 106° in a PM3 calculation). 6.45 6.46 a. For orbitals of BF3 and NH3, see Figures 5.32 and 5.30, respectively. b. The B–N bonding molecular orbital, shown below, is polarized toward the more electronegative nitrogen; the matching antibonding orbital, which has a node between the boron and nitrogen atoms, is polarized toward the boron. a. The sketch below is a first approximation of the postulated interaction responsible for the halogen bond with acetylene. Copyright © 2014 Pearson Education, Inc. 96 Chapter 6 Acid-Base and Donor-Acceptor Chemistry The donor orbital of acetylene is a pi-bonding orbital b. The acceptor orbital of Br 2 is its * orbital A key orbital to look for is one in which the adjacent lobes of the pi-bonding orbital of acetylene, shown an a, and the * orbital of Br2 have merged into a larger lobe (occupied by an electron pair) that links the two molecules. This bonding orbital should have a matching antibonding orbital (empty) at higher energy. Copyright © 2014 Pearson Education, Inc. Chapter 7 The Crystalline Solid State 97 CHAPTER 7: THE CRYSTALLINE SOLID STATE 7.1 a. Oh b. e. The image in Figure 7.10, which has D3d symmetry, actually consists of three unit cells. For an image of a single unit cell, which has point group C2h, see page 556 in the Greenwood and Earnshaw reference in the “General References” section. D4h c. Oh d. Td 7.2 The unit-cell dimension is 2r, the volume is 8r3. Since this cell contains one molecule 4 / 3r 3 whose volume is 4/3 π r3, the fraction occupied is 0.524 52.4%. 8r 3 7.3 The unit-cell length for a primitive cubic cell is 2r. Using the Pythagorean theorem, we can calculate the face diagonal as 2r 2 2r 2 2.828r and the body diagonal as 2.828r 2 2r 2 3.464r . Each corner atom contributes r to this distance, so the diameter of the body center is 1.464 r and the radius is 0.732 r, 73.2% of the corner atom size. 7.4 a. A face-centered cubic cell is shown preceding Exercise 7.1 in Section 7.1.1. The cell 1 contains 6 × 1 = 3 atoms at the centers of the faces and 8 × 8 = 1 atom at the corners, a 2 total of 4 atoms. The total volume of these 4 atoms 4 radius of each atom (treated as a sphere). 4 3 16 3 r r , where r is the 3 3 As can be seen in the diagram, the diagonal of a face of the cube = 4r. By the Pythagorean theorem, the dimension (length of a diagonal 4r side) of the cube = . 2 2 3 4r 32r 3 . The volume of the cube, therefore, = 2 2 16 3 r volume of spheres 3 0.740 . The ratio = 3 32r volume of cube 2 Therefore, 74.0% of the volume of the cube is occupied by the spheres. b. In a body-centered cube, the unit cell contains 1 atom in the center of 1 the cube and 8 × 8 = 1 atom at the corners, a total of 2 atoms. The total volume of these 4 8 2 atoms 2 r 3 r 3 , where r is the radius of each atom. 3 3 Because atoms touch along the diagonal, the diagonal of the cube = 4r. Using the diagonal 4r Pythagorean theorem, it can be shown that the dimension of the cube = . 3 3 Copyright © 2014 Pearson Education, Inc. 98 Chapter 7 The Crystalline Solid State 3 4r 64r 3 Therefore, the volume of the cube = 3 3 3 8 3 r volume of spheres 3 0.680 or 68.0% The ratio = 3 64r volume of cube 3 3 7.5 In the table below, CaF2 is considered to have a fluoride ion in the body center of the overall unit cell and calcium ions in the body centers of the subunits (labeled “Internal”). Compound Corners NaCl cations NaCl anions 8 × 1/8 CsCl cations CsCl anions 8 × 1/8 CaF2 cations CaF2 anions 7.6 Edges Face Centers 6 × 1/2 12 × 1/4 Body Centers Internal Total Type 1×1 4 4 MX MX 1×1 1 1 MX MX 4 8 MX2 MX2 4×1 8 × 1/8 12 × 1/4 6 × 1/2 1×1 LiBr has a formula weight of 86.845, and the unit cell contains four cations and four anions (or four formula units per molecular unit cell). 10 –6 m3 86.845 g mol–1 3 –1 25.07 cm mol 2.507 10 –5 m3mol –1 –3 3 3.464 g cm cm 2.507 10 –5 m3mol –1 4 units 1.665 10 –28 m3 unit cell 6.022 10 23 units mol–1 unit cell 3 1.665 10 –28 m3 5.502 10 –10 m = unit cell length 2(r++ r–) = 5.502 × 10–10 m; r++ r– = 2.751 × 10–10 m = 275.1 pm The sum of the ionic radii from Appendix B.1 is 261 pm. 7.7 CsCl has 8 Cl– at the corners of the unit cell cube, with the Cs+ at the center. r+/ r– = 173/167 =1.04. CaF2 has the same structure in a single cube of F– ions, but only half the cubes contain Ca2+. r+/ r– = 126/119 = 1.06. Both should have coordination number = 12 based on the radius ratios. Copyright © 2014 Pearson Education, Inc. Chapter 7 The Crystalline Solid State 7.8 Figure 7.8 shows the zinc blende unit cell, which contains four S atoms (net) in an fcc lattice and four Zn atoms in the body centers of the alternate smaller cubes. The diagrams below are each a view of two layers of such a cell, with • indicating a Zn atom in this layer and a indicating a S atom in the layer below. The next pair of layers either above or below these has the opposite pattern, and the third repeats the original. The fcc lattice pattern can be seen for the Zn atoms (four corners and face center on the top face, four face centers in the middle later, and four corners and face center on the bottom face). Each S atom (and each Zn atom) has two nearest neighbors in the layer above and two in the layer below, in the arrangement for a tetrahedral hole. Extending the patterns below shows the S fcc lattice. • • • • • • • 7.9 99 • • • • • The graphite layers have essentially the same energy levels as benzene, but each level becomes a wider band because of the large number of atoms. This leads to the energy levels shown at right, with the bands coming from the lowest energy π orbitals filled and those coming from the highest energy π orbitals empty. The difference between the highest occupied and lowest unoccupied bands is small enough to allow conduction electrons to make the jump and the electrons and holes to move within the bands. Conduction perpendicular to the layers is smaller, because there are no direct orbitals connecting them. In polycrystalline graphite, the overall conductance is an average of the two. (The π orbitals of benzene are shown in Figure 13.22.) • • * C6H6 Graphite In diamond, each carbon atom has four bonds to its nearest neighbors. The gap between these filled orbitals and the corresponding antibonding orbitals is larger, effectively limiting conductance. The bonding in carbon nanotubes resembles that of graphite, and conductivity is expected. The conductivity of carbon nanotubes is a function of the diameter of the tubes, and spans the range associated with semiconductors to metallic conductors, up to approximately 1000 times the conductivity of copper (Section 8.6.1). 7.10 Solutions of alkali halides in water conduct electricity. This does not prove that they are ionic as solids, but is suggestive of ions in the solid state. Their high melting points are also suggestive of ionic structures, and the molten salts also conduct electricity. Perhaps the most conclusive evidence is from X-ray diffraction studies, in which these compounds show uniform cation—anion distances. If they were molecular species, the interatomic distances within a molecule should be smaller than the interatomic distances between molecules. 7.11 Hg(I) appears in compounds such as Hg22+ units. The 6s1 4f 14 5d 10 structure of Hg+ forms and * molecular orbitals from the s atomic orbitals, allowing the two s electrons to pair in the orbital for a diamagnetic unit. Copyright © 2014 Pearson Education, Inc. 100 7.12 Chapter 7 The Crystalline Solid State a. Forming anions from neutral atoms results in the addition of an electron. More electrons means a larger size, due to increasing electron—electron repulsion. By the same argument, forming cations from neutral atoms results in removal of an electron and smaller size. The cations have fewer electrons and less electron-electron repulsions. b. 7.13 The oxide ion is larger than the fluoride ion because its nuclear charge is smaller; the oxygen nucleus has less attraction for the electrons. Radius ratios for the alkali halides, using ionic radii for CN = 6 (Appendix B.1): Ions Radii (pm) F– Cl– Br– I– Li+ 79 0.66 0.47 0.43 0.38 Radii (pm) 119 167 182 206 Na+ 107 0.90 0.64 0.59 0.52 K+ 138 1.16 0.83 0.76 0.67 Rb+ 166 1.39 0.99 0.91 0.81 Cs+ 181 1.52 1.08 0.99 0.88 These salts all actually exhibit the sodium chloride lattice (CN = 6), and radius ratios between 0.414 and 0.732 are expected (Table 7.1). The seven boldface entries are the only structures correctly predicted on the basis of these data. These data incorrectly predict LiI to have CN = 4, NaF, KCl, KBr, RbCl, RbBr, RbI, CsBr, and CsI to have CN = 8, and KF, RbF, CsF, and CsCl to have CN = 12. 7.14 The interionic distances for salts of the same cation increase as Br – > Cl– > F – as expected on the basis of the increasing sizes of these monoanions with increasing atomic number. A comparison of these interionic distances with the sum of the corresponding ionic radii is interesting. Ionic radii for CN = 6 are used (Appendix B.1) in determining the sums (pm) below. Salt Interionic Distance LiF LiCl LiBr 201 257 275 Sum of Ionic Radii 198 246 261 Salt Interionic Distance NaF NaCl NaBr 231 281 298 Sum of Ionic Radii 226 274 289 Salt Interionic Distance AgF AgCl AgBr 246 277 288 Sum of Ionic Radii 248 296 311 It is interesting that only the silver salts exhibit shorter interionic distances than predicted by the sum of the ionic radii. As the softness of the anion increases, the absolute contraction of the interionic distance relative to the sum of the ionic radii increases (from only 2 pm (AgF) to 23 pm (AgBr)). This suggests an increasing covalent contribution in the interaction with Ag + as the halide softness increases from F – to Br – . Estimation of the interionic distance using ionic radii clearly introduces more error as the softness of the interaction towards Ag + increases. For the lithium and sodium salts, the interionic distances are all longer than the sum of the ionic radii. The agreement between the distances on a percentage basis is best for LiF (the interionic distance is 1.5% greater than the sum of the ionic radii) and worst for LiBr (the interionic distance is 5.4% greater than the sum of the ionic radii). The agreement between these values tracks qualitatively with the hard-soft compatibility of these ions. The relative differences in these Copyright © 2014 Pearson Education, Inc. Chapter 7 The Crystalline Solid State 101 distances changes more dramatically with decreasing halide softness for the harder Li+ (from 1.5% to 5.4%) than with the less hard Na + (2.2% to 3.1%). 7.15 ½ Cl2 (g) Cl (g) 239/2 = 119.5 (Energies in kJ/mol) – Cl (g) + e Cl (g) –EA Na (s) Na (g) 109 Na (g) Na+ (g) + e– 496 (=5.14 eV × 96.4853 kJ/mol•eV) – – Na+ (g) + Cl (g) NaCl (s) Total: Na (s) + ½ Cl2 (g) NaCl (s) –789 (Lattice energy calculated below) –413 EA = 348.5 kJ/mol (The value in Appendix B.3 is 349 kJ/mol.) U 23 NMZ Z – e2 30 6.022 10 1.74756 1 –1 2.3071 10 –28 J m 1 1 –12 r0 274 274 10 m 4 0 r0 – 789 kJ/mol 7.16 CaO has charges of 2+ and 2–, and radii of 114 and 126 pm, total distance of 240 pm. KF has charges of 1+ and 1–, and radii of 138 and 119 pm, total distance of 257 pm. The distance in CaO is 7% smaller and the charge factor is four times as large, both leading to stronger interionic attraction and contributing to the hardness of the crystal. MgO has charges of 2+ and 2–, and radii of 86 and 126 pm, total distance of 212 pm, with NaCl structure and Madelung constant = 1.75. CaF2 has charges of 2+ and 1–, and radii of 126 and 119 pm, total distance of 245 pm with the fluorite structure and Madelung constant = 2.52. The size difference and charges favor stronger MgO interionic attraction, enough to overcome the Madelung constant difference. 7.17 MgCl2 has a rutile structure, so it will be used for the NaCl2 calculation. The lattice energy for a rutile structure with combined radii of 253 pm (either NaCl2 or MgCl2) is: U U NMZ Z – e 2 1 r0 4 0 r0 2.307 10–28 Jm 1 30 6.022 1023 mol –1 2.385 2 –1 253 10 –12 m 253 –2,310 kJ mol –1 The lattice energy of MgCl2 (and the hypothetical NaCl2) is negative. If the Born—Haber cycles of MgCl2 and NaCl2, respectively, were compared, important contributing terms that would differ Copyright © 2014 Pearson Education, Inc. 102 Chapter 7 The Crystalline Solid State would be the H sub and ionization energy values. The magnitudes of the second ionization energies vary tremendously. Na+ Na2+ + e–, removing an electron from a core 2p orbital, requires much more energy than the corresponding Mg+ Mg2+ + e– reaction that removes the second 3s valence electron. The two metals differ by about 2,800 kJ/mol in these steps (all in kJ/mol): Na (s) Na(g) Na (g) Na+(g) + e– Na+ (g) Na2+(g) + e– Totals: Na (s) Na2+(g) + 2e– Mg (s) Mg (g) Mg (g) Mg+(g) + e– Mg+ (g) Mg2+(g) + e– Mg (s) Mg2+(g) + 2e– 107 495 4562 5164 147 738 1451 2336 Therefore, it is extremely unlikely that NaCl2 can be made. The large second ionization energy of Na does not permit the H f for hypothetical NaCl2 to be negative despite the predicted large and negative lattice energy of this salt. Repeating the process for a sodium chloride lattice (sum of radii = 274 pm) to compare The lattice energy and the energies necessary to form Na+ and Mg+: U NMZ Z – e 2 1 r0 4 0 r0 U 2.307 10–28 Jm 1 30 6.022 1023 mol–1 1.748 1 –1 274 10 –12 m 274 –789 kJ mol–1 Na (s) Na(g) Na (g) Na+(g) + e– Na (s) Na+(g) + e– 107 495 602 Mg (s) Mg (g) Mg (g) Mg+(g) + e– Mg (s) Mg+(g) + e– 147 738 885 Unlike with NaCl2, the lattice energy associated with forming MgCl nearly compensates for the energy required to generate Mg + . From this perspective, the formation of MgCl is more realistic. However, the ionic bonding in MgCl2 is so much stronger (nearly triple the lattice energy) that MgCl is a very unlikely product of the reaction of Mg and Cl2 . Indeed, MgCl2 is the thermodynamic sink in this system. 7.18 ½ Br2 (l) ½ Br2 (g) ½ Br2 (g) Br (g) Br (g) + e– Br– K (s) K (g) K (g) K++ e– K+ (g) + Br– (g) KBr (s) Total 14.9 (Energies in kJ/mol) 95.1 –324.7 79.0 418.8 –687.7 –404.6 kJ mol–1 for K (s) + ½ Br2 (l) KBr (s) Copyright © 2014 Pearson Education, Inc. Chapter 7 The Crystalline Solid State 103 For a sodium chloride lattice with total radii of 320 pm: U U NMZ Z – e 2 1 r0 4 0 r0 2.307 10–28 Jm 1 30 6.022 1023 mol–1 1.748 1 –1 320 10 –12 m 320 –687.7 kJ mol–1 7.19 ½ O2 (g) O (g) O (g) + 2 e– O2– Mg (s) Mg (g) Mg (g) Mg2++ 2e– Mg2++ O2– MgO (s) Total 247 (Energies in kJ/mol) 603 37 2188 –3934 –859 kJ mol–1 for Mg (s) + ½ O2 (g) MgO (s) For a sodium chloride lattice with total radii of 212 pm: 7.20 U NMZ Z – e 2 1 r0 4 0 r0 U 6.022 10 23 mol –1 1.748 2 –2 30 –1 2.307 10 –28 Jm 1 –3934 kJ mol –12 212 212 10 m PbS has total radii of 303 pm. For the NaCl lattice: U NMZ Z – e 2 1 r0 4 0 r0 U 6.022 10 23 mol –1 1.748 2 –2 30 2.307 10 –28 Jm 1 –12 303 30310 m –2888 kJ mol–1 1/8 S8 (s) S2– (g) Pb (s) Pb (g) Pb (g) Pb+ (g) + e– Pb+ (g) Pb2+ (g) + e– Pb2+ (g) + S2– (g) PbS (s) 535 196 716 1450 U Total –98 kJ mol–1 for Pb (s) + 1/8 S8 (s) PbS (s) (Energies in kJ/mol) U = –2995 kJ mol–1, roughly 3.5% more exothermic than -2888 kJ mol–1. Copyright © 2014 Pearson Education, Inc. 104 7.21 Chapter 7 The Crystalline Solid State In ZnO or TiO, additional Zn or Ti would have two more electrons than the metallic ions. As a result, any nonstoichiometry in the direction of excess Zn or Ti would supply extra electrons, making an n-type semiconductor. In Cu2S, CuI, or ZnO, excess S, I, or O would have fewer electrons than the corresponding ions. Therefore, the result of excess nonmetals in the lattice would be a p-type semiconductor. 7.22 Vibrational motions of the atoms in the lattice become at least partly synchronized, with positive centers moving closer together. This concentration of positive charge can attract electrons, allowing two electrons to be closer to each other than would usually be the case. When the vibrations are synchronized, this attraction can ripple through the material, helping the electrons move. Apparently the whole system acts as if it is at ground state energies, so no net change in energy is needed to keep the process going indefinitely. 7.23 The general reaction is Na2Z + Ca2+ (aq) CaZ + 2Na+ (aq), where Na2Z is the original zeolite with sodium ions providing the positive charge. When hard water, containing Ca2+ or Mg2+, passes through the zeolite, the ions exchange, leaving only Na+ in the softened water. The zeolite can be regenerated by flushing with concentrated brine. The large Na+ concentration reverses the reaction above. 7.24 The ion C24– should have the following molecular orbitals (see Figure 5.7): 2p 2p 2p distortion 2p 2s 2s Distortion could result in removal of the degeneracy of the πu and πg* orbitals, giving a diamagnetic ion. 7.25 Gallium nitride has a larger band gap than gallium arsenide (continuing the trend in which gallium phosphide has a larger band gap than gallium arsenide; see Table 7.3) and emits higher energy light. Gallium nitride has grown rapidly in importance; it is used in high-energy lasers, LEDs that emit light in the high-energy part of the visible spectrum as well as at lower energies, and a variety of other electronics applications. Copyright © 2014 Pearson Education, Inc. Chapter 7 The Crystalline Solid State 105 7.26 The smaller the size of the quantum dots, the greater the separation between energy levels within the dots, and the higher the energy the photoluminescence. Consequently, the largest dots would produce the lowest energy emission bands. 7.27 First appearing in the literature in 2001, articles on medical applications of quantum dots and related topics have been more than doubling in number every two years. 7.28 The spectral effect of the L-cysteine/ Cd 2+ ratio on the absorbance of CdS quantum dots is shown in Figure 1 of this reference. The molar ratio 4:2 causes the highest blue shift of the absorption edge (363 nm) and affords the smallest CdS quantum dots. These QDs absorb UV-Vis radiation of highest energy among those studied in this paper. 7.29 These quantum dots were prepared via first generating a solution of CdCl2 2.5 H 2O and thiourea in a small volume of ultrapure water. An aqueous solution of 3-mercaptopropionic acid was added, followed by 1 M NaOH, until the pH = 10. The molar ratio of Cd 2+ /thiourea/3mercaptopropionic acid was typically 1/1.7/2.3. The solution was saturated with N 2 and transferred to an autoclave. The mixture was subjected to 100 ° C for specified times before being lowered to ambient temperature. The pressure above the sealed solution increased during the reaction. A reaction scheme is provided here: NaO2C CdCl2 HSCH2CH2CO2H NaO2C S NaO2C H2N S S S H2N (Chemicals in autoclave) S CO2Na S CO2Na O in H2O at pH=10 100oC CdS S NH2 CO2Na S CO2Na NH2 (Species in autoclave after the reaction) The diameters of these QDs depend on the reaction time, with longer times leading to larger QDs. Diameters were calculated via two mathematical approaches using band gap and UV-Vis data, respectively. Reaction Time (min) 45 60 90 120 180 Diameters (nm) from Band Gap Data 2.5(4) 2.7(4) 3.1(4) 3.5(4) 3.9(4) Diameters (nm) from UV-Vis Data 2.3(4) 2.5(4) 2.9(4) 3.1(4) 3.7(4) As impressively shown in Figures 2b and 2c of the reference, the smaller the QDs (the shorter the reaction time), the higher the energy of the emitted radiation when these QDs are excited with UV radiation. The emission peak maxima vary from 510 nm (smallest QDs, 45 min reaction time) to 650 mn (largest QDs, 180 min reaction time). Copyright © 2014 Pearson Education, Inc. 106 Chapter 7 The Crystalline Solid State 7.30 These authors employed two spectroscopic methods to assess CdSe and CdTe quantum dot size. UV-Vis spectra were obtained throughout the digestive ripening stage to monitor particle size. Fluorescence spectra were subsequently acquired to confirm the UV-Vis results. The UV-Vis spectra acquired after 4 and 8 hours of digestive ripening are in Figure 3 of the reference. The corresponding fluorescence spectra are in Figure 4. The lower energy fluorescence max of the CdTe QDs (572 nm) relative to that of the CdSe QDs (530 nm) indicates the larger size of the CdTe QDs (Figure 4). A fundamental concept for QDs, that decreasing size leads to higherenergy emitted radiation, is applied to make the size assessment. The relative sizes were further examined by transmission electron microscopy (TEM) experiments that indicated the average particle sizes of the CdSe and CdTe QDs to be 4.0 and 4.5 nm, respectively. 7.31 The metallic nature of BaGe3 was probed by measuring the temperature dependence of its electrical resistivity; the metallic nature was established since the resistivity decreases with decreasing temperature. Zero resistivity was observed below 4.0 K, suggesting that BaGe3 is a superconductor with this critical temperature. The Meissner effect was also observed at 4.0 K, providing further support of this critical temperature. The conduction bands of BaGe3 are mainly composed of Ge 4p orbitals with some contribution from Ba orbitals. The states near the Fermi level are primarily made up of Ge 4p and Ba 5d orbitals. It is interesting that the Ge 4 px , 4 p y , and 4 pz orbitals contribute to different relative extents to the conduction bands near the Fermi level, with large contributions of Ge 4 px and 4 p y orbitals near the Fermi level of BaGe3 . 7.32 Figure 3 of the reference provides a plot of Tc versus a-axis lattice constants. For the LnFeAsO- based materials, the highest Tc is associated with an a-axis lattice constant of approximately 3.91 Å. The Tc values reach a maximum when the lanthanide element lies roughly in the middle of the lanthanide series (Sm, Gd), and decreases as the masses of the lanthanide atoms increase and decrease, respectively, from these elements, with La, Ce, (the lightest lanthanides) and Er (a heavier lanthanide) lying at the extremes of the Tc curve for LnFeAsO-based materials. In this regard, a periodic trend is roughly observed. The authors suggest an upper limit on the a-axis constant (~4.0 Å) for the possibility of superconductivity. The lack of superconductivity of the two perovskite-based As materials ( (Sr3Sc 2O5 )(Fe 2As2 ) and (Ba 3Sc 2O5 )(Fe 2As2 ) ) is attributed to their very long a-axis constants of roughly 4.07 and 4.13 Å, respectively. 7.33 a. Si4O128– b. Si8O2416– c. [Si6O1710–]n Copyright © 2014 Pearson Education, Inc. Chapter 8 Chemistry of the Main Group Elements 107 CHAPTER 8: CHEMISTRY OF THE MAIN GROUP ELEMENTS 8.1 a. H2 H2+ 74.2 pm 106 pm 436 kJ/mol 255 kJ/mol These values are consistent with the molecular orbital descriptions. H2 has two electrons in the bonding orbital, H2+ has only one. The net attraction between the bonding electron and the nuclei are weaker in H2+ relative to that between the two bonding electrons and the nuclei in H2.Therefore, the bond distance is longer in H2+. b. 8.2 H3+ was described in Problems 5.21 and 5.43. It has a single pair of electrons in a 2-electron, 3-center bond, with bond orders of 1/3. He2+ has 2 electrons in the bonding orbital and 1 in the antibonding * orbital, with a bond order of ½. HeH+ has a bond order of one, but with a poorer match between the energy levels of the two atoms (–13.6 eV for H, –24.5 for He). 1s* 1s 1s* 1s 1s 1s 1s 1s He He2+ He+ HHe+ He H+ 8.3 If the isoelectronic series could be continued, the ion CsF4+ would be expected to have a similar – structure to XeF4 and IF4 , with two lone pairs in trans positions. However, a variety of factors – make such a species truly unlikely. For example, while the bonds in XeF4 and IF4 are best described as polar covalent, the tremendous difference in electronegativity between Cs and F suggest that ionic bonding would be necessary in CsF4+. However, complexes containing Cs(V) are unknown. 8.4 The equilibrium constants for formation of alkali metal ion complexes with cryptands are listed in Figure 8.7, with the formation constant for the sodium complex with cryptand [2.2.1] larger than for either the lithium or potassium complexes. A similar trend is found for the alkaline earths, with the maximum at strontium. Apparently the optimum size for an ion fitting in the cryptand is larger than sodium (atomic radius: 107 pm), but smaller than barium (149 pm) or potassium (138 pm), leaving strontium (132 pm) as the closest fit. 8.5 The diagram at the right shows the primary interactions forming 2p molecular orbitals. The other orbitals on the fluorine atoms 2s form lone-pair orbitals and π orbitals. The interactions are similar to those in carbon dioxide (Figure 5.25), but the poorer energy match between the valence orbitals leads to significantly less covalent character in the resulting bonds of BeF2. Be Copyright © 2014 Pearson Education, Inc. 2p F Be F F F 108 8.6 Chapter 8 Chemistry of the Main Group Elements A combination of the in-plane p orbitals of Cl and the s and in-plane p orbitals of Be form orbitals that link the Be and Cl atoms in 3-atom, 2-electron bonds. The atomic orbital interactions that result in bonding molecular orbitals involving these bridging chlorine atoms are shown below. The antibonding orbitals also formed by these interactions, along with other orbitals with significant contribution from the terminal chlorine atoms, are not shown. pz pz px Cl Be Be Cl pz px s x z pz px px s The same orbitals, shown as calculated using extended Hückel theory software, are below. Orbitals from the terminal chlorine atoms are included in these molecular orbitals. Copyright © 2014 Pearson Education, Inc. Chapter 8 Chemistry of the Main Group Elements 8.7 The much greater difference in orbital energies of B (–8.3 and –14.0 eV) and F (–18.7 and –40.2 eV) makes the BF bond weaker than that of CO. The s orbital of B and the p orbital of F can interact, but the p-p combination cannot interact significantly because of the large difference in p orbital energies. This mismatch results in weak single bonding, rather than the triple bond of CO. 109 2p 2s 2p 2s B BF F R 8.8 a. Evidence cited for a double bond includes an X-ray crystal structure that shows a very short boron–boron distance (156 pm) that is considerably shorter than compounds having boron—boron single bonds and comparable to the bond distance in compounds with double bonds between boron atoms. Density functional theory computations also showed a HOMO with boron—boron π bonding. N R H R B R R N B N H R R N R = isopropyl R R b. The product in this case was a silicon compound having a structure similar to the boron compound shown above, but lacking hydrogen atoms; nevertheless, it has bent geometry at the silicon atoms, leading to the zigzag structure shown at right. An X-ray crystal structure similarly shows a silicon–silicon distance consistent with a double bond, and density functional theory yields a HOMO with π bonding between the silicon atoms. N N R R Si R Si R N R N R 8.9 The combination of orbitals forming the bridging orbitals in Al2(CH3)6 is similar to those of diborane (Section 8.5.1). The difference is that the CH3 group has a p orbital (or sp3 hybrid orbital) available for bonding to the aluminum atoms, with the same symmetry possibilities as those of H atoms in B2H6. Copyright © 2014 Pearson Education, Inc. R 110 8.10 Chapter 8 Chemistry of the Main Group Elements D2h (pz) (px) (1s) Ag B2g B1u B3u E 2 2 2 1 1 1 1 C2(z) 2 –2 0 1 –1 1 –1 C2(y) 0 0 0 1 1 –1 –1 C2(x) 0 0 2 1 –1 –1 1 (xy) 0 0 2 1 –1 –1 1 i 0 0 0 1 1 –1 –1 (xz) 2 2 2 1 1 1 1 (yz) 2 –2 0 1 –1 1 –1 z2 xz xy x Reduction of the representations gives the following: Boron orbitals: a. (pz) = Ag + B1u b. (px) = B2g + B3u Hydrogen orbitals: 8.11 c. (1s) = Ag + B3u d. Treating each group orbital as a single orbital, the orbitals below have the indicated symmetries. D2h E C2(z) C2(z) C2(z) i (xy) (xz) (yz) Ag (pz) 1 1 1 1 1 1 1 1 B1u (pz) 1 1 –1 –1 –1 –1 1 1 B2g(px) 1 –1 1 –1 1 –1 1 –1 B3u (px) 1 –1 –1 1 –1 1 1 –1 Ag (s) 1 1 1 1 1 1 1 1 B3u 1 –1 –1 1 –1 1 1 –1 Carbon has two lone pairs in C(PPh3)2, with VSEPR predicting a tetrahedral electron group geometry. The bulky phenyl groups limit the impact of the nonbonding pairs and force a larger angle. Copyright © 2014 Pearson Education, Inc. Chapter 8 Chemistry of the Main Group Elements 8.12 111 The cations of CaC2, CeC2, and YC2 are not isoelectronic. While Ca 2+ has no valence electrons, the unusual 2+ group 3 ion Y 2+ and the lanthanide Ce 2+ contain 1 and 2 valence electrons, respectively. It has been suggested (Greenwood and Earnshaw, Chemistry of the Elements, 2nd ed., p. 299) that there is some transfer of this valence electron density to the π* orbitals of the dicarbide (or acetylide) ion, resulting in a longer bond. 8.13 Radioactive decay obeys a first order kinetic equation: x dx –kx; ln –kt dy x 0 The relationship of the half-life (the time at which x = ½ xo) and the rate constant is x 1 ln ln –kt1/ 2 –0.693 x 0 2 k 0.693 t1/ 2 0.693 1.2110 –4 y –1 t1/ 2 x ln ln.56 –1.21 10 –4 t x 0 k t 4.8 10 3 y 8.14 Ih symmetry includes 12 C5 axes, 12 C52 axes, 20 C3 axes, 15 C2 axes, an inversion center, 12 S10 axes, 20 S6 axes, and 15 mirror planes. One of the C5 axes and an S10 axis can be seen in the middle of the second fullerene figure (end view) in Figure 8.18. There are six of each of these, each representing two rotation axes, C5 and C54, for a total of 12. The same axes fit the 12 S52 axes. If the structure is rotated to line up two hexagons surrounded by alternating pentagons and hexagons, the 20 C3 and 20 S6 axes can be seen (one in the center, three in the first group around that hexagon, and three adjacent pairs in the next group, doubled because of C3 and C32). The C2 axes pass through bonds shared by two hexagons, with pentagons at each end. These original hexagons each have two more, making five, the next ring out has eight, and the perimeter (seen edge on) has two for a total of 20 C2. There are five mirror planes through the center of the pentagons surrounded by five hexagons (see the end view of Figure 8.18 again), and there are three sets of these for a total of 15. (All this is best seen using a model!) 8.15 a. Td b. Ih c. D5h d. zigzag: C2h armchair: D2h Copyright © 2014 Pearson Education, Inc. 112 Chapter 8 Chemistry of the Main Group Elements 8.16 Graphane, obtained from the hydrogenation of graphene, was first reported in 2009 (see J. Agbenyega, Mater. Today, 2009, 12, 13), and research interest has grown rapidly. 8.17 Printing on transparencies is suggested for this exercise. Rolling up of the diagram should show that more than one chiral form is possible. 8.18 Readers are encouraged to search for the most recent references in this promising area. At this writing (February 2013) the reference has been cited in 176 other articles in scientific journals. 8.19 The reference provides absorption spectra and colors of solutions containing mixtures of armchair-enriched single-walled carbon nanotubes with different diameters, ranging from 0.83 nm to 1.5 nm. The strongest absorption band for the nanotubes shifted to longer wavelength (lower energy) as the diameter increased, with the smallest absorbing at 407 nm and the largest at 785 nm, and the observed colors (complementary to the colors of the strongest absorption bands) changed accordingly. The observations were consistent with the energy levels of the nanotubes becoming closer as the diameters of the tubes increased. A similar size effect is observed for quantum dots (Section 7.3.2). 8.20 The increased stability of 2+ oxidation states as compared to 4+ is an example of the “inert pair” effect (see Greenwood and Earnshaw, Chemistry of the Elements, 2nd ed., pp. 226, 227, 374). In general, the ionization energy decreases going down a column of the periodic table, because of greater shielding by the inner electrons. In this family, removal of the second electron is fairly easy, as it is the first in the higher-energy p orbitals. However, the next two electrons to be removed are the s electrons, and they are not as thoroughly shielded in the ions. The effect is larger for the three lower members of the group because the d10 electrons are added between the s and p electrons. The lower electronegativity of C and Si make them more likely to form covalent bonds than ions. 8.21 a. H2ISiSiH2I has a C2 axis perpendicular to the plane of the diagram as drawn, a perpendicular mirror plane (the plane of the diagram), and an inversion center, C2h. b. The Si—H stretches have the reducible representation shown below, which reduces to = Ag +Bg +Au+Bu. Au and Bu are IR-active. C2 C2h E i h 4 0 0 0 1 1 1 1 Ag Rz 1 –1 1 –1 R x, R y Bg 1 1 –1 –1 Au z 1 –1 –1 1 x, y Bu 8.22 P4 (g) 2 P2 (g), H = 217 kJ mol–1 P4 has six P–P bonds, so six bonds are broken and two triple bonds are formed. H = bond dissociation energy (reactant) – bond dissociation energy (product) 217 = 6 × 200 – 2(bond dissociation energy of PP) Bond dissociation energy of PP = 492 kJ mol–1 The pπ orbitals in P2 do not overlap as effectively as those in N2 on the basis of the larger atomic radius of P relative to N, resulting in weaker π bonds in P2. Copyright © 2014 Pearson Education, Inc. Chapter 8 Chemistry of the Main Group Elements 8.23 a. b. c. 8.24 – N3 has molecular orbitals similar to those of CO2 (Section 5.4.2), with two occupied orbitals and two occupied π orbitals for a total of four bonds. In contrast to the situation in CO2 , the atomic orbitals involved in the – bonding in N3 have equal energies. 113 * n – Because the HOMOs of N3 (at right) are primarily composed of p orbitals of the terminal nitrogen atoms, H+ bonds at an angle to the N=N=N axis. The angle is 114°, larger than the 90° predicted by simple bonding to a p orbital. The H–N–N angle is 114º. This observation suggests that the first resonance structure (with an N(sp)—H bond) contributes less to the electronic ground state of HN3 relative to the second structure (with an N(sp3)—H bond. A decreased terminal N—N distance relative to the central N—N distance is consistent with these contributions. H 1+ 1+ N N 1– H N 1+ N 2– N N The electronegativity of H (2.300) is less than that of N (3.066). Hydrazine has two hydrogen atoms and an equally electronegative NH2 group bound to each nitrogen atom, while ammonia has three hydrogens bound to N. In this way, ammonia features three polar covalent bonds enriching the electron density at nitrogen, while hydrazine features two polar covalent N—H bonds and one nonpolar covalent bond (to the NH2 group). The N of ammonia is therefore more electron-rich, and hydrazine is the weaker aqueous base. Another factor is the potential for hydrogen bonding with water in the resulting conjugate acids. An ammonium ion (NH4+) features four excellent hydrogen bond donors per N atom, while hydrazinium ion features just three hydrogen atoms bound to a positively charged N atom. The enhanced opportunities for NH4+ solvation relative to N2H5+ contribute to NH3 being a stronger base. 8.25 The larger central atoms feature longer bond lengths, leading to lower concentrations of bonding electron density close to the Group 15 atom, reducing bonding pair-bonding pair repulsion. The lone pair subsequently exerts increasing influence down the column, forcing smaller angles as the atomic radius of the central atom increases. This decrease in bonding pair-bonding pair repulsion can also be explained using electronegativity arguments, with the most electronegative nitrogen atom drawing a larger share of the bonding electron density towards it relative to the other group 15 elements that progressively become less electronegative down the column. 8.26 One p orbital of each oxygen is used for the bond to nitrogen, and the p orbital perpendicular to the plane of the molecule is used for a π orbital. The remaining p orbital is in the plane of the molecule, so addition of a proton to this lone pair leaves the entire molecule planar. 8.27 From Table 8.10 N2O Cv N2O4 D2h – NO2 C2v HNO2 Cs NO N2O5 – NO3 HNO3 Cv D2h D3h Cs NO2 C2v NO+ Cv N2O22– C2h Copyright © 2014 Pearson Education, Inc. N2O3 Cs NO2+ Dh NO43– Td 114 8.28 Chapter 8 Chemistry of the Main Group Elements a. b. Computational work suggests that the electronic ground state of cis- N 2 F4 is 1.4 kcal/mol lower than that of trans- N 2 F4 at 298 K. Two trans-cis isomerization mechanisms are proposed. The simpler mechanistic possibility involves rotation about the N=N bond. A lower limit of 59.6 kcal/mol was calculated for this isomerization mechanism. The second possibility proceeds via a planar transition state that resembles [N 2 F]F , with relatively short and long N—F bonds, and a nitrogen–nitrogen triple bond, respectively. The transformation of nitrogen–nitrogen bonding electron density to nonbonding electron density at nitrogen completes the isomerization. An activation barrier of 68.7 kcal/mol was determined for this mechanism. The activation barrier for isomerization via N=N rotation is 9.1 kcal/mol less than the proposed barrier for the [N 2 F]F mechanism. c. As a strong Lewis acid, SbF5 can engage in a donor-acceptor interaction with a nonbonding pair on a fluorine atom (left). The structure on the right, where the F atom has been completely transferred, resulting in [N 2 F][SbF6 ] , is the limiting result of this interaction. F F F N Sb F F F F F N N F Sb F N F F F F While fluoride transfer is not envisioned to occur when SbF5 assists in the isomerization, this possibility is instructive since it suggests that the interaction on the left is expected to lengthen one N—F bond while shortening the other N—F bond and the N—N bond. These changes are predicted to lower the activation barrier of the donor-acceptor complex (left) towards formation of the planar isomerization transition state that resembles [N 2 F]F , with a relatively short and long N—F bonds, and a nitrogen–nitrogen triple bond, respectively. 8.29 The balanced reduction half-reactions, for acidic and basic conditions, respectively, follow. Acidic: H 3PO4 (aq) + 2 H + (aq) + 2e – H 3PO3 (aq) + H 2O(l) H 3PO3 (aq) + 2 H + (aq) + 2e– H 3PO2 (aq) + H 2O(l) 4 H 3PO 2 (aq) + 4 H + (aq) + 4e– P4 + 8 H 2O(l) P4 + 8 H + (aq) + 8e– 2 P2 H 4 P2 H 4 2 H +(aq) 2e– 2 PH 3 Basic: PO43– (aq) + 2 H 2O(l) + 2e – HPO32– (aq) + 3 OH – (aq) HPO32– (aq) + 2 H 2O(l) + 2e – H 2 PO2– (aq) + 3 OH – (aq) Copyright © 2014 Pearson Education, Inc. Chapter 8 Chemistry of the Main Group Elements 115 4 H 2 PO 2– (aq) + 4e – P4 + 8 OH – (aq) P4 + 12 H 2O(l) + 12e– 4 PH 3 + 12 OH – (aq) The Frost diagram for phosphorus in acidic solution features points at (5, –1.915; H 3PO4 ), (3, –1.363; H 3PO3 ), (1, –0.365; H 3PO2 ), (0, 0; P4 ), (–2, 0.200; P2 H 4 ), and (–3, 0.206; PH 3). The Frost diagram for phosphorus in basic solution features points at (5, –7.43; PO 43– ), (3, –5.19; HPO32– ), (1, –2.05; H 2 PO 2– ), (0, 0; P4 ), (–3, 2.67; PH 3). 8.30 The bonding between O2 units is proposed to involve interactions between the singly occupied π* orbitals of neighboring molecules. Such interactions would generate both bonding and antibonding molecular orbitals, with the π* Copyright © 2014 Pearson Education, Inc. π* 116 Chapter 8 Chemistry of the Main Group Elements originally unpaired π* electrons stabilized (i.e., lowered in energy) as they occupy the bonding orbitals in O8. One such interaction is shown at right. Diagrams of the four highest occupied molecular orbitals of O8 (which may be viewed as (O2) 4) are in the reference. 8.31 S2 is similar to O2, with a double bond. As a result, the bond is shorter than single bonds in S8. 8.32 MnF62– acts as a Lewis base, transferring two F ions to two SbF5 molecules: – – MnF62– + 2 SbF5 2 SbF6 + MnF4 The MnF4 intermediate can then lose an F atom, with two of these radicals coupling to form F2: 2 MnF4 F2 + 2 MnF3 8.33 a. Although the ClO3 groups in Cl2O7 are highly electronegative, their size is apparently responsible for the larger angle in Cl2O7 (118.6°) than in Cl2O (110.9°). O Cl O3Cl Cl 110.9° b. 118.6° 2- O O ClO3 O3Cr 126° CrO3 In the dichromate ion the highly electropositive Cr atoms allow the central oxygen to attract electrons more strongly than is possible in Cl2O7, resulting in stronger bonding pair-bonding pair repulsions around the central oxygen in [Cr2O7]2– and a larger bond angle (126°). – 8.34 I3 is linear because there are three lone pairs in a trigonal geometry on the central I. I3+ is bent because there are only two lone pairs on the central I. 8.35 B has only three valence electrons. Adding six from the hydrogen atoms gives B2H6 a total of 12 electrons. The 3-center 2-electron bonds result in four pairs around each boron atom and nearly tetrahedral symmetry at each boron atom. Iodine has seven valence electrons initially. In I2Cl6, five more are added Cl Cl Cl to each iodine, resulting in 12 electrons and octahedral I I Cl Cl Cl geometry around each iodine. 8.36 F + BrF3 BrF4 – – – KF acts as a base; BrF4 is the solvent anion. SbF5 + BrF3 BrF2+ + SbF6 8.37 a. BrF2+ is the cation of the solvent; SbF5 acts as the acid. Br2+ and I2+ have one less antibonding electron than Br2 and I2, so the cations have bond orders of 1.5 and shorter bonds than the neutral molecules: Br2 Br2+ b. – Bond Order 1.0 1.5 Bond Distance (pm) 228 213 Bond Order I2 1.0 I2+ 1.5 Bond distance (pm) 267 256 The most likely transition is from the πg* (HOMO) to u* (LUMO) (see Figure 5.7 for comparable energy levels of F2). Because the observed colors are complementary to the colors absorbed, Br2+ is expected to absorb primarily green light, and I2+ is expected to absorb primarily orange light. Because orange light is less energetic than green, I2+ has Copyright © 2014 Pearson Education, Inc. Chapter 8 Chemistry of the Main Group Elements 117 the more closely spaced HOMO and LUMO. 8.38 c. In both I2 and I2+ the HOMO and LUMO are the πg* and u*, respectively. I2 absorbs primarily yellow light (to give the observed violet color) and I2+ absorbs primarily orange (to give blue). Because yellow light is more energetic than orange, I2 has the more widely spaced HOMO and LUMO. a. In I2+ there are three electrons in the π* orbitals that result from interactions between 5p orbitals (these are the orbitals I I labeled 4πg* in Figure 6.10). I I Interaction between the singly I2+ I42+ occupied π* orbitals of two I2+ ions can lead to stabilization of electrons by formation of a bonding molecular orbital that can connect the I2+ units when I42+ is formed. b. I2+ Higher temperature would provide the energy to break the bonds between the I2+ units and favor the monomer. In addition, formation of two I2+ ions from I42+ would be accompanied by an increase in entropy, also favoring the monomer at higher temperatures. F O 8.39 There are three possibilities: F I F I O F O F F F The third structure, with the lone pair and double bonds in a facial F O arrangement, is least likely because it C2v Cs would have the greatest degree of electron-electron repulsions involving these regions of high electron concentrations. I O O F Cs Infrared data can help distinguish between the other two structures. In the first structure, only the antisymmetric stretch would be IR-active, giving a single absorption band. In the second structure, both the symmetric and antisymmetric stretches would be IR-active, giving two bands. Observation of two I–O bands is therefore consistent with the second structure. This structure, which has fewer 90° lone pair–double bond repulsions than the first structure, is also most likely by VSEPR considerations. Other experimental data are also consistent with the second structure. 8.40 Superhalogens are species which, like the halogens, have very high electron affinities. Classic superhalogens contain a central atom surrounded by highly electronegative atoms or groups of – – atoms. Examples include MnO4, AlCl4 , and BO2 . See K. Pradhan, G. L. Gutsev, P, Jena, J. Chem. Phys., 2010, 133, 144301; C. Sikorska, P. Skurski, Inorg. Chem., 2011, 50, 6384; and references cited therein. 8.41 These reactions take place in the gas phase. The initial product of Xe with PtF6 is believed – to be Xe+ PtF6 ; however, when these two ions are in close proximity, they may react further – to give [XeF]+, [Pt2F11] , and other products. SF6, if present in large excess, prevents the formation of these secondary products, apparently by acting as an inert diluent and preventing effective collisions between the desired products. Copyright © 2014 Pearson Education, Inc. 118 Chapter 8 Chemistry of the Main Group Elements 8.42 F F Xe O F F O O Xe O Xe O O O C4v F C2v 8.43 F F Xe F F C2v F D3h Two electrons are in the bonding orbital and two are in the nonbonding orbital: 5p 2p Xe 8.44 F Xe F F F Xe(OTeF5)4: The OTeF5 group has one electron available for bonding on the O, so the four groups form a square planar structure around Xe, with lone pairs in the axial positions. O=Xe(OTeF5)4: A square pyramidal structure, with O in one of the axial positions and a lone pair on the other, and OTeF5 groups in the square base. 8.45 Half reactions: – Mn2+ + 4 H2O MnO4 + 8 H+ + 5e– XeO64– + 12 H+ + 8e– Xe + 6 H2O Overall reaction: – 8 Mn2+ + 5 XeO64– + 2 H2O 8 MnO4 + 4 H+ + 5 Xe 8.46 – XeF5 has D5h symmetry. The reducible representation for Xe–F stretching is D5h E 5 2C5 0 2C52 0 5C2 1 h 5 2S5 0 2S53 0 which reduces to A1+ E1+ E2, with only E1 IR-active. O F Copyright © 2014 Pearson Education, Inc. 5v 1 F Xe F F Chapter 8 Chemistry of the Main Group Elements 8.47 a. 119 Point group: C4v C4v A1 A2 B1 B2 E E 18 1 1 1 1 2 2C4 2 1 1 –1 –1 0 C2 –2 1 1 1 1 –2 2v 4 1 –1 1 –1 0 b. = 4 A1 + A2 + 2B1 + B2 + 5E c. Translation: A1 + E (match x, y, and z) Rotation: A2 + E (match Rx, Ry, and Rz) Vibration: all that remain: 3 A1 + 2B1 + B2 + 3E 2d 2 1 –1 –1 1 0 z Rz (x, y), (Rx, Ry) 8.48 By the VSEPR approach, XeF22– would be F expected to have a steric number of 6, with four lone pairs on xenon. Two structures, F Xe Xe cis and trans, need to be considered. In F the cis structure there would be five lone pair–lone pair interactions at 90°, and in the F trans structure there would be only four such cis trans interactions. The trans structure, therefore, would be expected to be more likely. However, the number of lone pair–lone pair repulsions would still be high, likely making this a difficult ion to prepare. 8.49 The energy necessary for vibrational excitation decreases with reduced mass; consequently, the lowest energy band, at 1003.3 cm-1, is assigned to the Xe–D stretch of DXeOXeD, the deuterium analogue of HXeOXeH. The remaining unassigned bands, at 1432.7 and 1034.7 cm-1, are therefore due to the Xe–H and Xe–D stretches, respectively, of HXeOXeD, which is also formed. 8.50 a. HXeO63 2 H 2O(l) 2e HXeO4 4 OH o 0.94 V HXeO 4 3 H 2O(l) 6e Xe + 7 OH o 1.24 V b. The disproportionation reaction of interest is: 4 HXeO4 5 OH Xe + 3 HXeO63 3 H 2O(l) o 0.30 V G o nFE o (6 mol)(96485 8.51 a. C J )(0.30 ) 1.7 *105 J (170 kJ ) . C mol The synthesis of XeO 2 is very challenging to carry out. A special reactor composed of a perfluoro-ethylene/propylene copolymer must be rendered completely moisture free by extended treatment under vacuum. The reactivity of the container surface is then further reduced by treatment with F2 gas. Once the container is prepared, a small volume of water is added in an Ar gas atmosphere, and cooled to 0 °C. The reaction is initiated by Copyright © 2014 Pearson Education, Inc. 120 Chapter 8 Chemistry of the Main Group Elements slowly adding XeF4 crystals. The authors emphasize the importance of both the order of the addition (water first, then XeF4 ) and the speed of addition. The large and negative reaction H , coupled with the relatively small volume of solvent to dissipate the heat, can result in complications, including decomposition of the product; the mixture must be kept at 0 °C. Within 30 seconds after mixing, yellow XeO 2 precipitates from the solution. The XeO 2 solid is then separated from the reaction mixture via centrifugation of the mini-reactor itself at 0 °C. The resulting solid is thermally unstable at ambient temperature and must be kept cold (-78 °C) for long-term storage. b. VSEPR theory predicts monomeric XeO 2 to be a bent molecule with a persistent dipole moment that should permit solubility in water. The insolubility of obtained XeO 2 samples suggests an extended XeO 2 structure; individual monomeric units apparently do not persist in water. c. Application of isotopically labeled water in the synthesis facilitated XeO 2 characterization. The presence of exclusively Xe O bonding was inferred by preparing XeO 2 as described in (a) using H 216O and H 218O , respectively. The Raman spectrum of the H 216O -derived XeO 2 was identical to that of the H 218O -derived XeO 2 sample, except that all of the bands of the H 216O -derived XeO 2 sample were shifted to slightly higher energies. The identical spectral patterns, coupled with an isotopic shift in each band, suggested that only Xe O bonding was present. Identical Raman spectra were obtained for XeO 2 samples prepared with either H 216O or D 216O , suggesting no hydrogen in the samples; no isotopic shifts of any Raman bands were observed. These chemists were particularly interested in the possibility of Xe OH units, which was ruled out by the previous observations. 8.52 a. The [XeF]+ ion, with a total of 14 valence s and p electrons, would be expected to have a bond order of 1 (see Section 5.2.1). b. Using the VSEPR approach, the following structures can be drawn for the cations of compounds 1-4: F F F S N Xe F F F S F F N S 1 F 2 Xe F N F F F F H H F 3 S F F N 4 H Xe The [F3NXeF]+ ion, 1, with a triple bond between sulfur and nitrogen, has the shortest N–S distance (139.7 pm), and [F5SN(H)Xe]+, with a single bond, has the longest distance (176.1 pm). The other ions have bond distances between these extremes that are consistent with double bonds between sulfur and nitrogen. c. Compound 1 is expected to have a cation with linear bonding around nitrogen, as shown. However, its crystal structure shows significant bending, with an S–N–Xe angle of Copyright © 2014 Pearson Education, Inc. Chapter 8 Chemistry of the Main Group Elements 121 142.6°. This bending is attributed to close N…F contacts within the crystalline lattice. d. VSEPR would predict linear bonding with three lone pairs on xenon as, for example, in XeF2. The angle measured by X-ray crystallography is 179.6°. e. With significant S–N double bonding, these groups are equatorial. f. The S–Faxial bonds are longer, as in many other VSEPR examples (see Figures 3.17 through 3.19). Average distances for these ions are: (2: S–F (axial), 158.2 pm; S–F (equatorial), 152.4 pm) and (3: S–F (axial), 156.1 pm; S–F (equatorial), 151.8 pm). 8.53 a. b. D2h D4d 8.54 [FBeNe]+ The relative order of molecular orbitals may vary depending on the software used. A simple approach, using an extended Hückel calculation, gave the following results: F–Be bonding c. d. C1 C2v e. f. Cs C2 Several orbitals involved: HOMO: degenerate pair π bonding between Be and F Next orbital below HOMO bonding by the F pz and the Be s Be–Ne bonding One molecular orbital contributing significantly: Four orbitals below HOMO bonding by the Ne pz and the Be s and pz Overall, the interactions between Be and F are stronger than between Be and Ne. 8.55 In Xe2+, there are a total of 54 orbitals. We will classify only the higher-energy orbitals here, starting with the highest energy (as calculated using the relatively simple extended Hückel approach). The orbitals used are the 5s, 5p, and 4d. Molecular Orbital HOMO Type of Interaction * Atomic Orbitals pz HOMO – 1 * px HOMO – 2 HOMO – 3 HOMO – 4 py pz px py Molecular Orbital HOMO – 8 HOMO – 9 HOMO – 10 HOMO – 11 Type of Interaction * Atomic Orbitals d 2 2 , d xy d x 2 y 2 , d xy HOMO – 12 HOMO – 13 HOMO – 14 HOMO – 15 * d xz , d yz d xz , d yz dz2 HOMO – 5 HOMO – 6 * s HOMO – 16 * HOMO – 7 s HOMO – 17 Copyright © 2014 Pearson Education, Inc. x y dz2 122 Chapter 8 Chemistry of the Main Group Elements If this were neutral Xe2, there would be no bond, because every occupied bonding orbital would be offset by an occupied antibonding orbital. In Xe2+, one electron has been removed, so the bond order is 1 2 , making for a long bond. The large atomic radius of Xe further predisposes Xe2+ to having a long bond. It is noteworthy that bonding is exceedingly weak in this species since the overlap leading to formation of the and * orbitals is very small. 8.56 The reference by Steudel and Wong provides illustrations for the four highest occupied molecular orbitals. In addition to the interesting symmetry of these orbitals, the way in which the * orbitals of O2 are imbedded in the molecular orbitals of O8 should be noted. Copyright © 2014 Pearson Education, Inc. Chapter 9 Coordination Chemistry I: Structures and Isomers 123 CHAPTER 9: COORDINATION CHEMISTRY I: STRUCTURES AND ISOMERS 9.1 Hexagonal: C2v C2v D2h Hexagonal pyramidal: Cs Cs C2v Trigonal prismatic: Cs C2v C2 Trigonal antiprismatic: Cs C2 C2h The structures with C2 symmetry would be optically active. 9.2 a. dicyanotetrakis(methylisocyano)iron(II) or dicyanotetrakis(methylisocyano)iron(0) b. rubidium tetrafluoroargentate(III) or rubidium tetrafluoroargentate(1–) c. cis- and trans-carbonylchlorobis(triphenylphosphine)iridium(I) or cis- and transcarbon ylchlorobis(triphenylphospine)iridium(0) 9.3 d. pentaammineazidocobalt(III) sulfate or pentaammineazidocobalt(2+) sulfate e. diamminesilver(I) tetrafluoroborate(III) or diamminesilver(1+) tetrafluoroborate(1–) – (The BF4 ion is commonly called “tetrafluoroborate.”) a. tris(oxalato)vanadate(III) or tris(oxalato)vanadate(3–) b. sodium tetrachloroaluminate(III) or sodium tetrachloroaluminate(1–) c. carbonatobis(ethylenediamine)cobalt(III) chloride or carbonatobis(ethy lenediamine)cobalt(1+) chloride 9.4 d. tris(2,2-bipyridine)nickel(II) nitrate or tris(2,2-bipyridine)nickel(2+) nitrate (The IUPAC name of the bidentate ligand, 2,2-bipyridyl may also be used; this ligand is most familiarly called “bipy.”) e. hexacarbonylmolybdenum(0) (also commonly called “molybdenum hexacarbonyl”). The (0) is often omitted. a. tetraamminecopper(II) or tetraamminecopper(2+) b. tetrachloroplatinate(II) or tetrachloroplatinate(2–) c. tris(dimethyldithiocarbamato)iron(III) or tris(dimethyldithiocarbamato)iron(0) d. hexacyanomanganate(II) or hexacyanomanganate(4–) e. nonahydridorhenate(VII) or nonahydridorhenate(2–) (This ion is commonly called “enneahydridorhenate.”) Copyright © 2014 Pearson Education, Inc. 124 9.5 Chapter 9 Coordination Chemistry I: Structures and Isomers a. triamminetrichloroplatinum(IV) or triamminetrichloroplatinum(1+) b. diamminediaquadichlorocobalt(III) or diamminediaquadichlorocobalt(1+) c. diamminediaquabromochlorocobalt(III) or diamminediaquabromochlorocobalt(1+) d. triaquabromochloroiodochromium(III) or triaquabromochloroiodochromium(0) e. or dichlorobis(ethylenediamine)platinum(IV) or dichlorobis(ethylenediamine)platinum(2+) dichlorobis(1,2-ethanediamine)platinum(IV) or dichlorobis(1,2ethanediamine)platinum(2+) f. diamminedichloro(o-phenanthroline)chromium(III) or diamminedichloro(ophenanthroline)chro mium(1+) or diamminedichloro(1,10-phenanthroline)chromium(III) or diamminedichloro(1,10-phenanthroline)chromium(1+) g. or bis(2,2-bipyridine)bromochloroplatinum(IV) or bis(2,2bypyridine)bromochloroplatinum(2+) bis(2,2-bipyridyl)bromochloroplatinum(IV) or bis(2,2bipyridyl)bromochloroplatinum(2+) h. dibromo[o-phenylene(dimethylarsine)(dimethylphosphine)]rhenium(II) or dibrom o[o-phenylene(dimethylarsine)(dimethylphosphine)]rhenium(0) or dibrom o[1,2-phenylene(dimethylarsine)(dimethylphosphine)]rhenium(II) or dibrom o[1,2-phenylene(dimethylarsine)(dimethylphosphine)]rhenium(0) i. dibromochlorodiethylenetriaminerhenium(III) or dibrom ochlorodiethylenetriaminerhenium(0) or dibromochloro(2,2diam inodiethylamine)rhenium(III) or dibromochloro(2,2diam inodiethylamine)rhenium(0) 9.6 9.7 a. dicarbonylbis(dimethyldithiocarbamato)ruthenium(III) or dicarbonylbis(dimethyldithiocarbamato)ruthenium(1+) b. trisoxalatocobaltate(III) or trisoxalatocobaltate(3–) c. tris(ethylenediamine)ruthenium(II) or tris(ethylenediamine)ruthenium(2+) d. bis(2,2-bipyridine)dichloronickel(II) or bis(2,2-bipyridine)dichloronickel(2+) a. Bis(en)Co(III)-µ-amido-µ-hydroxobis(en)Co(III) N N Co N 4+ N N H2 N O H Co N N N Copyright © 2014 Pearson Education, Inc. Chapter 9 Coordination Chemistry I: Structures and Isomers b. 125 Diaquadiiododinitritopalladium(IV) I ONO H2O Pd H2O ONO H 2O ONO H 2O I I OH2 ONO OH2 H 2O ONO I I I OH2 I OH2 Pd ONO ONO ONO Pd I I I Pd I Pd ONO ONO I Pd ONO OH2 ONO OH2 H2O enantiomers c. Fe(dtc)3 S S S S Fe S S S S Fe S S S S = C – N S S S CH3 H S At low temperature, restricted rotation about the C—N bond can lead to additional isomers as a consequence of the different substituents on the nitrogen. These isomers can be observed by NMR. Copyright © 2014 Pearson Education, Inc. 126 9.8 Chapter 9 Coordination Chemistry I: Structures and Isomers a. triammineaquadichlorocobalt(III) chloride + H2O H3N H 3N Co Isomers are of the cation: + Cl Cl H3N NH3 H3N Co OH2 H2O NH3 Cl Co Cl Cl Cl cis trans mer b. H3N O Cr NH3 H3N 4+ NH3 H3N H3N H3N c. fac -oxo-bis(pentammine-chromium(III)) ion Cr NH3 NH3 NH3 potassium diaquabis(oxalato)manganate(III) – O O H2O Mn O O O O H2O Mn Isomers are of the anion: – O – H2O O O OH2 O OH2 O H2O trans Cl a. O Mn cis enantiomers 9.9 NH3 cis-diamminebromochloroplatinum(II) Pt Br NH3 I b. diaquadiiododinitritopalladium(IV) H2O Pt ONO ONO OH2 I c. tri--carbonylbis(tricarbonyliron(0)) O C O O C C Fe OC + NH3 O C CO Fe CC O O Copyright © 2014 Pearson Education, Inc. CO NH3 NH3 Chapter 9 Coordination Chemistry I: Structures and Isomers O 9.10 C CH2 O – H2 N O N = N O O M O N N O O N M N N O O N N M N N N O O N O mer M(AB)3 B A B M B A A B B A M A A B B A A M A A A B B A B mer [Pt(NH3)3Cl3]+ a. B M B fac 9.12 O M O fac 9.11 127 + NH 3 Cl Pt Cl + Cl NH3 H3N NH3 Cl NH3 Pt NH3 Cl Cl fac mer [Co(NH3)2(H2O)2Cl2]+ b. + NH3 H 2O Cl Co + Cl NH3 H 2O OH2 H 2O Cl Co NH3 Cl NH3 Cl + OH2 Cl Co Co Cl OH2 H2 O NH3 H 3N NH3 H3N Cl Co Cl enantiomers Copyright © 2014 Pearson Education, Inc. Co NH3 + H 2O OH2 Cl + NH3 OH2 NH3 Cl H 2O + NH3 OH2 Cl 128 Chapter 9 Coordination Chemistry I: Structures and Isomers [Co(NH3)2(H2O)2BrCl]+ c. + OH2 H 3N H 3N Co Br H 2O Cl H 2O Co OH2 + Br Co NH3 Br NH3 Cl Co + H 3N NH3 H 3N Co Cl OH2 Cl OH2 H2 O H2 O Br Cl + Co NH3 H3 N NH3 H3N + OH2 Co Cl Br enantiomers d. Br H2O Br Cl OH2 Co NH3 OH2 OH2 + NH3 NH3 H2O NH3 + NH3 Br OH2 H 2O + Cl enantiomers Cr(H2O)3BrClI OH2 H2 O H2O Cr Cl OH2 H2 O I I OH2 Cr Br Cl Br enantiomers Cl H2 O Cr OH2 I Br H 2O Cr OH2 Cl Br H2 O OH2 Cr I Br Cl I e. OH2 OH2 OH2 [Pt(en)2Cl2]2+ 2+ N N Cl Pt 2+ N N N N N Cl Pt N N Cl N Cl cis enantiomers Copyright © 2014 Pearson Education, Inc. 2+ Cl Pt Cl trans N N Chapter 9 Coordination Chemistry I: Structures and Isomers [Cr(o-phen)(NH3)2Cl2]+ f. + N Cl NH3 H 3N NH3 H3N N N Cr enantiomers N NH3 H3N 2+ N N N trans NH3 ligands 2+ N Pt Cl Cl trans Cl ligands N N NH3 Cr Cl 2+ Br Pt Cl N NH3 [Pt(bipy)2BrCl]2+ N Cr Cl Cl Cl Cl N N N Cr + + + N g. 129 Br Br Cl Cl N Pt N N N enantiomers h. Re(arphos)2Br2 Abbreviating the bidentate ligands As P : Br P As Re Br P P As As Re Br Br Re As Br P Br Br P P P P Re Re P As As As As P As As Br P Br Br Br Br P As As Re As As P Copyright © 2014 Pearson Education, Inc. Re P P As As Re P Br Br Br Br 130 Chapter 9 Coordination Chemistry I: Structures and Isomers i. Re(dien)Br2Cl Cl N Re N Br Cl Br Br Br Br Re N N N N N Br N Cl Re Cl N N N Br Re Br 9.13 Br N N N Cl Re N Br a. M(ABA)(CDC) C A A M B D C A D B M C B C A D C M B A A A A C C b. M(ABA)(CDE) C A B M A A D C A D E C D C D M B B A A M E E B B M A A A A M E E A A M A A B B E Copyright © 2014 Pearson Education, Inc. M M E C D C D C D C D Chapter 9 Coordination Chemistry I: Structures and Isomers 131 9.14 A A C B M B A C B M B A B B B C M C C A B C C M A A B B C M A A B B M C C C C B B B B M C C A B M A A B B M C C 9.15 B C A A M B C A A A A A C C M C A B A B C a. The “softer” phosphorus atom bonds preferentially to the soft metal Pd (see Section 6.6.1). b, c. Abbreviating the bidentate ligands N P : Cl P N Ni Cl P P N N Ni Cl Cl Ni N Cl P Cl Cl N Cl P P P P Ni N N N N P N P Ni P Cl Cl Cl Cl Ni P N N Ni N N P Copyright © 2014 Pearson Education, Inc. P P N N Ni P Cl Cl Cl Cl 132 9.16 Chapter 9 Coordination Chemistry I: Structures and Isomers a, b. Abbreviating the bidentate ligands N P and O S : Cl S O N Cl Cl M P P O O Cl P S N O M Cl Cl N P M Cl Cl Cl Cl M N P P N N O O M S S N N P P S Cl M Cl M P P S S M S Cl Cl Cl Cl M N N S S M O O O O Cl Cl Cl Cl – N 9.17 The single C–N stretching frequency indicates a trans structure for the cyanides (the symmetric stretch of the C—N bonds is not IR active), while the two C–O bonds indicate a cis structure for the carbonyls (both the symmetric and antisymmetric C–O stretches are IR active). As a result, the bromo ligands are also cis. C O C Br Co C N Copyright © 2014 Pearson Education, Inc. O C Br Chapter 9 Coordination Chemistry I: Structures and Isomers 9.18 133 There are 18 isomers overall, six with the chelating ligand in a mer geometry and 12 with the chelating ligand in a fac geometry. All are enantiomers. They are all shown below, with dashed lines separating the enantiomers. N P M As P P Br H2O M P P As As M Br Br NH3 H 3N N M P P As As N M NH 3 H3N OH2 H2 O M NH3 OH2 OH2 Br Br N N N N N N OH2 H2O Br Br M P P As As M NH3 H3 N Br Br M P P As As M OH2 H 2O NH3 H3N M NH3 NH3 OH2 OH2 Br Br NH3 NH3 OH2 OH2 NH3 NH3 M As Br OH2 N N NH3 M As N N N OH2 H2O M Br P P As As M Br N N NH3 H3N M b. 9.20 All are chiral if the ring in b does not switch conformations. 9.21 20b 20c top ring: , bottom ring: 9.22 a, b. 1. O Mo 2. S Cs Mo Mo Cs 3. Mo C3v O Mo Cr O S O O Mo Mo Se Se Cr W Mo Mo O Cr Se O C1 S O S O Mo S S Mo O W Cr W Se C1 O Mo Cr W Mo C1 O O W Mo Cr C1 Se C1 W Cs C1 S O O W Mo Mo Mo O Se S Mo Mo Mo Se Cr O Se O W C1 Copyright © 2014 Pearson Education, Inc. O Mo Cr W C1 M OH2 W S S S N Br O Mo W N Br Mo Mo O M O Mo O O S S Mo W d. Mo O Mo O c. As OH2 a. P As Br Br 9.19 P S P As P As P As 134 Chapter 9 Coordination Chemistry I: Structures and Isomers S Mo Mo Mo Cr O c. 9.23 S Se Cs O Mo Se W W O Cs O Cr Yes, provided the structure has no symmetry or only Cn axes. Examples are the structures with C1 symmetry in part a. The 19F doublet is from the two axial fluorines (split by the equatorial fluorine). The 19F triplet is from the equatorial fluorine (split by the two axial fluorines). The two doubly bonded oxygens are equatorial, as expected from VSEPR considerations. Point group: C2v 9.24 + F O O Os F F Examples include both cations and anions: – – – – – [Cu(CN)2] , [Cu2(CN)3] , [Cu3(CN)4] , [Cu4(CN)5] , [Cu5(CN)6] [Cu2(CN)]+, [Cu3(CN)2]+, [Cu4(CN)3]+, [Cu5(CN)4]+, [Cu6(CN)5]+ Based primarily on calculations (rather than experimental data), Dance et al. proposed linear structures such as the following: – NC—Cu—CN – NC—Cu—CN—Cu—CN – NC—Cu—CN—Cu—CN—Cu—CN + Cu—CN—Cu + Cu—CN—Cu—CN—Cu + Cu—CN—Cu—NC—Cu—CN—Cu [Cu(CN)2] : [Cu2(CN)3] : [Cu3(CN)4] : [Cu2(CN)] : [Cu3(CN)2] : [Cu4(CN)3] : Where 2-coordinate copper appears in these ions, the geometry around the Cu is linear, as expected from VSEPR. 9.25 The bulky mesityl groups cause sufficient crowding that the phosphine ligands can show HC chirality (C3 symmetry) and can be considered as similar to left-handed (PL) and right-handed (PR) propellers. If two P(mesityl)3 phosphines are attached in a linear arrangement to a gold atom, three isomers are possible: 3 PL—Au—PL PR—Au—PR PL—Au—PR H 3C mesityl (PR—Au—PL is equivalent to PL—Au—PR, as can best be seen with models.) NMR data at low temperature support the presence of these isomers, which interconvert at higher temperatures. Copyright © 2014 Pearson Education, Inc. CH3 Chapter 9 Coordination Chemistry I: Structures and Isomers 9.26 The point group is D3h. A representation based on the nine 1s orbitals of the hydride ligands is: D3h A1 E A2 E E 9 1 2 1 2 2C3 0 1 –1 1 –1 3C2 1 1 0 –1 0 2S3 0 1 –1 –1 1 h 3 1 2 –1 –2 3v 3 0 0 1 0 Re z2 (x, y), (x2–y2, xy) z (xz, yz) 135 = H The representation reduces to 2 A1 + 2 E + A2 + ECollectively these representations match all the functions for s (totally symmetric, matching A1), p, and d orbitals of Re, so all the s, p, and d orbitals of the metal have suitable symmetry for interaction. (The strength of these interactions will also depend on the match in energies between the rhenium orbitals and the 1s orbital of hydrogen.) 9.27 NN NN NN O Cr O O Cr O O Cr O O OO OO OO Mn O O O OO O Mn O O O O O OO OO O Mn O O O O O OO O O Mn O OO O O Cr O O Cr O O Cr O NN NN NN 9.28 N N N N N O O O O N N N N O N N N Co O O N Co N O O N N N N Co O O O N N N N O O N Co N O O O N N O O O N N N Co O O N N N Co O O N N N N Copyright © 2014 Pearson Education, Inc. 136 9.29 9.30 Chapter 9 Coordination Chemistry I: Structures and Isomers a. Cu(acacCN)2: D2h b. C6 tpt: C2v All four metal-organic frameworks studied (MOF-177, Co(BDP), Cu-BTTri, Mg 2 (dobdc) ) are significantly more effective at adsorbing carbon dioxide relative to adsorbing hydrogen. This is attributed, in part, to the higher polarizability of CO 2 relative to that of H 2 . The formation of an induced dipole in these gases by exposed cations within MOFs is an important prerequisite for adsorption. The two MOF properties that most strongly correlate with CO 2 adsorption capacity are MOF surface area and MOF accessible pore volume. As these values (tabulated below) increase, the CO 2 adsorption capacity increases. MOF Surface Area ( m2 g ) Accessible Pore Volume ( cm3 g ) MOF-177 4690 1.59 Co(BDP) 2030 0.93 Cu-BTTri 1750 0.713 Mg 2 (dobdc) 1800 0.573 The graphs in Figure 1 of the reference clearly indicate that Mg 2 (dobdc) adsorbs the most CO 2 at 5 bar. The arrangement and concentration of open Mg 2+ cation sites on the Mg 2 (dobdc) surface is hypothesized to render this MOF more susceptible to CO 2 adsorption. This MOF, along with Cu-BTTri, which also features exposed metal sites, are identified as the best prospects for CO 2 H 2 separation. 9.31 The synthesis and application of amine-functionalized MOFs for CO 2 adsorption is the general topic of the reference. While the M 2 (dobdc) series of MOFs were proposed as excellent candidates for this functionalization (on the basis of their relatively large concentration of exposed metal cation sites), their amine-functionalization proved difficult. This was attributed to the relatively narrow MOF channels that may hinder amine diffusion into M 2 (dobdc) . One hypothesized solution was to prepare a MOF with the M 2 (dobdc) structure-type, but with larger pores. The wider linker dobpdc (below, along with dobdc for comparison) was used in the hope of obtaining MOFs with larger pores. O O O O O O O O O O O O dobdc dobpdc Copyright © 2014 Pearson Education, Inc. Chapter 9 Coordination Chemistry I: Structures and Isomers 137 Amine-functionalized Mg 2 (dobpdc) was prepared by mixing H 4 (dobpdc) , magnesium bromide, and a small solvent volume (a mixture of N,N’-diethylformamide and ethanol) in a Pyrex container. The mixture was heated in a microwave reactor, and the M 2 (dobpdc) collected by filtration after cooling. Dried samples of Mg 2 (dobpdc) were then heated for roughly one hour at 420 °C under dynamic vacuum. After this “activation” step, Mg 2 (dobpdc) was stirred with an excess of N,N’-dimethylethylenediamine (mmen) in hexanes for one day. Subsequent heating under vacuum resulted in removal of residual solvents to afford mmen-functionalized Mg 2 (dobpdc) . The “activation” step was found necessary to completely remove residual N,N’diethylformamide from the Mg 2+ coordination sites. 9.32 This reference discusses application of porphyrin-containing MOFs where the porphyrin provides a binding site for Fe(III) and Cu(II). The precursor to the porphyrin linker (TCPP) is provided below; the resulting carboxylates of this linker permit its incorporation into the MOF. HOOC COOH N H N N H N HOOC COOH The metallation options include premetallation and postmetallation. In premetallation, H 4 -TCPP-Cu and H 4 -TCPP-FeCl , respectively, are used as reactants for the MOF synthesis. In this case, the porphyrin linker and its bound metal ion are installed simultaneously into the MOF. This general approach afforded MOF-525-Cu, MOF-545-Fe, and MOF-545-Cu. MOF-525-Fe could not be obtained via this strategy. For this MOF, postmetallation was employed, via the reaction of MOF-525 with Fe(III) chloride; Fe(III) ions were introduced into the MOF-525 porphyrin linkers via this method. In terms of similarities and differences, MOF-545 can be metallated with both Fe(III) and Cu(II) via a premetallation strategy, while MOF-525 requires alternate procedures for incorporation of Cu(II) (premetallation) and Fe(III) (postmetallation), respectively. Copyright © 2014 Pearson Education, Inc. 138 Chapter 10 Coordination Chemistry II: Bonding CHAPTER 10: COORDINATION CHEMISTRY II: BONDING 10.1 10.2 a. Tetrahedral d 6, 4 unpaired electrons b. [Co(H2O)6]2+, high spin octahedral d 7, 3 unpaired electrons c. [Cr(H2O)6]3+, octahedral d 3, 3 unpaired electrons d. square planar d 7, 1 unpaired electron e. 5.1 BM = = a. [M(H2O)6]3+ with 1 unpaired electron. b. [MBr4]– with the maximum number of unpaired electrons (5). n n 2 ; n 4.2 4 M = Ti d1: M3+ with 5 d electrons: M = Fe c. Diamagnetic [M(CN)6]3–. The strong field cyano ligand favors low spin. M3+ with 6 d electrons: M = Co d. [M(H2O)6]2+ having LFSE = o . Both high spin d4 and d9 have the correct o. 3 5 0.6 High spin d4: M = Cr 0.6 d9: M = Cu – 0.4 10.3 a. K3[M(CN)6] M is first row transition metal, 3 unpaired electrons. M3+ with 3 d electrons: M = Cr Copyright © 2014 Pearson Education, Inc. – 0.4 Chapter 10 Coordination Chemistry II: Bonding b. 139 [M(H2O)6]3+ M is second row transition metal, LFSE = 2.4 o. This can be achieved with a low spin d6 configuration. M3+ with 6 d electrons: M = Rh 0.6 – 0.4 c. [MCl4]– M is first row transition metal, 5 unpaired electrons. M3+ with 5 d electrons: Fe d. MCl2(NH3)2 M is third row d8 transition metal; two M–Cl stretching bands in IR. Second and third row d8 complexes are often square planar. The presence of two M–Cl stretching bands implies cis geometry of the chloro ligands. Cl M with 8 d electrons: M = Pt 2+ NH3 Pt Cl NH3 10.4 Angular overlap calculations for d 8 and d 9 ions show no energy difference between D4h and Oh when exclusively considering interactions. Both d 8 geometries have energies of -3e; both d 9 geometries have energies of -6e. In general, stability constants decrease as more ligands are added; the sequence for nickel is the common one. The huge drop in stability constant between the second and third ethylenediamine on Cu2+ is a result of the d 9 Jahn-Teller effect. The first two en ligands add in a square-planar geometry, with water molecules in the axial positions, and the coordination of these two mondentate ligands allows for the Jahn-Teller distortion. Adding a third en ligand requires a geometry change and with a preference for uniform M—N bond distances towards the six nitrogen atoms. This is counter to the Jahn-Teller distortion, and the third addition is much less favorable than the first two. 10.5 [M(H2O)6]2+ M is first row transition metal, µ = 3.9 BM. The magnetic moment implies 3 unpaired electrons, which would give rise to 3 3 5 = 3.9 BM. There are two possibilities, d3 and d7: M2+ , d3 : M = V 10.6 a. [Cr(H2O)6]2+ b. 4– M2+ , d7: M = Co 4(6) 4.9 B [Cr(CN)6] n=2 2(4) 2.8 B – c. [FeCl4] d. [Fe(CN)6]3– 1(3) 1.7 B n=4 e. [Ni(H2O)6]2+ f. [Cu(en)2(H2O)2]2+ n = 1 n=5 5(7) 5.9 B n=1 Copyright © 2014 Pearson Education, Inc. n=2 2(4) 2.8 B 1(3) 1.7 B 140 10.7 Chapter 10 Coordination Chemistry II: Bonding Fe(H2O)4(CN)2 is really [Fe(H2O)6]2[Fe(CN)6], all containing Fe(II). [Fe(H2O)6]2+ is high spin d 6, with = 4.9 µB; [Fe(CN)6]4– is low spin d 6 , with = 0 µB. The average value is then 2 × 4.9/3 = 3.3 µB. 2.67 unpaired electrons gives 2.67 4.67 3.53 µB. 10.8 Co(II) is d 7. In tetrahedral complexes, it is generally high spin and has 3 unpaired electrons; in octahedral complexes, it is also typically high spin and also has 3 unpaired electrons; in square planar complexes, it has 1 unpaired electron. The magnetic moments can be calculated as n(n 2) 3.9, 3.9, and 1.7 B , respectively. 10.9 For the red compounds (Me and Et at high temperatures, Pr, pip, and pyr at all temperatures), the larger magnetic moment indicates approximately 5 unpaired electrons, appropriate for high-spin Fe(III) species. At low temperatures for the Me and Et compounds, the magnetic moment indicates 3 to 4 unpaired electrons, an average value indicating an equilibrium mixture of high and low spin species. The low spin octahedral complexes have 1 unpaired electron. Increasing the size of the R groups changes the structure enough that it is locked into high-spin species at all temperatures. 10.10 Both [M(H2O)6]2+ and [M(NH3)6]2+ should show the double-humped curve of Figure 10.12, with larger values for the NH3 compounds. Therefore, the difference between these curves shows the same general features as in Figure 10.12. 10.11 These can be verified using the approach introduced in Chapters 4 and 5, in which a character of +1 is counted for each operation that leaves a vector unchanged, a character of –1 for each operation that leaves a vector in its position, but with the direction reversed. The general transformation matrix for rotation is provided in Figure 4.16. Application is of this matrix is necessary to deduce the characters involving rotations. 10.12 The e column gives d orbital energies for complexes involving donor ligands only; the Total column gives energies for complexes of ligands that act as both donors and acceptors: a. ML2, using positions 1 and 6: z2 x2–y2 xy xz yz b. eπ 0 0 0 –2 –2 e 2 0 0 0 0 Total 2e 0 0 –2 eπ –2 eπ ML3, using positions 2, 11, 12: z2 x2–y2 xy xz yz e 0.75 1.125 1.125 0 0 eπ 0 –1.5 –1.5 –1.5 –1.5 Total 0.75e 1.125 e – 1.5 eπ 1.125 e – 1.5 eπ –1.5 eπ –1.5 eπ Copyright © 2014 Pearson Education, Inc. Chapter 10 Coordination Chemistry II: Bonding c. ML5, C4v, using positions 1, 2, 3, 4, 5: z x2–y2 xy xz yz eπ 0 –1.5 –1.5 –3.5 –3.5 e 2.75 1.125 1.125 0 0 z x2–y2 xy xz yz Total 2.75e 1.125 e – 1.5 eπ 1.125 e – 1.5 eπ –3.5 eπ –3.5 eπ ML8, cube, positions 7, 8, 9, 10, doubled for the other four corners: e 0 0 2.67 2.67 2.67 2 z x2–y2 xy xz yz 10.13 Total 2e 3e –4eπ –3eπ –3eπ ML5, D3h, using positions 1, 2, 6, 11, 12: 2 e. eπ 0 0 –4 –3 –3 e 2 3 0 0 0 2 d. 141 eπ –5.33 –5.33 –1.78 –1.78 –1.78 Total –5.33 eπ –5.33 eπ 2.67 e – 1.78 eπ 2.67 e – 1.78 eπ 2.67 e – 1.78 eπ – Metal d orbitals, NH3 influence eπ Total e 2 1 0 z 1e 2 2 3 0 x –y 3e 0 0 0 xy 0 0 0 xz 0 0 0 yz Metal d orbitals, Cl influence eπ Total e 2 2 0 z 2e 2 2 0 0 0 x –y 0 0 0 xy 0 2 2 eπ xz 0 2 2 eπ yz Ligand NH3: Ligand Cl : 1 2 3 4 5 6 – e 0 –1 –1 –1 –1 0 Total 0 –1e –1e –1e –1e 0 1 2 3 4 5 6 e –1 0 0 0 0 –1 eπ –2 0 0 0 0 –2 Total –1e – 2 eπ 0 0 0 0 –1e – 2 eπ Overall energy = –8 e(NH3) – 4 e(Cl) – 8 eπ(Cl) + 4 eπ(Cl) = –8 e(NH3) – 4 e(Cl) – 4 eπ(Cl) The metal electrons are still unpaired, one in the dxy orbital and one each in the dxz and dyz, raised by π interaction with Cl–. Four of the ligand orbitals are lowered by e(NH3) and two are lowered Copyright © 2014 Pearson Education, Inc. 142 Chapter 10 Coordination Chemistry II: Bonding by e(Cl) + 2 eπ(Cl); each contains a pair of ligand electrons. The precise orbital energies cannot be calculated since the angular overlap parameters for NH3 and Cl towards Cr(III) are not provided. Table 10.13 does provide the necessary parameters for trans-[Cr(NH 3 )4 F2 ]+ . 10.14 a. If we assume these ligands have similar donor abilities, the energies below are obtained for the molecular orbitals with high d orbital character. z2 metal d orbitals b. x2-y2 xy xz yz 2.75e 1.125e For consideration of L as a -acceptor in the axial position, the identical energy level diagram is obtained regardless of whether L is assigned to position 1 or 6. The xz and yz orbitals are each stabilized by e. z2 x2-y2 2.75e xy 1.125e metal d orbitals e xz yz For consideration of L in an equatorial position, the angular overlap result is different depending on whether the -acceptor is placed in position 2 or position 11/12. This is problematic since all three equatorial positions are indistinguishable. For L in position 2: z2 2.75e x2-y2 xy metal 1.125e e d orbitals 1.125e yz xz Copyright © 2014 Pearson Education, Inc. e Chapter 10 Coordination Chemistry II: Bonding 143 For L in position 11 or 12: z2 metal d orbitals 2.75e xy 2 x -y2 b xz c yz a: 1.125e e b: 1.125e e a c: 0.25e 0.75e Since all three equatorial sites are equivalent, the optimum approach is to consider the weighted average of these two equatorial diagrams, with the second one contributing twice as much as the first since the second diagram is obtained with L in both position 11 and position 12. This average diagram is below: z2 2.75e xy x2-y2 metal d orbitals yz c. xz 1.125e e 0.5e The axial/equatorial preference for the -acceptor ligand depends on the number of metal valence electrons. The table below assumes low-spin configurations, reasonable on the basis of the -acceptor ligand. Alternate configuration energies would be obtained for high-spin cases. This analysis predicts that the axial position would be preferred for the -acceptor ligand in d 1 to d 7 cases. There is no difference between the configuration energies for the d 8 to d 10 cases. This analysis includes neither the exchange energy nor the coulombic energy of repulsion contributions. Ligand positions are strongly influenced by steric considerations, also not accounted for in this generic case. Equatorial L Configuration Energy 0.5e Preference: Axial or Equatorial L ? d1 Axial L Configuration Energy e d2 2e e Axial d3 3e 1.5e Axial d4 4e 2e Axial Metal Valence Electron Count Copyright © 2014 Pearson Education, Inc. Axial 144 10.15 Chapter 10 Coordination Chemistry II: Bonding d5 1.125e 4e 1.125e 2.5e Axial d6 2.25e 4e 2.25e 3e Axial d7 3.375e 4e 3.375e 3.5e Axial d8 4.5e 4e 4.5e 4e None d9 7.25e 4e 7.25e 4e None d 10 10e 4e 10e 4e None The angular overlap diagrams for the molecular orbitals with high d orbital character in square pyramidal (left) and trigonal bipyramidal (right) geometries with -donor ligands are below. In each of these geometries, the -donor ligands pairs are stabilized by 1e; these levels are omitted here. x2-y2 z2 z2 3e 2e metal d orbitals xy xz yz x2-y2 xy xz yz 2.75e 1.125e The energies of the resulting configurations, and the preference between these coordination geometries, depend on electron count and whether the complex is low spin or high spin. The low spin configuration energies (ignoring both the exchange energy and the coulombic energy of repulsion contributions) and preferences are: Low Spin Preference 0e Trigonal Bipyramidal Configuration Energy 0e d2 0e 0e None d3 0e 0e None d4 0e 0e None d5 0e 1.125e Square Pyramidal d6 0e 2.25e Square Pyramidal d7 2e 3.375e Square Pyramidal d8 4e 4.5e Square Pyramidal d9 7e 7.25e Square Pyramidal d 10 10e 10e None Metal Valence Electron Count Square Pyramidal Configuration Energy d1 Copyright © 2014 Pearson Education, Inc. None Chapter 10 Coordination Chemistry II: Bonding 145 The corresponding high spin preferences (assuming that configurations with five unpaired electrons are lower in energy than five-electron configurations with electron pairs) are: Metal Valence Electron Count Square Pyramidal Configuration Energy High Spin Preference 0e Trigonal Bipyramidal Configuration Energy 0e d1 d2 0e 0e None d3 0e 1.125e Square pyramidal d4 2e 2.25e Square pyramidal d5 5e 5e None d6 5e 5e None d7 5e 5e None d8 5e 6.125e Square pyramidal d9 7e 7.25e Square Pyramidal d 10 10e 10e None None Although steric configurations are not part of this angular overlap analysis, the trigonal bipyramidal geometry is never preferred relative to square pyramidal in these cases with five donor ligands. For a set of five -donor/-acceptor ligands, the energy levels are as follows for square pyramidal (left) and trigonal bipyramidal (right). x2-y2 z2 z2 3e 2.75e 2e metal d orbitals x2-y2 xy 1.125e e 3.5e 4e xy xz yz xz yz The energies of the resulting configurations, and the preference between these coordination geometries, depend on electron count and whether the complex is low spin or high spin. We will assume only low spin configurations here since ligands with combined -donor/-acceptor capabilities tend to stabilize low spin complexes. The low spin configuration energies (ignoring both the exchange energy and the coulombic energy of repulsion contributions) and preferences are: Copyright © 2014 Pearson Education, Inc. 146 Chapter 10 Coordination Chemistry II: Bonding Low Spin Preference Metal Valence Electron Count Square Pyramidal Configuration Energy d1 4e Trigonal Bipyramidal Configuration Energy 3.5e d2 8e 7e Square Pyramidal d3 12e 10.5e Square Pyramidal d4 16e 14e Square Pyramidal d5 20e 1.125e 15.5e Square Pyramidal d6 24e 2.25e 17e Square Pyramidal d7 2e 24e 3.375e 18.5e Square Pyramidal d8 4e 24e 4.5e 20e d9 7e 24e 7.25e 20e TBP>SP by only 0.5e 4e Square Pyramidal d 10 10e 24e 10e 20e Square Pyramidal Square Pyramidal Although steric configurations are not part of this analysis, the square pyramidal geometry is clearly preferred relative to trigonal pyramidal in all these cases with five -donor/-acceptor ligands, except with a d 8 metal where the preference is slight. 10.16 a. Seesaw, using positions 1, 6, 11, and 12 ( donor only) Position z2 x2 – y2 xy xz yz 1 6 11 12 Total 1 1 1/4 1/4 2.5 0 0 3/16 3/16 0.375 0 0 9/16 9/16 1.125 0 0 0 0 0 0 0 0 0 0 Trigonal pyramidal, using positions 1, 2, 11, and 12 ( donor only) Position z2 x2 – y2 xy xz yz 1 2 11 12 Total 1 1/4 1/4 1/4 1.75 0 3/4 3/16 3/16 1.125 0 0 9/16 9/16 1.125 0 0 0 0 0 0 0 0 0 0 Copyright © 2014 Pearson Education, Inc. Chapter 10 Coordination Chemistry II: Bonding b. Number of d Electrons 1 2 3 4 5 6 7 8 9 10 10.17 Energies of d Configurations (in units of e) Seesaw Trigonal Pyramidal Low Spin High Spin Low Spin High Spin 0 0 0 0 0.375 0.750 1.875 3 5.5 8 0 0 0.375 1.5 4 4 4 4.375 5.5 8 0 0 0 0 1.125 2.25 3.375 4.5 6.25 8 0 0 1.125 2.25 4 4 4 5.125 6.25 8 c. The seesaw geometry is favored for d 5—d 9 low spin and d 3, d 4, d 8, and d 9 high spin configurations. For other configurations, both geometries have identical energies according to this approach. (The values shown are for the d orbitals only; each interaction also stabilizes a ligand orbital by e.) a. The energies of the molecular orbitals with high d orbital character are as follows: z2 3.5e xy x2-y2 2.25e metal d orbitals yz xz The new positions (13 and 14 below) are opposite 11 and 12, and have the same values in the table. z2 is affected most strongly, since two ligands are along the z axis; x2 – y2 and xy are also strongly influenced, since they are in the plane of the six ligands. xz and yz are not changed, since they miss the ligands in all directions. Position z2 x2 – y2 Xy Xz yz 1 2 6 11 12 4 13 14 Total 1 1/4 1 1/4 1/4 1/4 1/4 1/4 3.5 0 3/4 0 3/16 3/16 3/4 3/16 3/16 2.25 0 0 0 9/16 9/16 0 9/16 9/16 2.25 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 Copyright © 2014 Pearson Education, Inc. 147 148 Chapter 10 Coordination Chemistry II: Bonding b. The D6h character table indicates that z2 has A1g symmetry, x2 – y2 and xy have E2g symmetry, and xz and yz exhibit E1g symmetry. c. The energies of the molecular orbitals with high d orbital character are as follows: z2 3.5e xy x2-y2 2.25ee metal d orbitals 5e yz z2 0 0 0 0 0 0 0 0 0 3.5 e Position 1 2 6 11 12 4 13 14 Total Overall d. 10.18 x2–y2 0 0 0 3/4 3/4 0 3/4 3/4 3 eπ 2.25 e - 3 eπ xz xy 0 1 0 1/4 1/4 1 1/4 1/4 3 eπ 2.25 e - 3 eπ xz 1 1 1 1/4 1/4 1 1/4 1/4 5 eπ -5 eπ yz 1 0 1 3/4 3/4 0 3/4 3/4 5 eπ -5 eπ If we assume that each low spin configuration will have the maximum number of paired electrons, then the configurations with asymmetrically occupied degenerate orbitals are those for d1, d3, d5, and d7 metals. These are expected to give rise to Jahn-Teller distortion. The ammine Co(III) complex is considerably more stable and is less easily reduced, with the difference primarily in the 3+ species. In addition, the metal in [Co(NH3)6]3+ is surrounded by ammonia molecules, which are more difficult to oxidize than water. This makes transfer of electrons through the ligand more difficult for the ammine complexes. [Co(H2O)6] [Co(H2O)6]2+ d (1s) d 7 (hs) –2.4o –0.8o o 16,750 8,400 [Co(NH3)6]3+ d 6 (1s) –2.4o 24,000 –57,600 [Co(NH3)6]2+ d 7 (hs) –0.8o 10,200 –8,160 LFSE 3+ 6 LFSE –40,200 –6,720 Copyright © 2014 Pearson Education, Inc. Difference 33,480 49,440 Chapter 10 Coordination Chemistry II: Bonding – 149 – 10.19 Cl has the lowest o value and fairly good π donor properties that reduce o. F is next, with less π donor ability. Water has very small π donor ability (only one lone pair not involved in bonding), and ammonia and en have nearly insignificant π donor nor acceptor ability (no lone pairs, antibonding orbitals with the wrong shapes and energies for π bonding). – CN has good π acceptor properties, making o largest for this ligand. 10.20 Ammonia is a stronger field ligand than water. It is a stronger Lewis base ( donor) than water. Water also has a lone pair that can act as a π donor (which leads to a reduction in the ligand field splitting). These factors result in the less electronegative nitrogen on ammonia being a better donor atom than the oxygen atom of water. In the halide ions, all have the same valence electronic structure, so the electronegativity is the determining factor in ligand field strength. Fluoride is also a stronger Brønsted base than the other halides. 10.21 a. Compression moves dz2 up in energy and lowers the ligand energies of positions 1 and 6. dz2 eg d x2–y2 dxz, dyz t2g b. Stretching reverses the changes. In the limit of a square planar structure, dz2 is affected only through interactions with the ring in the xy plane. dxy dx2–y2 eg d z2 dxy t2g dxz, dyz Oh D4h 10.22 Cr3+ has three singly occupied t2g orbitals and two empty eg orbitals. As a result, its complexes exhibit no Jahn-Teller distortion. Mn3+ has one electron in the eg orbitals; its complexes show Jahn-Teller distortion. 10.23 a. 4– 3+ 4– 3+ [Co(CO)4] [Cr(CN)6] [Fe(H2O)6] [Co(NO2)6] [Co(NH3)6] – [MnO4] 2+ [Cu(H2O)6] b. LFSE (in o) 0 0 2 2.8 –1.6 5 5.9 0 3 3.9 –0.8 0 0 –2.4 0 0 0 1 1.7 –0.6 n – 0 The two tetrahedral ions ([Co(CO)4]– and [MnO4]–) have zero LFSE (10 or 0 d electrons, respectively) and have ligand equipped to participate in significant π bonding, with CO as an acceptor and O2– as donor. With the exception of [Fe(H2O)6]3+, the others have LFSE values that favor octahedral structures. [Fe(H2O)6]3+ has LFSE = 0 for either Copyright © 2014 Pearson Education, Inc. 150 Chapter 10 Coordination Chemistry II: Bonding octrahedral or tetrahedral shapes, but water is a slight π donor and Fe(III) is only a moderately good π acceptor. As a result, electrostatics favors six ligands. c. The difference in the LFSEs can be used to assess the relative stabilities of these geometries for Co(II) and Ni(II): For Co2+: High-spin octahedral d 7 has LFSE = –0.8o Tetrahedral d 7 has LFSE = –1.2t = –0.53o For Ni2+: High-spin octahedral d 8 has LFSE = –1.2o Tetrahedral d 8 has LFSE = –0.6t = –0.27o Co(II) (d 7) has only –0.27o favoring the octahedral shape, while Ni(II) (d 8) has –0.93o. Therefore, Co(II) compounds are more likely to be tetrahedral than are Ni(II) compounds. 10.24 a. The ligand field stabilization energies and the LFSEO LFSET differences are below. h d The LFSE is more negative for an octahedral field than the corresponding LFSE for a tetrahedral field, regardless of the ligand field strength. The magnitude of the difference ( LFSEO LFSET ) provides some insight into preferences between these geometries for h d these ions. The likelihood of tetrahedral geometry increases as the energy difference between the octahedral and tetrahedral LFSE decreases. Note that this solution considers both high and low spin tetrahedral complexes even though high spin configurations are more common in tetrahedral complexes. The rightmost column provides the difference between the strong field octahedral LFSE and the weak field tetrahedral configurations. Fe 2+ (d 6 ) Co2+ (d 7 ) Ni2+ (d 8 ) Octahedral Field LFSE Weak Strong Tetrahedral Field LFSE Weak Strong 2 o 5 4 o 5 6 o 5 8 45 o 16 o 45 24 o 45 12 5 o 9 o 5 6 o 5 LFSEO LFSET h 48 45 o 36 o 45 24 o 45 Weak — Weak Strong — Strong 0.222 o 1.333 o 0.444 o o 0.667 o 0.667 o d Strong (L.S.)— Weak (H.S.) 2.22 o 1.44 o 0.667 o The magnitudes of the LFSE O LFSET differences predict the preference for h d tetrahedral geometry with weak field ligands to be Fe 2+ Co2+ Ni2+ . The corresponding preference for tetrahedral geometries with a strong field ligand set is predicted as Ni2+ Co2+ Fe 2+ . Note that this ranking is the same with strong field ligands whether the comparison is made to the corresponding high or low spin tetrahedral configuration (the “strong—strong” and “strong—weak” differences offer the same preference rankings). These calculations suggest that the octahedral preference for Fe 2+ and Co2+ is even stronger with a strong field ligand set than with a weak field ligand set. Copyright © 2014 Pearson Education, Inc. Chapter 10 Coordination Chemistry II: Bonding 151 This approach predicts that the Ni2+ preference for octahedral geometry is independent of the ligand field strength. This LFSE analysis does not match the observation that the number of tetrahedral complexes for these ions ranks as Co2+ Fe 2+ Ni2+ , but does suggest that Co2+ is more likely to afford tetrahedral complexes relative to Ni2+ when the ligand field is weak. b. The energies of these electronic configurations on the basis of the angular overlap model are below. Note that this solution also considers both high and low spin tetrahedral complexes even though high spin configurations are more common in tetrahedral complexes. The rightmost column provides the difference between the strong field octahedral and the weak field tetrahedral electronic configurations, respectively. As in a, the electronic stabilization afforded by the octahedral field is greater than that provided by a tetrahedral field, with the preference for an octahedral geometry over tetrahedral geometry generally more pronounced with strong field ligands (with Ni2+ being an exception with preferences for octahedral geometry that are independent of the ligand field strength). Octahedral Weak Strong Tetrahedral Weak Strong Weak Oh Td Strong Fe 2+ (d 6 ) 6e 12e 4e 5.33e 2e 6.67e Strong (L.S.) —Weak (H.S.) 8e Co 2+ (d 7 ) 6e 9e 4e 4e 2e 5e 5e Ni2+ (d 8 ) 6e 6e 2.67e 2.67e 3.33e 3.33e 3.33e The magnitudes of the Oh Td angular overlap differences predict the preference for . The tetrahedral geometry with weak field ligands to be corresponding preference for tetrahedral geometries with a strong field ligand set is predicted as Ni2+ Co2+ Fe 2+ regardless of whether low spin/low-spin (strong— strong) or low-spin/high spin (strong—weak) configurations are compared. This angular overlap analysis does not match the observation that the number of tetrahedral complexes for these ions ranks as Co2+ Fe 2+ Ni2+ , but does suggest that the tetrahedral preference is when the ligands are weak. Neither of these approaches is highly predictive in these cases. The complexity of the effects that govern geometric preferences cannot be encompassed in any single approach. 10.25 The energy levels of the square planar d orbitals are shown in Figure 10.14; those of the octahedral orbitals are shown in Figure 10.5. The seventh, eighth, and ninth d electrons in an octahedral complex go into the highest orbitals, raising the total energy of the complex. In a square planar complex, the seventh and eighth metal d electrons occupy a relatively nonbonding orbital. When strong field -acceptor ligands are employed, the o between the t2g and eg orbitals increases in octahedral complexes, further stabilizing configurations with 6 or less metal Copyright © 2014 Pearson Education, Inc. 152 Chapter 10 Coordination Chemistry II: Bonding d electrons. In these cases, the energy of the square planar complex may be more favorable even though the total ligand electron energy is less than in octahedral complexes because there are only four ligands rather than six. 10.26 a. Square pyramidal complexes have C4v symmetry. C4v A1 B1 E 2C4 1 1 –1 0 E 5 1 1 2 The z2 and x2–y2 orbitals are the major d orbitals used in the bonding, so the d orbital energies are as shown at right. The z2 orbital is less involved because it is directed primarily toward one ligand, and the x2–y2 orbital is directed toward four ligands. C2 1 1 1 –2 2v 3 1 1 0 2 d 1 1 –1 0 dx2–y2 d z2 dxz, dyz d dxy M b. D5h A1 E1 E2 A2 z, z2 x – y2 (x, y), (xz, yz) 2 ML5 5L Pentagonal bipyramidal complexes have D5h symmetry. E 7 1 2 2 1 2C5 2 1 2 cos 72° 2 cos 144° 1 2C52 2 1 2 cos 144° 2 cos 72° 1 5C2 1 1 0 0 –1 h 5 1 2 2 –1 2S5 0 1 2 cos 144° 2 cos 72° –1 2S53 0 1 2 cos 72° 2 cos 144° –1 5v 3 1 0 0 1 z2 (x, y) 2 (x – y2, xy) z The representation reduces to 2 A1+ E1 + E2 + A2The d orbitals are in three sets, including a nonbonding pair (dxz and dyz) in addition to the A1 and degenerate E2. dxy, dx2–y2 dz2 d dxz, dyz M ML7 Copyright © 2014 Pearson Education, Inc. 7L Chapter 10 Coordination Chemistry II: Bonding 10.27 153 The angular overlap parameters are (Cl), 5430 cm 1 ; e (Cl), 1380 cm 1 ; e (PPh3 ), 3340 cm 1 ; e (PPh3 ), 310 cm 1 . Application of these values using Tables 10.10 and 10.11 with the ligand positions arbitrarily defined as shown, provides these results: Stabilization of Donor Pairs (7) 1e (Cl) 2e (Cl) 8190 cm1 (8) 1e (PPh3 ) 3340 cm1 (9) 1e (PPh3 ) 3340 cm 1 1e (PPh3 ) 3340 cm 1 (10) Cl (7) Co Ph3P (8) PPh3 (10) PPh3 (9) The three acceptor orbitals associated with the PPh 3 ligands are destabilized by 2 e (PPh3 ) 207 cm 1 . 3 The energies of the molecular orbitals with high metal d orbital character are shown below. The triphenylphosphine and chloride interactions and the chloride interactions contribute to the destabilization of these orbitals, while the -acceptor capability of triphenylphosphine slightly lowers the energies of the xy, xz, and yz orbitals. z2 x2 y2 xy xz yz 2 e (Cl) 2e (PPh3 ) 300 cm1 3 2 e (Cl) 2e (PPh3 ) 300 cm1 3 1 2 2 1e (PPh3 ) e (Cl) e (Cl) e (PPh 3 ) 5250 cm1 3 9 3 1 2 2 1e (PPh3 ) e (Cl) e (Cl) e (PPh 3 ) 5250 cm1 3 9 3 1 2 2 1e (PPh3 ) e (Cl) e (Cl) e (PPh 3 ) 5250 cm1 3 9 3 (Continued) Copyright © 2014 Pearson Education, Inc. 154 Chapter 10 Coordination Chemistry II: Bonding The energy level diagram is: 207 cm-1 8 9 10 -acceptor orbitals of PPh3 xy xz yz 5250 cm-1 xy xz yz z2 x2-y2 z2 x2-y2 300 cm-1 7 8 3340 cm-1 8190 cm-1 8 9 10 9 10 Cl- PPh3 PPh3 PPh3 7 The large degree of stabilization of the chloride donor pair relative to the PPh3 pairs is surprising; triphenylphosphine is generally considered a stronger field ligand relative to chloride. The steric bulk associated with the coordination of three PPh 3 ligands to the Co(I) center may attenuate the donor ability (and acceptor ability) of these ligands in CoCl(PPh3 )3 . The increased -donation of PPh 3 and increased -donation of Cl in NiCl2 (PPh 3 )2 relative to CoCl(PPh3 )3 , respectively, is likely linked to the higher oxidation state of Ni(II) relative to the Co(I). The reduced steric bulk about the Ni(II) center in NiCl2 (PPh 3 )2 relative to around the Co(I) center in CoCl(PPh3 )3 likely plays a role in increasing the overlap necessary for the -interactions. 10.28 a. The MO diagram is similar to that for CO, shown in Figure 5.13, with the atomic orbitals of fluorine significantly lower than the matching orbitals of nitrogen. The NF molecule has 2 more electrons than CO; these occupy π* orbitals and have parallel spins. b. Because its highest occupied sigma orbital (labeled 5a1 for CO) is significantly concentrated on the nitrogen, NF should act as a strong donor. In addition, because the π* orbitals are also strongly concentrated on nitrogen, NF should act as a π acceptor at nitrogen. Because the π* orbitals are singly occupied, NF is likely to be a weaker π acceptor than CO, whose π* orbitals are empty. Overall, NF is likely to be relatively high in the spectrochemical series, but lower than CO. 10.29 a. When these compounds are oxidized, in the positively charged products there is less acceptance by the CO ligands than in the neutral compounds. Because CO is acting less as a acceptor in the products, the products have less donation into their π* orbitals, which are antibonding with respect to the C–O bonds. This means that the oxidation products have stronger and shorter C–O bonds than the reactants. In the reference cited, the C–O bonds are calculated to be shortened by 0.014 to 0.018 Å in the PH3 complex and by 0.015 to 0.020 Å in the NH3 complex. Copyright © 2014 Pearson Education, Inc. Chapter 10 Coordination Chemistry II: Bonding b. 155 In the phosphine compound, oxidation reduces the acceptance by the phosphine (as it reduces the acceptance by the carbonyls), and there is weaker Cr–P bonding. Therefore, the Cr–P bond is longer in the product (by 0.094 Å). In the ammine compound, the NH3 is not a acceptor; it is a donor only. Because oxidation increases the positive charge on the metal, there is stronger attraction between the metal and the sigma-donating NH3, shortening the Cr–N bond (by 0.050 Å). 2– 10.30 The structure of [ReH9] (Figure 9.35) has a D3h point group. D3h A1 E A2 E 2C3 0 1 –1 1 –1 E 9 1 2 1 2 h 3 1 2 –1 –2 3C2 1 1 0 –1 0 2S3 0 1 –1 –1 1 3v 3 1 0 1 0 x2 + y2, z2 (x, y) (x2 – y2, xy) z (xz, yz) The representation reduces to 2A1 + 2E + A2 + EThese representations match atomic orbitals on Re as follows: A1 E (px, py), (dx2– y2, dxy) A2 pz E 10.31 a. s, dz2 (dxz, dyz) Group orbitals: H b. H Group orbital–central atom interactions: d z2 H s H c. Fe Fe pz H H The hydrogen orbitals (potential energy = –13.6 eV) are likely to interact more strongly with iron orbitals that have a better energy match, the 3d (–11.7 eV) and 4s (–7.9 eV). The 3p orbitals of iron are likely to have very low energy (ca. –30 eV) and not to interact significantly with the hydrogen orbitals. (Potential energies for transition metal orbitals can be found in J. B. Mann, T. L. Meek, E. T. Knight, J. F. Capitani, and L. C. Allen, J. Am. Chem. Soc., 2000, 122, 5132.) Copyright © 2014 Pearson Education, Inc. 156 Chapter 10 Coordination Chemistry II: Bonding 10.32 a. In octahedral geometry the s orbitals on the hydride ligands would generate the representation: Oh A1g Eg T1u E 6 1 2 3 8C3 0 1 –1 0 6C2 0 1 0 –1 3C2(=C42) 2 1 2 –1 6C4 2 1 0 1 i 0 1 2 –3 6S4 0 1 0 –1 8S6 0 1 –1 0 3h 4 1 2 1 6d 2 1 0 1 (2z2–x2–y2, x2–y2) (x, y, z) b. The representation reduces to A1g + Eg + T1u, as shown above. c. The matching orbitals on Cr are: A1g: s ; Eg: dx2-y2, dz2 ; T1u: px, py, pz. The orbital potential energies for the Cr 3d orbitals and H 1s orbitals are -10.75 eV and -13.61 eV, respectively. The energy compatibility of these orbitals suggests the possibility of a relatively robust interaction between the Cr dx2-y2, dz2 orbitals and the group orbitals comprised of hydrogen 1s orbitals. dx2–y2, dz2 o 3d dxy, dxz, dyz s Eg Cr 10.33 a. D4h b. 1. Other group orbitals (A1g, T1u) CrH6 6H Because a deuterium atom has only its 1s electron to participate in bonding, there are only three platinum group orbitals that can potentially interact with deuterium: Group Orbitals: Group Orbital–D Interactions d z2 pz s Pt D Pt Pt D Pt 2. In each case, sigma interactions could occur, as shown above. 3. The strongest interactions are most likely between the dz2 orbitals of Pt and the 1s orbital of D. Lobes of the 5dz2 orbitals point toward the D, and their energy (–10.37 eV) is a good match for the valence orbital potential energy of hydrogen Copyright © 2014 Pearson Education, Inc. Chapter 10 Coordination Chemistry II: Bonding 157 (–13.61 eV). Both bonding and antibonding molecular orbitals would be formed: D Pt D Pt Pt Pt 10.34 MnO4– has no d electrons, while MnO42– has 1. The slight antibonding effect of this electron is enough to lengthen the bonds. In addition, there is less electrostatic attraction by Mn(VI) in – MnO42– than by Mn(VII) in MnO4 . 10.35 The oxidation product is sketched at right, with the Pd—N distances (pm) provided. The longer Pd—N axial distances (243 pm) relative to the equatorial Pd—N distances (202 and 210 pm) illustrate the Jahn-Teller distortion. The reference indicates that the HOMO has extensive character of the Pd 4d 2 orbital. This orbital also features significant sp 3 character at the axial z nitrogen atoms. Using a ligand with more electron-rich axial nitrogen atoms could possibly increase the energy of the HOMO. One proposal would be the introduction of electron-releasing methyl substituents at the (now) methylene carbons the tertiary amine nitrogen atoms. One ligand candidate is below. One challenge in “ligand design” is that modifications of a molecule will sometimes prevent its subsequent metal binding (for example, if steric bulk near the donor atoms increases too much). And the synthesis of new ligands often presents significant research challenges, even when the desired modification appears minor. H 3C CH3 N N N N H 3C CH3 Copyright © 2014 Pearson Education, Inc. N 243 N 210 N 202 Pd Cl CH3 243 N 158 10.36 Chapter 10 Coordination Chemistry II: Bonding The dimeric structure of [MnL1(N 3 )(OCH 3 )]2 is below, with an abbreviated sketch of the Mn(III) coordinate sphere with bond lengths (pm) at right. The reference states that the three donor atoms of the meridionally coordinated Schiff base ligand and the bridging oxygen atom 189 pm from Mn(III) form an equatorial plane in this distorted octahedral geometry. The significantly longer bond length to the other bridging oxygen atom (223 pm) suggests the presence of Jahn-Teller distortion, but the possibility of a trans effect of the azide ligand (Section 12.7) must also be considered. CH3 N O CH3 H O N3 N Mn N O CH3 O H O CH3 Mn 189 O 223 N3 N 189 H3C 199 Mn N 217 O 210 CH3 N N3 10.37. The structure of [Pd(tacn)(Htacn)]3+ is surprising since both non-coordinated nitrogen atoms are positioned on the same side of the plane formed by the four nitrogen atoms bound to the Pd(II) center (syn). This creates more steric hindrance than would be provided by an anti structure. The authors suggest that the orientation of these nitrogen atoms maximizes hydrogen bonding with the nitrate counterions and water in the solid state. An alternate explanation not explicitly stated is an intramolecular hydrogen bond involving both N(1) and N(1) . The palladium-nitrogen core of [Pd(tacn)2 ]3+ is sketched at right (bond lengths in Å with estimated standard deviations), with the slight elongation in the Pd—N(axial) bonds indicating Jahn-Teller distortion. The three unlabeled bonds are equivalent to those trans to them by symmetry; they have the identical bond distances. 10.38 a. Oh A1g Eg T1u E 6 1 2 3 N 2.180(9) N 2.118(9) 2.111(9) N Pd N N N 8C3 0 1 –1 0 6C2 0 1 0 –1 6S4 2 1 0 1 3C2(=C42) 2 1 2 –1 i 0 1 2 –3 6S4 0 1 0 –1 8S6 0 1 –1 0 3h 4 1 2 1 6d 2 1 0 1 b. The representation reduces to A1g + Eg + T1u, as shown above. c. The matching orbitals on Ti are: A1g: s ; Eg: dx2-y2, dz2 ; T1u: px, py, pz Copyright © 2014 Pearson Education, Inc. (2z2–x2–y2, x2–y2) (x, y, z) Chapter 10 Coordination Chemistry II: Bonding 159 d. dx2–y2, dz2 o 3d dxy, dxz, dyz s Eg Ti 10.39 10.40 10.41 [TiH6]2– Other group orbitals (A1g, T1u) 6H a. The t2g molecular orbitals should involve titanium’s dxy, dxz, and dyz orbitals, the the eg orbitals should involve the metal’s dx2-y2, dz2 orbitals. b. A difference from Figure 10.5 that should be noted is that the t2g orbitals of the metal are no longer nonbonding in TiF63– but are involved in the formation of bonding and antibonding molecular orbitals. In comparison with Figure 10.7, the t2g orbitals with greatest metal character should form molecular orbitals that are higher rather than lower in energy in comparison with the dxy, dxz, and dyz orbitals of the free metal ion. a. The iron orbitals matching the representation e are the dx2‐y2 and dz2 ; the orbitals matching t2 are the dxy, dxz, and dyz. The d orbital–ligand orbital interactions should appear more indirect than in the case of an octahedral transition metal complex because the lobes of these orbitals do not point directly toward the ligands. b. The results for FeCl4– should be simpler than shown in Figure 10.19 because the chloro ligand does not have π* orbitals to be taken into consideration (the LUMO orbitals of CO). Results may vary considerably, depending on the software used. It is probably more likely that selecting the π-acceptor ligand (CN–), a -donor, and a π-donor will show a trend matching that in Table 10.12 than if a set of three of -donor or π-donor ligands is used—but examining each type can be instructive. Copyright © 2014 Pearson Education, Inc. 160 Chapter 11 Coordination Chemistry III: Electronic Spectra CHAPTER 11: COORDINATION CHEMISTRY III: ELECTRONIC SPECTRA 11.1 a. p3 There are (6!)/(3!3!) = 20 microstates: MS –3/2 +2 +1 ML 0 1 – –1 –2 – 0 –1 – –1/2 – – + 1 1 0 – – 1+ 1 –1 – – 0+ 1 0 – – + 1 0 –1 – – 1 0+ –1 – – + 1 0 –1 – – –1+ –1 1 – – –1 0 0+ – – –1+ –1 0 1/2 – 1 1 0+ – 1+ 1 –1+ – 1+ 0 0+ – + + 1 0 –1 – 1+ 0 –1+ – 1 0+ –1+ – –1+ –1 1+ – + –1 0 0+ – –1+ –1 0+ 3/2 + 1+ 0+ –1+ Terms: L = 0, S = 3/2: 4S (ground state) L = 2, S = 1/2: 2D L = 1, S = 1/2: 2P b. p1d1 There are 6! 10! 1!5! 1!9! –1 – – 1 2 – – 1 1 – – 0 2 – – 1 0 – – 0 1 – – –1 2 – – 1 –1 – – 0 0 – – –1 1 – – –1 0 – – 0 –1 – – 1 –2 – – –1 –1 – – 0 –2 – – –1 –2 3 2 1 ML 0 –1 –2 –3 Terms: L = 3, S = 1 L = 3, S = 0 L = 2, S = 1 60 microstates: 3 F (ground state) F 3 D 1 MS 0 – + – 1 2 , 1+ 2 – – 1+ 1 , 1 1+ – + – 0 2 , 0+ 2 – – 1+ 0 , 0+ 1 – + – + 1 0, 0 1 – – –1 2+, –1+ 2 – – 1+ –1 , 1 –1+ – – 0 0+, 0+ 0 – – + + –1 1 , –1 1 – – –1+ 0 , 0+ –1 – + – –1 0 , 0 –1+ – – 1 –2+, 1+ –2 + – – –1 –1 , –1 –1+ – – 0+ –2 , 0 –2+ – – –1 –2+, –1+ –2 L = 2, S = 0 L = 1, S = 1 L = 1, S = 0 1 1 2+ 1+ 1+ 0+ 2+ 1+ 0+ 0+ 1+ –1+ 2+ + 1+ –1+ 0+ 0+ –1+ 1+ –1+ 0+ 0+ –1+ 1+ –2+ –1– –1+ 0+ –2+ –1+ –2+ 1 D P 1 P 3 The two electrons have quantum numbers that are independent of each other, because the electrons are in different orbitals. Because they have different l values, the electrons can have the same ml and ms values. Copyright © 2014 Pearson Education, Inc. Chapter 11 Coordination Chemistry III: Electronic Spectra 11.2 For p3: L = 0, S = 3/2, the term 4S has J = 3/2 only (|L+S| = |L–S|). Therefore, the ground state is 4S3/2. For p1d1: L = 3, S = 1, the term 3F has J = 4, 3, 2. Since both levels are less than half filled, the state having lowest J has lowest energy, and the ground state is 3F2. 2! 10! 20 microstates: 11.3 a. s1d1 There are 1!1! 1!9! MS No index entries found. mmmm M –1 0 +1 +2 0– 2– 0 – 2 +, 0 + 2 – 0+ 2+ +1 0– 1– 0 – 1 +, 0 + 1 – 0+ 1+ – – – + + – 0 0 0 0 0, 0 0 0+ 0+ ML –1 0– –1– 0– –1+, 0+ –1– 0+ –1+ – – – + + – –2 0 –2 0 –2 , 0 –2 0+ –2+ b. Terms: L = 2, S = 1: 3D; L = 2, S = 0: 1D The 3D, with the higher spin multiplicity, is the lower energy term. 10! 14! 140 microstates: 11.4 a. d1f1 There are 1!9! 1!13! MS No index entries found. mmmm M –1 0 +5 2– 3– 2– 3+, 2+ 3– 2– 2– 2– 2+, 2+ 2– +4 1– 3– 1– 3+, 1+ 3– – – 2 1 2– 1+, 2+ 1– – – +3 1 2 1– 2+, 1+ 2– – – 0 3 0– 3+, 0+ 3– – – 2 0 2– 0+, 2+ 0– 1– 1– 1– 1+, 1+ 1– +2 – – 0 2 0– 2+, 0+ 2– – – –1 3 –1– 3+, –1+ 3– – – 2– –1+, 2+ –1– 2 –1 – – 1 0 1– 0+, 1+ 0– +1 0– 1– 0– 1+, 0+ 1– – – –1 2 –1– 2+, –1+ 2– – – –2 3 –2– 3+, –2– 3– – – 2 –2 2– –2+, 2+ –2– ML – – 1 –1 1– –1+, 1+ –1– – – 0 0 0 0 – 0 +, 0 + 0 – –1– 1– –1– 1+, –1+ 1– – – –2 2 –2– 2+ , –2+ 2– – – –2 1 –2– 1+, –2+ 1– – – –1 0 –1– 0+, –1+ 0– – – –1 0 –1 0– –1+, 0+ –1– 1– –2– 1– –2+, 1+ –2– – – 2 –3 2– –3+, 2– –3– c. continued Copyright © 2014 Pearson Education, Inc. +1 2+ 3+ 2+ 2+ 1+ 3+ 2+ 1+ 1+ 2+ 0+ 3+ 2+ 0+ 1+ 1+ 0+ 2+ –1+ 3+ 2+ –1+ 1+ 0+ 0+ 1+ –1+ 2+ –2– 3+ 2+ –2+ 1+ –1+ 0+ 0+ –1+ 1+ –2+ 2+ –2+ 1+ –1+ 0+ 0+ –1+ 1+ –2+ 2– –3+ 161 162 Chapter 11 Coordination Chemistry III: Electronic Spectra –2 –3 –4 –5 11.5 –2– 0– –1– –1– 0– –2– 1– –3– –2– –1– –1– –2– 0– –3– –2– –2– –1– –3– –2– –3– –2– 0+ , –2+ 0– –1– –1+, –1+ –1– 0– –2+, 0+ –2– 1– –3+, 1+ –3– –2– –1+, –2+ –1– –1– –2+, –1+ –2– 0– –3+, 0+ –3– –2– –2+, –2+ –2– –1– –3+, –1+ –3– –2– –3+, –2+ –3– –2+ 0+ –1+ –1+ 0+ –2+ 1+ –3+ –2+ –1+ –1+ –2+ 0+ –3+ –2+ –2+ –1+ –3+ –2+ –3+ b. Terms: L = 5, S = 1: 3H; L = 5, S = 0: 1H; L = 4, S = 1: 3G; L = 4, S = 0: 1G; L = 3, S = 1: 3F; L = 3, S = 0: 1F ; L = 2, S = 1: 3D; L = 2, S = 0: 1D; L = 1, S = 1: 3P; L = 1, S = 0: 1P c. The lowest energy term is the 3H. For this term, J has the values 6, 5, and 4. Because the subshells are less than half full, the lowest value of J provides the lowest energy: 3 H4. a. From Problem 11.1a, for a p3 configuration there are three terms: 4S, 2D, and 2P. The J values for each of these are determined below. 2P For 4S: L = 0, S = 3/2 ; Because J = L + S, L + S – 1, …|L – S|, the quantum number J can only be 3/2 and there is a single state for 4S: 4S3/2 For 2D: L = 2, S = 1/2. 28839.31 28838.92 c 2D 19233.18 19224.46 e Possible J values are 5/2 and 3/2, and the two possible states are 2D5/2 and 2D3/2. 4S For 2P: L = 1, S = 1/2. Possible J values are 3/2 and 1/2, with states 2P3/2 and 2P1/2. The lowest energy state is 4S3/2 (highest multiplicity). 2D5/2 and 2D3/2 are next, at 19233.18 and 19224.46 cm–1, and 2P3/2 and 2P1/2 are the highest energy at 28838.92 and 28839.31 cm–1. b. The difference in energy between the 4S and 2D states is 2e. From the average of the two nearly degenerate 2D states, e = 9614.41 cm–1. The difference in energy between the averages of the 2D and 2P states is c = 28839.12 – 19228.82 = 9610.30 cm–1. Copyright © 2014 Pearson Education, Inc. 0 Chapter 11 Coordination Chemistry III: Electronic Spectra 11.6 a. s1f1 There are 2! 14! 1!1! 1!13! 28 microstates: MS 0 –1 11.8 + – 2 – 1 – 0 – 0 3, 0 3 0 – 0 2+, – 0+ 2 – 0+ 2+ – 0 1+, – 0+ 1 – 0+ 1+ – 0 – 0 +, 0 + 0 – 0+ 0+ – 0+ –1 – 0+ –1+ + + 3+ 0 2 0 1 0 0 0 –1 0 –1 – – 0 –1+, –2 0 –2 – – 0 –2+, 0+ –2 – – 0+ –2+ –3 0 –3 – – 0 –3+, 0+ –3 – – 0+ –3+ 3 F (ground state) F b. Terms: L = 3, S = 1 L = 3, S = 0 c. The 3F term, with the higher spin multiplicity, has the lower energy. This term has J = 2, 3, 4; the lowest energy term, including J, is 3F2. a. 2 D has L = 2 and S = 1/2, so ML = –2, –1, 0, 1, 2 and MS = –1/2, 1/2 b. 3 G has L = 4 and S = 1, so ML = –4, –3, –2, –1, 0, 1, 2, 3, 4 and MS = –1, 0, 1 c. 4 F has L = 3 and S = 3/2, so ML = –3, –2, –1, 0, 1, 2, 3 and MS = –3/2, –1/2, 1/2, 3/2 a. 2 D with J = 5/2, 3/2 fits an excited state of d3, 2D3/2 b. 3 G with J = 5, 4, 3 fits an excited state of d4, 3G3 c. 4 F with J = 9/2, 7/2, 5/2, 3/2 fits the ground state of d7, 4F9/2 1 L A 0.10 l 1.00cm mol cm A mol 0.10 A lc; c 2.6 L L l 0.038 1.00cm mol cm 0.038 11.9 11.10 3 – 1 – 3 ML 11.7 – a. __ 24,900 cm–1 1 1 24,900 cm –1 4.02 10 –5 cm = 402 nm Copyright © 2014 Pearson Education, Inc. 163 164 Chapter 11 Coordination Chemistry III: Electronic Spectra 2.998 10 m s 100 cm m 7.46 10 s c b. 8 –1 –1 14 –1 –5 4.02 10 cm 366 nm 2.998 10 m s 10 nm m 8.19 10 s c 8 –1 366 nm 9 –1 14 –1 E h (6.626 10 –34 Js)(8.19 1014 s –1 ) 5.4310 –19 J 11.11 J values are included in these answers: a. d 8 Oh MS = 1 = S Spin multiplicity = 2 + 1 = 3 Max ML = 2+2+1+1+0+0–1–2 = 3 = L, so F term. b. d 5 Oh high spin MS = 5/2 = S Spin multiplicity = 5 + 1 = 6 Max ML = 2+1+0–1–2 = 0 = L, so S term. J = 5/2 3 J = 4,3,2 6 F4 S5/2 d 5 Oh low spin MS = 1/2 = S Spin multiplicity = 1 + 1 = 2 Max ML = 2+2+1+1+0 = 6 = L, so I term. J = |L ± S| = 11/2, 13/2 2I11/2, 2I13/2 (J is uncertain in this case; the usual rule does not apply because the level is exactly half full) 11.12 c. d 4 Td MS = 2 = S Spin multiplicity = 4 + 1=5 Max ML = 2+1+0–1 = 2 = L, so D term. d. d 9 D4h MS = 1/2 = S Spin multiplicity = 1+1=2 Max ML = 2+2+1+1+0+0–1–1–2 = 2 = L, so D term. a. [M(H2O)6]2+ There are two possibilities, M = Ni (d8) and M = Ti (d2). The complex [Ni(H2O)6]2+ is well known. However, titanium strongly prefers the 3+ oxidation state, and [Ti(H2O)6]2+ has not been well characterized. A third possibility would be low spin [Cr(H2O)6]2+ —if this complex were low spin. However, stronger field ligands than H2O are necessary for complexes of Cr(II) to be low spin. b. [M(NH3)6]3+ There are several possibilities, for M = Ti (d1), V (d2), Cr (d3), and Co (d6), among first row transition metals that commonly exhibit a 3+ oxidation state. J = 4, 3, 2, 1, 0 5 D0 J = |L ± S| = 5/2, 3/2 2 D5/2 Excitation of the single d electron from the t2g to eg levels in [Ti(NH3)6]3+ leads to asymmetric occupation of the eg, a configuration susceptible to Jahn-Teller distortion (Section 10.5). Like [Ti(H2O)6]3+ (Figure 11.8), [Ti(NH3)6]3+ shows splitting of its absorption band (excitation from 2T2g to 2Eg (Figure 11.11)). Similarly, in [V(NH3)6]3+, excitation of an electron from t2g to eg would give asymmetric occupation of the eg, a configuration that could potentially give rise to distortion and splitting of absorption bands. However, spectra of d3 octahedral complexes typically do not show such splitting (see spectrum of [V(H2O)6]3+ in Figure 11.8), but have broad overlapping bands that dominate the visible spectra. Copyright © 2014 Pearson Education, Inc. Chapter 11 Coordination Chemistry III: Electronic Spectra 165 Excitation of a t2g electron to an eg level in the d3 complex [Cr(NH3)6]3+ leads to asymmetric occupation of the eg, potentially giving rise to band splitting in the absorption spectrum. The low-spin d6 complex [Co(NH3)6]3+ does show splitting consistent with distortion of the excited state resulting from excitation of a t2g electron to the previously empty eg level. Such splitting is also observed in the d6 iron (II) complex [Fe(H2O)6]2+ (Figure 11.8). The last three transition metals in this row, Ni, Cu, and Zn, do not exhibit stable 3+ complexes of formula [M(NH3)6]3+. Comparable complexes for Mn and Fe have not been well characterized. c. [M(H2O)6]2+ If M = Zn, all t2g and eg orbitals are filled, and consequently no d–d transitions are possible. As a consequence, the d10 complex [Zn(H2O)6]2+ is colorless. The d5 complex [Mn(H2O)6]2+ is nearly colorless (very pale pink) because it has no excited state of the same spin multiplicity (6) as the ground state (see Tanabe-Sugano diagram in Figure 11.7). The electronic spectrum (Figure 11.8) shows absorption bands approximately two orders of magnitude smaller than for other first row [M(H2O)6]2+ complexes. 11.13 [Ni(H2O)6]2+ For d 8 ions, the energy of the lowest energy band is o, so o = 8,700 cm–1. The bands are split due to Jahn-Teller distortion in the excited state. 11.14 a. [Cr(C2O4)3]3– is Cr(III), d 3. o is equal to the lowest energy band, so o = 17,400 cm–1. b. [Ti(NCS)6]3– is Ti(III), d 1. o is the energy of the single band; o = 18,400 cm–1. The band is split due to Jahn-Teller distortion of the excited state. c. [Ni(en)3]2+ is Ni(II), d 8. The lowest energy band corresponds to o; o = 11,200 cm–1. d. [VF6]3– is V(III), d 2. Following the example on pp. 427 - 428, we find the ratio 2/1 and then o /B: 2/1 = 1.57 at o /B = 26. From the Tanabe-Sugano diagram at o /B = 26, 1: E/B = 24.1 E = 24.1 B = 14,800 cm–1 B = 614 cm–1 –1 2: E/B = 37.0 E = 37.0 B = 23,250 cm B = 628 cm–1 –1 –1 Average B = 621 cm , o = 26 B = 16,100 cm (On the basis of the data in this and following problems, the values of B and o should be rounded to two significant digits; additional digits are shown here to assist in checking calculations.) e. V(III) is a d 2 ion. Again following the example on pp. 427 - 428, 2 = 21,413 cm–1 and 1 = 14,409 cm–1, 2/1 = 1.49. From the Tanabe-Sugano diagram at o/B = 34.5, 1: E/B = 29 E = 29 B =14,409 cm–1 B = 497 cm–1 –1 2: E/B = 44 E = 44 B = 21,413 cm B = 487 cm–1 Average B = 492 cm–1, o = 34.5 B = 16,970 cm–1 11.15 [Co(NH3)6]2+, d 7. As in Problem 11.14d: 2/1 = 2.34 at o/B = 11. From the Tanabe-Sugano diagram at o = 11, 1: E/B = 10 E= 10 B = 9,000 cm–1 B = 900 cm–1 –1 2: E/B = 22.5 E= 22.5 B = 21,100 cm B = 938 cm–1 Average B = 919 cm–1, o = 11 B = 10,100 cm–1 Copyright © 2014 Pearson Education, Inc. 166 Chapter 11 Coordination Chemistry III: Electronic Spectra 11.16 a. t2g4eg2 The t2g level is a triply degenerate asymmetrically occupied state, so it is T. b. t2g6 c. t2g3eg3 This is an excited state, with the t2g level uniformly occupied and the eg level doubly degenerate, so it is E. t2g5 This is a triply degenerate state, T. (Vacancies in the orbitals can be treated similarly to electrons.) d. e. eg This is a nondegenerate state, completely occupied, so it is A. Another excited state, this is doubly degenerate, E. 11.17 The complexes with potential degeneracies are those with d 1, d 2, d 4, d 7 and d 9, low-spin d 5, and high-spin d 6 configurations. The strongest effects are with high-spin d 4, low-spin d 7, and d 9 complexes, corresponding to [Mn(NH3)6]3+, [Ni(NH3)6]3+, and the unknown [Zn(NH3)6]3+. Weaker effects might be seen with [Ti(NH3)6]3+, [V(NH3)6]3+, low-spin [Fe(NH3)6]3+, and highspin [Co(NH3)6]3+, all of which are unknown or unstable. Other ligands are needed to stabilize ions containing these 3+ metal ions. 11.18 The 5d orbitals of Re are higher in energy than the 3d orbitals of Mn, so an LMCT excitation – requires more energy for ReO4 . In addition, since the molecular orbitals derived primarily from – – – the 3d orbitals of MnO4 are lower in energy than the corresponding MO’s of ReO4 , MnO4 is better able to accept electrons; it is a better oxidizing agent. 11.19 The order of energy of the charge transfer bands is I < Br < Cl. LMCT in low-spin d 6 [Co(NH 3 )5X]2+ can be approximated (since this complex does not have Oh symmetry) as excitation into the LUMO (empty eg orbitals of high z 2 and x 2 y 2 character) from lower energy, relatively nonbonding (or weakly bonding -donor) orbitals with high halide valence orbital character. It is reasonable to approximate the LUMO energies as very similar for this series of Co(III) cations. However, the lower-energy orbitals with high halide valence orbital character will be lowest for the most electronegative chloride, and highest for the least electronegative iodide. This difference results in [Co(NH 3 )5I]2+ having the lowest-energy predicted LMCT band. From an HSAB perspective, iodide is softer and can lose an electron most easily in an LMCT process. 11.20 Comparing [Fe(CN)6]3– (low spin d 5) and [Fe(CN)6]4– (low spin d 6), where CN– is a donor and a π acceptor: the t2g orbitals of [Fe(CN)6]3– contain 5 electrons, allowing LMCT from the ligand orbitals to either t2g or eg levels. The t2g levels of [Fe(CN)6]4– are full, so only the higher eg levels are available for LMCT. MLCT transitions (t2g π*) are also possible for either complex. 11.21 a. O2– and Cl– are both and π donors, and the metal ions are Cr(V) and Mo(V) (the ligands are Cl– and O2–). Metal d orbitals, influence of Cl– ligands: z2 2 x – y2 xy xz yz e 2 3 0 0 0 eπ 0 0 4 3 3 Copyright © 2014 Pearson Education, Inc. Total 2e 3e 4eπ 3eπ 3eπ Chapter 11 Coordination Chemistry III: Electronic Spectra 167 Metal d orbitals, influence of O2– ligands: z2 x2 – y2 xy xz yz e 1 0 0 0 0 E 1 1 1 2 2C4 1 –1 –1 0 C4v A1 B1 B2 E eπ 0 0 0 1 1 C2 1 1 1 –2 Total e 0 0 eπ eπ 2v 1 1 –1 0 2d 1 –1 1 0 z2 x – y2 xy (xz, yz) 2 Overall, from lower energy (top) to higher energy (bottom), the d orbital energies are: xy xz and yz z2 x2 – y2 4eπ(Cl) 3eπ(Cl) + eπ(O) 2e(Cl) + e(O) 3e(Cl) b. The symmetry labels of the orbitals are given in the C4v character table. c. The lowest d orbital has an energy of 2eπ(Cl); the next have energies of 2eπ(Cl) + eπ(O). Because these are d 1 complexes, the transitions given provide the HOMO/LUMO gaps in these complexes. The interactions between metal and ligand are generally stronger for a second row transition metal than for the first row, raising the LUMO energy in the Mo case. It is reasonable to expect the larger Mo(V) to offer better overlap with the ligand donor orbitals relative to the smaller Cr(V). 11.22 [V(CO)6]– < Cr(CO)6 < [Mn(CO)6]+ As the nuclear charge on the metal increases, the metal orbitals are drawn to lower energies. Consequently, the MLCT bands should increase in energy. 11.23 a. s 0.65 n n 2 At 80K: s 5.2 n n 2 At 300K: b. 11.24 a. n = 0.19 n = 4.3 The complex is near the low-spin – high-spin boundary of the d 6 Tanabe-Sugano diagram. High spin becomes increasingly favored as the temperature increases. ML2, using positions 1 and 6, with O2– both a and π donor: 2 z x2 – y2 xy xz yz e 2 0 0 0 0 eπ 0 0 0 2 2 Copyright © 2014 Pearson Education, Inc. Total 2e 0 0 2eπ 2eπ 168 Chapter 11 Coordination Chemistry III: Electronic Spectra b. c. 11.25 If this is a high-spin complex, there are 4 electrons in the lowest levels (xy, x2 – y2), 3 in the next two (xz, yz), and 1 in the highest (z2). Electronic transitions can be either from the middle levels to the top, and from the bottom levels to the middle and the top—three possibilities in all. According to the reference, the transitions seen are from the middle and the bottom levels to the top level. 2e 1 2e 2 Assigning the transitions as in part b: E = 2e = 16,000 cm–1 E = 2eπ = 9,000 cm–1 so e = 8,000 cm–1 and eπ = 4,500 cm–1 The reference provides a much more detailed analysis, including a discussion of the paramagnetism of this complex and other factors related to its electronic spectrum. Re(CO)3(P(OPh3))(DBSQ) Re(CO)3(PPh3)(DBSQ) Re(CO)3(NEt3)(DBSQ) 18,250 cm–1 17,300 cm–1 16,670 cm–1 NEt3 is the strongest donor ligand in this series. Therefore, the metal in the complex Re(CO)3(NEt3)(DBSQ) has the greatest concentration of electrons and the greatest tendency for electron transfer to acceptor orbitals. Since this complex also has the lowest-energy charge transfer band, we may assign this as MLCT. 11.26 11.27 a. RuO42– has the highest value of t. t increases with the oxidation state of the metal and in general is greater for second row than for first row metals. The overall trend is RuO42– > FeO42– > MnO43– > CrO44–. b. The nuclear charge of iron is greatest in this isoelectronic series and exerts the strongest attraction for bonding electrons. As a result, FeO42– has the shortest metal-oxygen distance, 165 pm, in comparison with 170 pm for MnO43– and 176 pm for CrO44– and RuO42–. c. As the nuclear charge of the metal increases, the metal orbitals’ energies are decreased (further stabilized). Consequently, less energy is needed to excite electrons from ligand orbitals to these metal orbitals. These are LMCT absorptions. Aqueous solutions of Ni(NO3)2 contain the green [Ni(H2O)6]2+ ion; nitrate is the counterion. Addition of aqueous NH3 replaces the H2O ligands in [Ni(H2O)6]2+ to give blue [Ni(NH3)6]2+. As a bidentate ligand, en can replace two NH3 ligands; three en ligands can therefore replace all six NH3 ligands to form violet [Ni(en)3]2+: 2+ observed color: complementary color: [Ni(H2O)6] green red NH3 2+ [Ni(NH3)6] blue orange en [Ni(en)3]2+ violet yellow The complementary colors in this series have increasing energies, indicating that en has the strongest effect on o, and H2O has the weakest effect. This is consistent with the positions of these ligands in the spectrochemical series. Copyright © 2014 Pearson Education, Inc. Chapter 11 Coordination Chemistry III: Electronic Spectra 11.28 a. 169 These colors are most likely the consequence of LMCT transitions, from orbitals that are primarily from the oxide ligands to orbitals that are primarily from the metal: t2 e d LMCT MO4– M7+ 11.29 – b. In TcO4 , the separation between the donor orbitals of the O2– ligands and the – – acceptor orbitals is greater than in MnO4 . As a consequence, TcO4 absorbs light – of higher energy (green) than MnO4 (yellow). Actually, most of the absorption by – TcO4 is in the ultraviolet, with the pale red color a result of a tail of the absorption band extending into the visible. c. The metal-ligand interactions in MnO42– (Mn(VI)) are weaker than in MnO4 (Mn(VII)), and the separation of donor and acceptor orbitals in MnO42– is smaller, – meaning that less energy (red light) is necessary for excitation than in MnO4 (yellow). 2– Also worth noting: the Mn—O bond distance is longer in MnO4 (165.9 pm) than in – MnO4 (162.9 pm), an indication of weaker bonding in the former. a. At 350 nm: A 2.34 lc 1.00cm 2.00 10 –4 M – –1 2,660 L mol –1 cm –1 At 590 nm: A 0.370 lc 1.00cm 2.00 10 –4 M At 1540 nm: A 0.0016 lc 1.00cm 2.00 10 –4 M –1 1,850 L mol cm b. –1 At 514 nm: A 0.532 lc 1.00cm 2.00 10 –4 M 11,700 L mol cm –1 11.30 4O2– –1 –1 8.0 L mol cm Because of their high intensity, the bands at 350, 514, and 590 nm are probably charge transfer bands. However, the low molar absorptivity of the band at 1540 nm indicates that it is probably a d–d transition (see examples in Figure 11.8). These are all d 8 complexes, with three excited states of the same spin multiplicity as the ground state. For d 8, o = energy of the lowest energy band. B can be calculated by using the method of Problem 11.14. o is the difference between the transitions 3A2 3T1 and 3A2 3T2; a graph of 2/1 versus o/B needs to be prepared for d 8 Ni2+ in order to calculate B, as described on pages 427 - 428. A plot of the ratio of the highest energy band to the lowest energy band is shown on the next page. Copyright © 2014 Pearson Education, Inc. 170 Chapter 11 Coordination Chemistry III: Electronic Spectra Species [Ni(H2O)6]2+ [Ni(NH3)6]2+ [Ni(OS(CH3)2)6]2+ [Ni(dma)6]2+ 11.31 o/B 8.8 12.5 8.4 8.0 Ratio 3.06 2.62 3.11 3.14 o(cm–1) 8500 10500 7728 7576 B(cm–1) 970 840 920 950 a. These are both high-spin d 7 complexes, for which the ground-state term symbol is 4F (see Figure 11.7) b. There are three possible transitions, all originating from the 4T1(4F) and going to the 4 T2, 4A2, and 4T1(4P) levels. c. For [Co(bipy)3]2+, the ratio comes at 2 22,000 cm–1 1.95. From Figure 11.14, this 1 11,300 cm–1 o 17, at which: B for the 22,000 cm–1 band: E 22,000 cm1 28 and B 786 cm1 B 28 for the 11,300 cm–1 band: E 11,300 cm1 15 and B 754 cm1 B 15 Average B = 770 cm–1; o 17; o = 13,100 cm–1 B 4 For a d 7 complex, LFSE = – o –10,500 cm–1 5 o for [Co(NH3)6]2+ was calculated in Problem 11.15; o = 10,100 cm–1. 4 For this d 7 complex, LFSE = – o –8,080 cm–1 5 d. These bands should be broad (see Figure 11.8). e. The molecular orbital energy level diagrams should be similar to Figure 10.5, with 6 electrons in t2g orbitals and 1 electron in an eg level in each case. The separation between t2g and eg orbitals should be larger in the bipy complex. Copyright © 2014 Pearson Education, Inc. Chapter 11 Coordination Chemistry III: Electronic Spectra 11.32 These absorption bands correspond to LMCT transitions. As the nuclear charge increases (Fe Ni), the acceptor (largely d) orbitals decrease in energy, enabling the charge transfer transitions to occur at lower energy in the nickel complex. The target complex structure is at right. The conjugated phenylacetlyene linker is first and foremost a conjugated system to allow electronic communication between the separated metal fragments; the linker is to permit the OC CO OC 11.33 171 Fe CO Fe CO OC S S N delocalization of the [Ru(terpy)2 ]2+ MLCT excited state towards the iron atoms. The reference mentions literature precedence for the effectiveness of phenylacetlyene linkers in increasing covalently attached [Ru(terpy)2 ]2+ excited state lifetimes, advantageous to facilitate effective transfer to the iron-sulfur fragment. The rigidity of the linker permits precise control over the distance between the Fe and Ru centers, another way to modulate the effectiveness of the electronic communication between these metal centers. Finally, from a practical standpoint, the acetylene substituent in an intermediate serves as a convenient functional group in the multi-step synthesis of the complex sketched. The most intense bands in the provided spectrum are 2+ 1- F F F P F F F 1- F F F P N F F N F N Ru N N N assigned to * intraligand transitions of terpyridine (310 nm) and MLCT of the covalently attached [Ru(terpy)2 ]2+ fragment (500 nm). The complex above is unable to induce proton reduction since the intramolecular electron transfer from the [Ru(terpy)2 ]2+ excited state (via MLCT) to the iron-sulfur fragment is not thermodynamically favorable. The authors speculate on various strategies to render this electron transfer favorable, including substitution of the carbon monoxide molecules with different ligands. 11.34 The excellent electron withdrawing ability of the imide function is hypothesized to stabilize the resulting reduced dithiolate diiron complex upon electron transfer from the photogenerated zinc porphyrin excited state. This proposed stabilization is expected to render the electron transfer from the zinc photosensitizer more thermodynamically favorable. The well-established utility of the imide functional group for the coupling reactions necessary to synthesize the complex is also advantageous. The lack of ground state electronic interaction between the naphthalene monoimide (NMI) dithiolate-tethered zinc porphyrin and the iron centers was inferred by infrared spectroscopy. The infrared spectrum in the carbonyl region of the NMI ditholate diiron complex is essentially unchanged upon covalent attachment of the zinc porphyrin. The presence of this zinc substituent has an insignificant impact on the electron density at the iron centers. Copyright © 2014 Pearson Education, Inc. 172 Chapter 12 Coordination Chemistry IV: Reactions and Mechanisms CHAPTER 12: COORDINATION CHEMISTRY IV: REACTIONS AND MECHANISMS 12.1 [Cr(H2O)6]2+ is labile, and has 4 unpaired electrons, with 1 in the anti-bonding eg orbital. Occupation of this orbital renders substitution easier by leading to relatively weak chromium(II)– aqua ligand bonds. [Cr(CN)6]4– is inert. It has all 4 metal valence electrons in the bonding t2g levels. These orbitals are rendered bonding in character due to π-backbonding with the cyanide ligands. 12.2 If the rate constants for different entering ligands are significantly different, it suggests that increasing the coordination number is important in the rate-determining step. A necessary assumption to make this assertion is that the mechanism via which the complex undergoes substitution does not vary with the entering ligand. 12.3 Pentachlorooxochromate(V) is a d 1 complex; it should be labile, with vacancies in the t2g levels. Hexaiodomanganate(IV) is a d 3 complex; it should be inert. Hexacyanoferrate(III) is a low-spin d 5 complex; it should be inert (vacant eg levels). Hexammineiron(III) is a d 6 high-spin complex; with partly occupied orbitals in both levels; it should be labile. 12.4 The [Fe(CN)6]4– ion is a low-spin d6 complex, with a maximum LFSE of –2.4 o. It is a notably kinetically inert complex, hence its low reactivity towards ligand substitution that would release the potentially toxic cyanide. 12.5 [Fe(H2O)6]3+ and [Co(H2O)6]2+ are high-spin species; the electrons in the upper eg levels render them labile. [Cr(CN)6]4– is a d 4 low-spin species. The t2g levels are unequally occupied and the eg are vacant, which makes it a borderline complex in terms of substitution rate. [Cr(CN)6]3–, [Fe(CN)6]4–, and [Co(NH3)5(H2O)]3+ are low-spin species with the t2g levels either half filled or completely filled. This, combined with empty eg levels, indicates inert species. LFAE approximations suggest that the activation energies for substitution reactions with these ions are relatively large. 12.6 The general rate law for square planar substitution is: Rate = (k1 + k2[Cl–]) [[Pt(NH3)4]2+]. The general procedure is to measure either the disappearance of [Pt(NH3)4]2+ or the formation of [Pt(NH3)3Cl]+ to find the reaction rate. The most convenient method is by UV or visible absorbance spectra, using a wavelength where there is a large difference in absorbance between the two species. One experimental strategy would be to attempt to establish pseudo-first order conditions, by using a significantly large chloride concentration, so that its concentration can be assumed to not change appreciably while the reaction rate is measured. This may make the reaction appear first order in Pt reactant. Measure the change in reactant and product concentrations using spectroscopy. A plot ln([Pt(NH3)42+]) versus time will afford a straight line with slope kobs if the pseudo-first order approximation is valid when excess chloride is used. Assuming this approximation holds, repeat several times with different (but all sufficiently high) chloride concentrations. A graph kobs vs. [Cl–] will result in a straight line with intercept = k1 and slope = k2. Copyright © 2014 Pearson Education, Inc. Chapter 12 Coordination Chemistry IV: Reactions and Mechanisms 12.7 173 a. DMSO exchange is extremely fast since this is present in large excess as the solvent. b. Saturation kinetics occurs with these entering ligands, as increasing [X ] results in apparent plateaus of the rate constant. The largest of these rate constants (for nitrate) is slightly more than an order of magnitude larger than that of thiocyanate, suggesting the incoming ligand may play a significant role in influencing the rate. The limiting rate constant at high [X ] suggests that either a D mechanism, I d mechanism (if we assume that these substitution reactions are irreversible), or one involving a preassociation complex are viable options. c. Application of a steady-state approximation for a possible D mechanism provides a rate 2+ law of Rate = kobs Co(en)2 (NO 2 )(DMSO) where kobs k1 when the concentration of entering ligand is high. A vital issue is whether these k1 values are sufficiently similar when the chloride, nitrate, and thiocyanate trials are compared to justify assignment of a D mechanism. Ideally, k1 would be expected to be independent of the incoming ligand if a D mechanism were operative. For the possibility of assessing whether an I d mechanism is operative with these data, it is necessary to assume that the substitution is irreversible. In this case, when the concentration of entering ligand is high, pseudo-first order conditions can be assumed, and the rate law simplifies to 2+ Rate = kobs Co(en)2 (NO 2 )(DMSO) where kobs k1[X ] . d. The similarity in the k1 values when those of chloride, nitrate, and thiocyanate are compared for both D (where kobs k1 ) and I d possibilities (where kobs k1[X ] assuming the substitution reaction is irreversible) suggests that these mechanisms are both viable possibilities. However, the variation in these rate constants with different entering ligands indicates that the entering ligand likely plays some role in the rate determining step, rendering an I d mechanism as perhaps a better hypothesis. The possibility of a preassociation complex mechanism could also be argued, but more data is needed to properly evaluate this possibility. 12.8 a. Because the rate is independent of the concentration of 13CO, the rate determining step is most likely: Cr(12CO)6 Cr(12CO)5 + 12CO Cr(12CO)5 then reacts rapidly with 13CO. b. These terms describe two pathways to product, a dissociative pathway (as in part a) and an associative pathway: Dissociative: Cr(CO)6 Cr(CO)5 + CO Cr(CO)5 + PR3 Cr(CO)5(PR3) Associative: Cr(CO)6 + PR3 (slow) (fast) Cr(CO)5(PR3) + CO Copyright © 2014 Pearson Education, Inc. rate = k1[Cr(CO)6] rate = k2[Cr(CO)6][PR3] 174 Chapter 12 Coordination Chemistry IV: Reactions and Mechanisms With two mechanistic pathways leading to the same product, the overall rate is the sum of the rates of both. c. Bulky ligands will tend to favor the first order (dissociative pathway) because the crowding around the metal will favor dissociation and hinder association with incoming ligands. (This effect is discussed further in Section 14.1.1) 12.9 [Cu(H2O)6]2+ is a d 9 complex, subject to Jahn-Teller distortion. Therefore, different rates are observed for exchange of axial and equatorial water molecules. 12.10 These exchange reactions are interesting since the reactants and products have the same structures and energies. The hypothesized mechanism for [Li(H 2O)4 ]+ water exchange is shown below. In A, a water molecule in the second coordination sphere interacts weakly with the lithium center and one aqua ligand via hydrogen bonding. This solvated [Li(H 2O)4 (H 2O)]+ species is more thermodynamically stable than separated [Li(H 2O)4 ]+ and H2O. In the transition state B, the incoming water molecule approaches the metal center and pushes two aqua ligands towards the axial positions of the approximately trigonal bipyramidal intermediate C, with the incoming H2O in an equatorial site. Calculations suggest that the molecular volume of intermediate C is less than that of B, supporting a limiting associative mechanism (A). H 2O H 2O Li OH2 OH2 OH2 O H H 2O H H2O Li OH2 OH2 A B O H 2O H H Li H 2O O H H OH2 C + The mechanism for ammonia ligand exchange for [Li(NH 3 )4 ] is proposed to follow an alternate mechanism that also begins with a solvated [Li(NH 3 )4 (NH 3 )]+ that is more thermodynamically stable than separated [Li(NH 3 )4 ]+ and NH 3 . In the lowest energy solvated [Li(NH 3 )4 ]+ species (D), the NH 3 molecule is proposed to interact simultaneously with N—H bonds of two ammine ligands. Transition state E is approximately trigonal bipyramidal with the entering and leaving ammine ligands positioned in axial sites. An associative interchange mechanism ( I a ) is inferred on the basis of molecular volume calculations that indicate that the volume of E is nearly the same as that of D (the bonds in E are long). NH3 NH3 H3 N Li Li NH3 NH3 H 3N NH3 D H3 N N H H H Copyright © 2014 Pearson Education, Inc. NH3 E Chapter 12 Coordination Chemistry IV: Reactions and Mechanisms 12.11 175 A plot of kobs versus [As(C6H5)3] exhibits a linear relationship, with intercept k1 = 2.3 × 10–5 s–1 and slope k2 = 2.06 × 10–5 M–1s–1. Rate = (k1 + k2[As(C6H5)3])[Co(NO)(CO)3] This reaction shows both first and second order kinetics; the ligand substitution likely occurs via two mechanisms. The first order reaction appears to be a dissociative reaction or a solventassisted dissociation of CO, followed by a fast addition of As(C6H5)3. The other path shows first order dependence on As(C6H5)3, perhaps caused by an associative reaction. 12.12 HNP is a measure of basicity, with more basic molecules having a smaller HNP. The reactions appear to be associative, with the rate increasing with the basicity of the incoming ligand. The two lines are a consequence of the different natures of the ligands; numbers 111 are all phosphorus ligands, while 12-14 are organic nitrogen compounds. The half neutralization potential provides a relative measure of basicity within compounds of similar structure, but is not an absolute measure of the reactivity of the compounds. 12.13 a. G° = H° – TS° = 10300 – 298 × 55.6 = –6300 J/mol = –6.3 kJ/mol G° = –RT lnK; ln K= – G°/RT = 6300/(8.3145 × 298.15) = 2.54; K = 12.7 b. The cis isomer has the higher bond energy (actually, the lower overall energy and collectively stronger bonding), since rearrangement to the trans isomer is endothermic. Since the phosphines are better π acceptors, the cis isomer should be the more stable. If the phosphines are mutually trans, they compete for overlap with the same d orbitals, resulting in weaker bonds to Pt(II). When these phosphines are mutually cis, each can use one of the pair of dxz and dyz orbitals and avoid competition. c. The free phosphine must aid the isomerization via an associative mechanism. Since benzene is the solvent, and it is very nonpolar, it is not likely to assist the reaction. Only the second term of the rate equations of Problem 12.6 is significant here, with phosphine playing the role of the entering ligand. Copyright © 2014 Pearson Education, Inc. 176 12.14 12.15 Chapter 12 Coordination Chemistry IV: Reactions and Mechanisms Two factors need to be considered: 1. The best π acceptors (CO, PPh3) slow the dissociation of CO in the cis position; π donors (halides) speed the dissociation. 2. Halides are weaker donors relative to phosphines and carbon monoxide. As a result, there is less electron density available for -back-bonding in the complexes with halide ligands. This results in weaker M—CO bonds in these complexes, which lowers the activation energy for CO dissociation by raising the electronic ground state of the starting complex. The -donor ability of these halides also assists in stabilizing the squarepyramidal transition state, further contributing to a lowering of the activation energy for CO dissociation. Trans-Pt(NH3)2Cl2 reacts with thiourea (tu) to form trans-[Pt(NH3)2(tu)2]2+; the first Cl– is displaced and the strong trans effect of tu leads to preferential replacement of the second Cl–. Both chlorides of cis-Pt(NH3)2Cl2 are initially replaced. The tu’s are then trans to the NH3’s, which are replaced because of the strong trans effect of tu, resulting in [Pt(tu)4]2+. 12.16 12.17 a. [Pt(CO)Cl3]– + NH3 trans-[Pt(CO)(NH3)Cl2] CO is the stronger trans director b. [Pt(NH3)Br3]– + NH3 cis-[Pt(NH3)2Br2] Br is the stronger trans director c. [Pt(C2H4)Cl3]– + NH3 trans-[Pt(C2H4)(NH3)Cl2] C2H4 is the stronger trans director. a. Two sets of reactions, with examples from Figure 12.13 identified: [PtCl4]2– + 2NH3 cis-[PtCl2(NH3)2] + 2 Cl– (b) cis-[PtCl2(NH3)2] + 2 py cis-[Pt(py)2(NH3)2]2+ + 2 Cl– (h) cis-[Pt(py)2(NH3)2]2+ + 2 NO2– trans-[Pt(NO2)2(NH3)(py)] + NH3 + py (e) trans-[Pt(NO2)2(NH3)(py)] + CH3NH2 py NH2CH3 Pt O 2N + + NO2- NH3 (g) (Continued on next page) Copyright © 2014 Pearson Education, Inc. Chapter 12 Coordination Chemistry IV: Reactions and Mechanisms 177 [PtCl4]2– + 2 py cis-[PtCl2(py)2] + 2 Cl– (b) cis-[PtCl2(py)2] + 2 CH3NH2 cis-[Pt(CH3NH2)2(py)2]2+ + 2 Cl– (h) cis-[Pt(CH3NH2)2(py)2]2+ + 2 NO2– trans-[Pt(NO2)2(CH3NH2)(py)] + py + CH3NH2 (c) trans-[Pt(NO2)2(CH3NH2)(py)] + NH3 O2 N NH2CH3 + Pt + py NO2- NH3 (a) b. Reaction with Cl2 puts Cl both above and below the plane, with the products + Cl H 3N H3 N NH2CH3 and Pt py + Cl NH2CH3 Pt py O2N NO2 Cl Cl Reaction with one mole of Br– replaces one Cl–, with the products + Br H 3N NH2CH3 Pt py NO2 Cl 12.18 12.19 + Br H3 N and NH2CH3 Pt py O2N Cl a. The large negative entropy of activation implies that the activated complex is much more ordered than the reactants. This suggests an associative pathway. b. The iodo ligand leads to the fastest rate, involving bond breaking trans to the halogen, and therefore has the strongest trans effect. V2+ is d 3, and likely to be inert. V3+ is d 2, and labile. At low [H+] (slightly basic conditions), an equilibrium mixture of [V(H 2O)6 ]3+ and [V(H 2O)5OH]2+ may persist. The hydroxo ligand of [V(H 2O)5OH]2+ may permit a bridging interaction with labile [V(H 2O)6 ]2+ to form [V(H 2O)5 ( -OH)V(H 2O)5 ]4+ to facilitate an inner sphere electron transfer mechanism. A proposed rate law is below, with a and b determining the relative rates of the two paths, where term a is associated with an outer sphere electron transfer and term b is associated with an inner sphere pathway. In path a, the bimolecular rate-determining step is the outer sphere electron Copyright © 2014 Pearson Education, Inc. 178 Chapter 12 Coordination Chemistry IV: Reactions and Mechanisms transfer while in path b the rate-determining step is the bimolecular formation of the bridging hydroxo complex. Rate = (k1 + k2/[H+])[V2+][V3+], so a = k1 [V2+][V3+] and b = k2[V2+][V3+] 12.20 [Cr(H2O)6]2+ is labile, but [Co(NH3)6]3+ is inert, and the NH3 ligands are not well equipped to bridge two metals. Therefore, the reaction is likely to be an outer-sphere reaction. 12.21 This reaction is inner sphere, so the X– serves as a bridging ligand and a conduit for the electron – – – – – transfer. The order of rates is Br > Cl > N3 > F > NCS . The difference in reaction rate is most strongly correlated to the varying abilities of these bridging ligands to accommodate – electron transfer between the metal centers. Br works best, and this could be rationalized by – estimating that the hard-soft compatibility of Cr2+ and Br is optimum among these ligands, leading to a relatively robust bridging interaction. Chloride is slightly harder, leading to a slightly – lower electron transfer rate. F is apparently too hard; the high effective nuclear charge its – valence electrons experience render it a poor conduit for electron transfer to Cr2+. The NCS ligand offers a very slow electron transfer, the Cr—NCS—Cr bridge features one interaction with better HSAB compatibility than the other; the resulting bridge must not be ideal for electron transfer despite the presence of a conjugated -system that can lower the activation barrier for electron transfer. Azide also has a conjugated system, but the Cr-NNN-Cr interactions apparently do not have an optimal HSAB match (nitrogen is too hard) to permit a suitable interaction. It is somewhat surprising that azide results in a faster electron transfer rate relative to thiocyanate on the basis of solely HASB arguments. The data in this problem are from D.L. Ball and E. L. King, J. Am. Chem. Soc., 1958, 80, 1091. 12.22 a. NSe has the stronger trans effect. The longer Os—N1 distance is consistent with weakening of this bond by the ligand trans to it, the NSe ligand. b. The short bond distance and large Os—N—Se angle (164.7°) suggest that the ligand has NSe+ character; the cation NSe+ would have a bond order of 3 and a very short bondbond distance. It is also worth noting that the N–Se stretching vibration in this complex is much higher (by 211 cm-1) than in gas phase NSe. a. The more rapid reactivity of the Mn complex is consistent with the general observation that first row transition metal complexes are generally more substitutionally labile than second and third row complexes. b. The negative volume of activation is consistent with an A (or Ia) mechanism. c. Occurrence of two infrared bands is more consistent with a fac isomer; a mer isomer is predicted by symmetry to have three IR-active bands: 12.23 mer isomer (C2v): C2v A1 B1 E 3 1 1 C2 1 1 –1 (xz) 3 1 1 (yz) 1 1 –1 Copyright © 2014 Pearson Education, Inc. z x Chapter 12 Coordination Chemistry IV: Reactions and Mechanisms 179 The representation reduces to 2 A1 + B1, all IR-active. Three IR-active carbonyl bands are expected. fac isomer (C3v): C3v A1 E E 3 1 2 2 C3 0 1 –1 3v 1 1 0 z (x, y) The representation reduces to A1 + E, both IR-active. Two IR-active carbonyl bands are expected. 12.24 These data support ring-closing reactions with dissociative activation. The uniformly positive S act values suggest transition states in the rate determining steps where cyanide ligand dissociation is more prominent than coordination of the pyridine-carboxylate nitrogen atom to the Mn(V) center. The similarity in the rate constants (the average k1 is 1.03 103 M 1s1 with a standard deviation of only 1.3 104 M 1s1 ) is remarkable. These rate constants appear independent of the different nitrogen atom basicities and the charges of these pyridinecarboxylates. Varying these properties of the entering ligand has very little impact on the speed of cyanide substitution. This is also consistent with dissociative activation. 12.25 a. The consumption rates of U 4+ and [PtCl6 ]2- were monitored with UV-Vis spectroscopy, by monitoring the absorptions of U 4+ ( max 648 nm) and [PtCl6 ]2- ( max 300 nm) with time. Reaction orders with respect to both of these ions were determined to be 1 on the basis of the linear relationships of the plots of ln( A At ) versus time, where A and At are the absorbances at time t and infinite time for U 4+ and [PtCl6 ]2- , respectively. These plots exhibited linear relationships for more than three half-lives. b. One hypothesis is that [U(H 2O) n ]4+ initially forms a H 2- complex with [PtCl6 ] that subsequently loses two protons to yield a neutral intermediate where O PtCl6 (H2O)n-2U O [PtCl6 ]2- is coordinated to [U(H 2O) n-2 (OH)2 ]2+ via H hydroxide bridges (shown at right). The inner-sphere electron transfer of two electrons from U(IV) to Pt(IV) is proposed to proceed via this complex. The activation barrier for electron transfer via this intermediate may be reduced (leading to a faster reaction) since the U fragment closely resembles the ultimate uranium(VI)-containing product [(H 2O) n-2 UO 2 ]2+ . Copyright © 2014 Pearson Education, Inc. 180 12.26 Chapter 12 Coordination Chemistry IV: Reactions and Mechanisms a. These complexes have been established as outer-sphere reactants in previous work. b. Using the reaction of [Co(edta)] and [Fe(CN)6 ]4 as an example, the rate law is proposed as Rate k1 [Fe(CN)6 ]4 [Co(edta)] , first order with respect to each complex. To create a graph such as that shown in Figure 3 of the reference, a series of reactions were conducted with varying, but always excess [Fe(CN)6 ]4 and a smaller fixed concentration of [Co(edta)] . The concentration of [Fe(CN)6 ]4 is assumed to not change significantly during these trials, allowing the rate law to be expressed as Rate kobs [Co(edta)] where kobs k1 [Fe(CN)6 ]4 . The kobs rate constants were determined and then plotted versus the [Fe(CN)6 ]4 concentration for each trial. The linear relationship and the zero y-intercept support the validity of the pseudo-first order conditions, and that the overall bimolecular rate law is operative in these electron transfer reactions. The relationship between kobs and k1 is kobs k1 [Fe(CN)6 ]4 , and the slope of the plot of kobs versus [Fe(CN)6 ]4 provides the bimolecular rate constant . This is a beautiful example of the application of pseudo-first order conditions. k1 c. The hypothesis is that the reactants become compartmentalized, hindering their ability to encounter each other for electron transfer, in the presence of reverse micelle (RM) microemulsions. [Co(edta)] is believed to penetrate into the interfacial region of the RM, while [Fe(CN)6 ]4 access to this region is limited, confining [Fe(CN)6 ]4 to the aqueous solution. The physical separation of these complexes may be the dominant factor in hindering the electron transfer reaction, but the possibility that the reduction potentials of these complexes may be changed (for example, that [Co(edta)] may be rendered a weaker oxidizing agent when embedded in the interfacial region of the RM) is also mentioned. Copyright © 2014 Pearson Education, Inc. Chapter 13 Organometallic Chemistry 181 CHAPTER 13: ORGANOMETALLIC CHEMISTRY 13.1 Method A: = Donor-Pair Method, Method B: = Neutral-Ligand Method a. 13.2 Fe(CO)5 A: 8 + 5×2 = 18 + B: 8 + 5×2 = 18 b. [Rh(bipy)2Cl] A: 7 + 2×4 + 2 = 17 B: 9 + 2×4 + 1 – 1 = 17 c. (5-Cp*)Re(=O)3 A: 6 + 0 + 3×4 = 18 B: 5 + 7 + 3×2 = 18 d. Re(PPh3)2Cl2N A: 2 + 2×2 + 2×2 + 6 = 16 B: 7 + 2×2 + 2×1 + 3 = 16 e. Os(CO)(CPh)(PPh3)2Cl A: 7 + 2 + 3 + 2×2 + 2 = 18 B: 8 + 2 + 3 + 2×2 + 1 = 18 f. Ru(CE)Cl2 (NC5H 4 NMe 2 )2 (C3H 4 N2 (C8 H9 )2 ) A: 6 2 2 2 2 2 2 18 B: 8 2 2 1 2 2 2 18 All of these compounds have 16-electron valence configurations. a. Ir(CO)Cl(PPh3)2 A: 8 + 2 + 2 + 2×2 = 16 B: 9 + 2 + 1 + 2×2 = b. RhCl(PPh3)3 A: 8 + 2 + 3×2 = 16 B: 9 + 1 + 3×2 = 16 c. [Ni(CN)4] 2– A: 8 + 4×2 = 16 B: 10 + 4×1 + 2 = 16 d. cis-PtCl2(NH3)2 A: 8 + 2×2 + 2×2 = 16 a. [M(CO)7]+ A: 18 – 7×2 = 4 = M+, V b. H3CM(CO)5 A: 18 – 2 – 5×2 = 6 = M+, Mn B: 18 – 1 – 5×2 = 7 = M, Mn c. M(CO)2(CS)(PPh3)Br A: 18 – 2×2 – 2 – 2 – 2 = 8 = M+, Co B: 18 – 2×2 – 2 – 2 – 1 = 9 = M, Co d. [3-C3H3)(5-C5H5)M(CO)]– A: 18 – 4 – 6 – 2 = 6 = M+, Mn – B: 18 – 3 – 5 – 2 = 8 = M , Mn e. (OC)5M=C(OCH3)C6H5 A and B: 18 – 5×2 – 2 = 6 = M, Cr f. [4-C4H4)(5-C5H5)M]+ g. (3-C3H5)(5-C5H5)M(CH3)(NO) A: 18 – 2 – 6 – 2 – 2 = 6 = M, Cr B: 18 – 3 – 5 – 1 – 3 = 6 = M, Cr 16 B: 10 + 2×2 + 2×1 = 16 13.3 A: 18 – 4 – 6 = 8 = M2+, Ni Copyright © 2014 Pearson Education, Inc. B: 18 – 7×2 = 4 = M+, V B: 18 – 4 – 5 = 9 = M+, Ni Chapter 13 Organometallic Chemistry 182 13.4 13.5 h. [M(CO)4I(diphos)]– a. [Fe(CO)2(5-C5H5)]2 A: 7 + 2×2 + 6 = 17, single Fe–Fe B: 8 + 2×2 + 5 = 17, single Fe–Fe b. [Mo(CO)2(5-Cp)]22– A: 5 + 2×2 + 6 + 1 = 16, double Mo=Mo B: 6 + 2×2 + 5 + 1 = 16, double Mo=Mo a. [M(CO)3(NO)]– linear M–N–O: A: 18 – 4×2 – 2 – 4 = 4 = M, Ti B: 18 – 4×2 –1 – 4 = 5 = M–, Ti Calculating for each metal atom: bent M–N–O: 13.6 b. [M(PF3)2(NO)2]+ linear NO: c. [M(CO)4(2–H)]3 As a triangular structure with three M–M bonds: A: 18 – 4×2 – 2 – 2 = 6 = M+, Tc B: 18 – 4×2 – 1 – 2 = 7 = M, Tc d. M(CO)(PMe3)2Cl A: 16 – 2 – 2×2 – 2 = 8 = M+, Rh B: 16 – 2 – 2×2 – 1 = 9 = M, Rh A: 18 – 2×2 – 2×2 = 10 = M–, Rh B: 18 – 2×2 – 2×3 = 8 = M+, Rh Method B works better for calculating overall charge. a. 13.7 A: 18 – 3×2 – 2 – 2 = 8 = M, Ru B: 18 – 3×2 – 3 = 9 = M–, Ru A: 18 – 3×2 – 2 = 10 = M, Pd B: 18 – 3×2 – 1 = 11 = M–, Pd [Co(CO)3]z 9 + 3×2 = 15, z = 3– z b. [Ni(CO)3(NO)] c. [Ru(CO)4(GeMe3)]z 10 + 3×2 + 3 = 19, z = 1+ d. [(3-C3H5)V(CNCH3)5]z 3 + 5 + 5×2 = 18, z = 0 e. [(5-C5H5)Fe(CO)3]z 5 + 8 + 3×2 = 19, z = 1+ f. [(5-C5H5)3Ni3(3–CO)2]z a. [(5-C5H5)W(CO)x]2, assuming a single W–W: A: 6 + 5 + 1 + x × 2 = 18, x = 3 B: 5 + 6 + 1 + x × 2 = 18, x = 3 b. ReBr(CO)x(CO2C2H4) c. [(CO)3Ni–Co(CO)3]z A and B: 3 × 2 + 10 + 2 + 9 + 3×2 – z = 36, z = 3– d. [Ni(NO)3(SiMe3)]z B: 10 + 3×3 + 1 – z = 18, z = 2+ 8 + 4×2 + 1 = 17, z = 1– 3 × 5 + 3×10 + 2×2 = 49, z = 1+, assuming three Ni–Ni bonds; calculating for each Ni: 5 + 10 + 2(2/3) + 2 = 18 1/3; charge per Ni = 1/3+, overall charge = 1+ (Each triply bridging CO can be considered to donate 2 electrons overall, 2/3 electron to each metal.) A: 6 + 2 + x × 2 + 2 = 18, x = 4 B: 7 + 1 + x × 2 + 2 = 18, x = 4 Copyright © 2014 Pearson Education, Inc. Chapter 13 Organometallic Chemistry 13.8 e. a. [(5-C5H5)Mn(CO)x]2 Upper ring: 5-Cp Co b. M: 183 B: 5 + 7 + 2 + x × 2 = 18, x = 2 5 electrons 9 electrons 14 electrons; 4 more needed, so lower ring must be 4 1-C7 ring 2 PEt3 GeR3 1 electron 4 electrons 1 electron M: 6 electrons 7-C7 ring 5-C5 ring 7 electrons 5 electrons 12 electrons 4 more needed to reach 16: Ti 10 more needed to reach 16: Pt O C 13.9 Co (group 9): Ni (group 10): Co CO CO Co O C O C Ni Co C O Ni Cu (group 11): Cu CO Other compounds with different hapticities for cyclopentadienyl ligands may also be possible. 13.10 Figure 10.19 gives an MO diagram for Ni(CO)4 and cites additional references. The HOMOs (t2) are strongly bonding, and there is a large energy gap between the HOMOs and the LUMOs in this 18-electron molecule. 13.11 The energy of stretching vibrations depends on the square root of the force constant divided by the reduced mass (Section 13.4.1). The reduced masses are 14.73 for 16O and 16.41 for 18O, so (14.73/16.41)1/2 = 0.947, and the 18 O complex has vibrational energy of 0.947 ×975 cm–1 = 924 cm–1. The value given in the reference is 926 cm–1. 13.12 Sulfur is less electronegative than oxygen. Therefore, the tungsten in W(S)Cl2(CO)(PMePh2)2 has greater electron density and a greater tendency to participate in -backbonding with CO; a lower energy (CO) is expected (actual value: 1986 cm–1). 13.13 Adding electrons to a carbonyl complex puts more electrons into the back-bonding t2g orbitals. As a result, the V–C bonds in [V(CO)6]– are strengthened and the distance shortened (but the C–O bonds are weakened by having more electrons in the t2g orbitals, that are antibonding with respect to the C–O bonds.). Copyright © 2014 Pearson Education, Inc. Chapter 13 Organometallic Chemistry 184 Sigma Donor 13.14 Pi Acceptor a. M C N b. M C N R R R R R R M PR3 M M S C N PR3 c. 13.15 13.16 M S C N a. If NO is counted as a linear donor, each of these has 18 electrons. The increasing nuclear charge (from formally Cr(0) to Mn(I) to Fe(II)) results in progressively less backbonding to the ligands and an increase in N–O bond order, leading to the higher NO stretching frequencies. The Cr species has an exceptionally low NO stretching frequency. It may have bent NO coordination, rendering it a 16 electron ion formally containing Cr(II). The Fe complex has a rather high NO stretching frequency, suggesting that π-backbonding is not a very large contribution to the electronic ground state of this complex. b. The low energy band indicates bent NO coordination; the higher energy band is from the linear ligand. This complex has one of each, with angles of 138° and 178° (Greenwood and Earnshaw, Chemistry of the Elements, 2nd ed., pp. 450-52). a. The CO2 molecular orbitals are shown in Figure 5.25. b. The π orbitals of 1,3,5-hexatriene are shown in Figure 13.21. c. cyclo-C4H4: Copyright © 2014 Pearson Education, Inc. Chapter 13 Organometallic Chemistry 13.17 d. cyclo-C7H7: a. Using the group theoretical method described in Chapter 4: The four CO ligands are in a square planar arrangement, but the Mo atom is above them. The reducible representation , shown below, can be derived for the C–O stretching vibrations. reduces to A1 + B1 + E, with A1 and E infrared active, so there are two bands visible (the E bands are degenerate). C4v A1 B1 E 2C4 0 1 –1 0 E 4 1 1 2 C2 0 1 1 –2 2v 2 1 1 0 b. x d xy p x, d xz y py , dyz s, p z, dz2 Copyright © 2014 Pearson Education, Inc. 185 2d 0 1 –1 0 z (x, y) Chapter 13 Organometallic Chemistry 186 13.18 13.19 a. From bottom to top: 0-node, left, is A1, 0 node, right is A2 1-node, far left and third from left are E1 1-node, second from left and far right are E1 2-node, far left and third far left are E2 2-node, second from left and far right are E2 b. s, dz2 A 1 c. The matching orbitals are shown in part b; there is no match in the Fe orbitals for the E2 ligand orbitals. a. The π orbitals of benzene are shown in Figure 13.22. b. Group orbitals (dxy, dx2–y2) E2 (dxy, dyz) E1 pz A2 (px, py) E 1 3-Node Cr 11 12 7 8 9 10 3 4 5 6 1 2 2-Node 1-Node 0-Node Copyright © 2014 Pearson Education, Inc. Chapter 13 Organometallic Chemistry c. Matching Cr orbitals: 1 2 3 4 5 6 d. 7 8 9 10 11 12 s, dz2 pz py dyz px dxz dxy none d x2 – y2 none none none Energy level diagram 3-Node 4p 2-Node 4s 1-Node 3d 0-Node Cr 13.20 a. Cr(C6H6 ) 2 Group Orbitals Point group: C2v (in orientation shown) Fe b. The reducible representation is : C2v A1 A2 B1 B2 E 10 1 1 1 1 C2 0 1 1 –1 –1 v (xz) 2 1 –1 1 –1 v (yz) 0 1 –1 –1 1 Copyright © 2014 Pearson Education, Inc. z x y x 2 , y 2, z 2 xy xz yz 187 188 c. Chapter 13 Organometallic Chemistry = 3A1 + 2A2 + 3B1 + 2B2 Representations A1 A2 B1 B2 Matching orbitals of Fe s, pz, dz2 dxy px, dxz py, dyz 13.21 2C3 0 1 1 –1 E 3 1 1 2 C3v A1 A2 E 3v 1 1 –1 0 z (x, y) = A1 + E, both IR active. There are three vibrations, but two are degenerate, so two C–O stretching bands are expected in the IR spectrum. 13.22 Ni(CO)4 = A1 + T2 8C3 1 1 0 E 4 1 3 Td A1 T2 3C2 0 1 –1 6S4 0 1 –1 6d 2 1 1 (x, y, z) A1 is IR inactive, T2 is IR active: 1 band. Cr(CO)6 Oh E 8C3 6C2 6 C4 3 C2 i 6S4 8S6 A1g Eg T1u 6 1 2 3 0 1 –1 0 0 1 0 –1 2 1 0 1 2 1 2 –1 0 1 2 –3 0 1 0 –1 0 1 –1 0 = A1g + Eg + T1u Only T1u is IR active: 1 band Copyright © 2014 Pearson Education, Inc. 3h 4 1 2 1 6d 2 1 0 1 (x, y, z) Chapter 13 Organometallic Chemistry C O O C C O a. O C M C C 13.24 O O 13.23 M C O C O 4C2 0 1 –1 0 0 0 C O C O D4 d E 2S8 2C4 2S83 C2 A1 B2 E1 E2 E3 10 1 1 2 2 2 0 1 –1 2 1 1 0 –2 0 0 1 –1 2 1 1 –2 2 –2 2 0 – 2 189 – 2 0 2 4d 4 1 1 0 0 0 z (x, y) (Rx, Ry) = 2A1 + 2B2 + E1 + E2 + E3 2B2 and E1 are IR active: 3 bands W(CO)5(NCC3H7) W(CO)4(NCC3H7)2, cis W(CO)3(NCC3H7)3, probably fac 2077, 1975, 1938 cm–1 2107, 1898, 1842 cm–1 1910, 1792 cm–1 Pentacarbonyls have three active IR stretches, tetracarbonyls have one (other ligands trans) or four (other ligands cis), and tricarbonyls have two (fac) or three (mer) (see Table 13.10). It appears that two of the tetracarbonyl bands have similar enough energies for their bands to overlap, so only three appear in the spectrum (or one band may be so weak in intensity that it does not appear), but that the nitriles must be cis. The evidence for the fac isomer of the tricarbonyl is not conclusive based on the IR spectrum alone, since again two bands of the mer isomer might overlap. 13.25 b. The stretching energy of the CO trans to the nitrile ligand is lower than that for CO cis due to reduced competition for the π backbonding electrons (CO is a better πacceptor). In the second complex, two are cis and two are trans. In the third complex, all three are trans, so the band energy is lower. In general, the trend is to lower energies as CO is replaced by butyronitrile; butyronitrile is not as effective a π acceptor as CO, so the π-acceptor nature of CO is enhanced as more nitrile ligands are added. c. As more CO ligands are replaced by nitriles, which are stronger at donating electrons to Mo, the remaining CO ligands become stronger π acceptors, and the Mo–C bonds become stronger. In Mo(CO)3(NCC3H7)3 the Mo–C bond is so strong that the complex cannot react further with butyronitrile. a. The bands at lower energy are for the CO ligands. The C–O stretches, which involve a greater change in dipole moment, are more intense than the C–N stretches. Because CO has a slightly greater reduced mass than CN, C–O stretches should occur at lower energies than C–N stretches; the energy of vibrational levels is inversely proportional to Copyright © 2014 Pearson Education, Inc. 190 Chapter 13 Organometallic Chemistry reduced mass . Also, the greater π-acceptor ability of the CO ligand is a contributing factor to the lower energy stretches for this ligand. b. 2– The trans-[Fe(CO)2(CN)4] complex should have a single IR-active C–O stretch (only the antisymmetric stretch is IR active) and a single IR-active C–N stretch (see Table 13.7). The cis complex should have two IR-active C–O stretches (both symmetric and antisymmetric) and four IR-active C–N stretches. In addition to its 3– single carbonyl stretch, [Fe(CO)(CN)5] would be expected to have three IR-active C–N stretches. The correct identifications are therefore: Complex 2– A: cis-[Fe(CO)2(CN)4] 2– B: trans-[Fe(CO)2(CN)4] 3– C: [Fe(CO)(CN)5] Predicted on basis of symmetry C–O stretches C–N stretches 2 1 1 4 1 3 In complexes A and C, some of the C–N stretches are too weak to be seen or the bands overlap; otherwise, the expected numbers of bands and the observed spectra match. 13.26 a. The CO ligand can be in either the axial or the equatorial position on the trigonal bipyramidal complex. These environments are not equivalent, so the two isomers should have different carbonyl stretching bands. b. The CO stretch absorbs at a higher energy in Fe(CO)(PF3)4, implying that the C–O bond is stronger in that compound. If the CO bond is stronger, the Fe–C bond is correspondingly weaker, indicating that the PF3 ligands are better π acceptors and therefore higher in the spectrochemical series than CO. O 13.27 C Et3P Ru PEt3 PEt3 PEt3 C Et3P Ru C C O O PEt3 O One C–O stretch (antisymmetric) (minor isomer) Two C–O stretches (symmetric and antisymmetric) (major isomer) 13.28 From Table 13.10, [Co(CO)3(PPh3)2]+ must be trigonal bipyramidal with all three CO ligands in the equatorial plane, giving rise to a single C–O absorption band. 13.29 By extrapolation of the positions of the C–O bands for [Mn(CO)6]+ (2100 cm–1) and [Fe(CO)6]2+ (2204 cm–1), one might predict a comparable band for [Ir(CO)6]3+ near 2300 cm–1. Actual value: 2254 cm–1. 13.30 a. In these three cations, the high metal oxidation states significantly reduce backbonding. Copyright © 2014 Pearson Education, Inc. Chapter 13 Organometallic Chemistry 191 This effect is greatest for [Hg(CO)2]2+, with a charge of 1+ per CO, and weakest for [Os(CO)6]2+, with a charge of 1/3+ per CO. Therefore, [Os(CO)6]2+ should have the strongest backbonding, weakest C–O bond, and lowest energy carbon-oxygen stretching vibration. The actual energies are listed on the next page. (CO), cm–1 [Hg(CO)2]2+ 2278 [Pt(CO) 4]2+ 2244 [Os(CO) 6]2+ 2190 b. A reducible representation for this ion based on C–O stretching vibrations would have the following characters: D2d E 2S4 C2 2 A1 2 B2 E 6 2 2 2 0 2 –2 0 2 2 2 –2 2C2 0 2 –2 0 2d 4 2 2 0 z (x, y) There should be three carbon–oxygen stretching bands, two of B2 symmetry and one of E symmetry (a degenerate pair). With a charge 2+, this complex should exhibit infrared bands near those of [Fe(CO)6]2+ (2204 cm–1). Actual bands are observed at 2173 cm–1 (E) and at 2187 and 2218 cm–1 (B2). 13.31 a. The representation based on the set of six C–O vibrations in Oh symmetry is: Oh E 6 8C3 0 6C2 0 6C4 2 3C2(=C42) 2 I 0 6S4 0 8S6 0 3h 4 6d 2 This representation reduces to A1g + Eg + T1u (see Exercise 10.4). Of these, A1g and Eg match squared functions and are therefore Raman active. (A1g matches the band at 2015 cm–1 and Eg matches the band at 2119 cm–1.) b. 13.32 J, K, and L are, respectively, the products of substitution of CO by py: Mo(CO)5py, cis-Mo(CO)4py2, and fac-Mo(CO)3py3. As the number of pyridine ligands increases, the strong donation by this ligand increases the backbonding by the CO ligands, resulting in successive lowering in energy of the C–O vibrations. Symmetry analysis of each of these complexes shows more expected Raman-active bands than reported in the article. Additional weak bands can be seen in the Raman spectra (shown in the reference). (5-C5H5)Cr(CO)2(NS) 1962, 2033 cm–1 (5-C5H5)Cr(CO)2(NO) 1955, 2028 cm–1 NS is a stronger π acceptor, so the CO backbonding is reduced in (5-C5H5)Cr(CO)2(NS) and the CO bond is strengthened and has a higher stretching energy. 13.33 The energy for a stretching vibration is proportional to k , where k is the force constant and Copyright © 2014 Pearson Education, Inc. Chapter 13 Organometallic Chemistry 192 µ is the reduced mass (Section 13.4.1). In this case, the reduced masses for the N–S stretches are: 14 N–S: 14.00 32.06 9.745 14.00 32.06 15 N–S: 15.00 32.06 10.22 . 15.00 32.06 Because the energy is inversely proportional to the square root of the reduced mass, we can write: E( 15 N–S) E( 14 N–S) (14 NS) (15 NS) 9.745 0.9765 The expected position of the N–S stretch 10.22 in the 15NS complex is therefore 0.9765 1284 cm–1 = 1254 cm–1. The reported value is 1248 cm–1. 13.34 Because the reduced mass of 13CO is greater than the reduced mass of 12CO, the separation between vibrational energy levels should be less for 13CO, and the 13CO complex should therefore show an infrared band at lower energy than 2199 cm–1. Actual value: 2149 cm–1. 13.35 a. Mo(CO)6 + Ph2 PCH 2PPh 2 b. (5-C5H5)(1-C3H5)Fe(CO)2 H2 C h O C Ph 2 P Mo P Ph 2 C O CO + 2 CO CO (5-C5H5)(3-C3H5)Fe(CO) + CO The allyl ligand can bond in either 1 or 3 fashion. Loss of CO converts the reactant from 18 electrons to 16 electrons; rearrangement of allyl from 1 to 3 returns it to 18 electrons. c. (5-C5Me5)Rh(CO)2 [(5-C5Me5)Rh(CO)]2 + 2 CO A double Rh–Rh bond is needed; CO ligands may be bridging or terminal (the electron count is the same for both modes). d. V(CO)6 + NO V(CO)5(NO) + CO The compound changes from 17 to 18 electrons. e. – [(CO)5WCC6H5]+ + F + F2BOC2H5 W(CO)5=C(C6H5)(OC2H5) + BF3 AlR 3 O O f. C [(5-C5H5)Fe(CO)2]2 + 2 Al(C2H5)3 Fe (See Figure 13.17) C Fe C O O v(CO), cm 13.36 P(t-C4H9)3 –1 1923 (best donor, poorest π acceptor) Copyright © 2014 Pearson Education, Inc. R3 Al C Chapter 13 Organometallic Chemistry P(p-C6H4Me)3 P(p-C6H4F)3 P(C6F5)3 13.37 13.38 13.39 a. Fe(CO)4(PF3) b. [Re(CO)6]+ 193 1965 1984 2004 (best π acceptor) (PF3 is a strong π acceptor) c. Mo(CO)3(PCl3)3 (PCl3 is the best π acceptor among the phosphines) a. Oxidation of these low-spin d6 octahedral complexes causes loss of 1 t2g electron, resulting in a change in LFSE of 0.4 (from –2.4 to –2.0 ). b. The C—O distance decreases. As the oxidation state of Cr increases from 0 to 1+, the CO ligands become weaker pi acceptors, and there is less occupation of π* orbitals in the ligands; consequently, the bonding in these ligands is stronger in the cations than in the neutral complexes (in the reference, the calculated decrease in C—O distances is in the range of 0.015 - 0.020 Å). c. The Cr—P distance increases (calculated increase: 0.094Å) as the phosphine ligand becomes a weaker pi acceptor; there is less backbonding in the cation than in the neutral complex. d. The Cr—N distance decreases. In order from highest to lowest : Mo(CO)4 (F2PCH2CH2PF2) Mo(CO)4 ((C2F5)2PCH2CH2P(C2F5)2) Mo(CO)4 ((C6F5)2PCH2CH2P(C6F5)2) Mo(CO)4 (Ph2PCH2CH2PPh2) Mo(CO)4 (Et2PCH2CH2PEt2) has strongest acceptor diphosphine has strongest donor diphosphine 13.40 Coordinated N2 has a lower stretching energy than free N2. N2 can act as a π acceptor, weakening the N–N bond and lowering the energy of the stretching vibration. The stretching vibration of free N2 is IR inactive; there is no change in dipole moment on stretching. 13.41 At the higher temperature, the –O–CH3 group should rotate rapidly enough to show only a single, average environment, so each CH3 has a single peak. At low temperatures, this rotation can be restricted, and cis and trans isomers result (see Figure 13.45). The larger peaks represent the more prevalent isomer. 13.42 At high temperature, the C5H5 rings undergo rapid 1,2 shifts to give a single average proton signal on the NMR time scale. At sufficiently low temperature, the different environments of the 1 and 5 rings can be seen. The relative intensities are a = 1, b = 2, c = 2, and d = 5. O C d d O C Fe a d b d d c b c 13.43 PF3 is a stronger π acceptor than PCl3. As a result, the chromium in Cr(CO)5(PCl3) has a greater electron density, and CO acts as a stronger π acceptor in this complex. As a result: a. Cr(CO)5(PF3) has the stronger and shorter C–O bonds. Copyright © 2014 Pearson Education, Inc. Chapter 13 Organometallic Chemistry 194 13.44 b. Cr(CO)5(PCl3) has the higher energy Cr–C bands; since CO acts as a better π acceptor in this complex, the Cr–C bond is strengthened. a. [Fe(NO)(mnt)2]– has less electron density on Fe, less backbonding to NO, and a stronger N–O bond with higher stretching frequency. b. (CO)5Cr:NN:Cr(CO)5 N2 acts as a π acceptor toward both metals, significantly weakening the N–N bond. c. Ta=CH2 has a shorter Ta–C bond because it is a double bond. d. CrCCH3 has a triple bond, which is shorter than either of the Cr–C bonds to CO ligands. e. [Fe(CO)4]2- has more π backbonding because Fe has the lowest nuclear charge of the metals in this isoelectronic series; this reduces the CO bonding and the energy of the C–O vibration. Figure 13.51 shows the highest occupied orbital in the four-atom pi system of the ruthenium analogue to the dimanganese complexes in this problem. This orbital, with two nodes between the metals, is shown below. 13.45 Mn C C Mn A similar orbital is involved in the dimanganese complexes. Oxidation of the neutral complex, by removing electrons from this orbital, should weaken and lengthen the carbon–carbon bond (removing electrons from an orbital that is C–C bonding) and strengthen and shorten the manganese–carbon bonds (removing electrons from an orbital that is Mn–C antibonding). 13.46 13.47 a. MX 6 b. ML3X 6 ML 4 X5 c. ML 4 2– ML 4 X 2 d. ML5 X2 e. ML4 X2 f. ML 4 X2+ ML 3X3 a. ML3X 3 3+ MX6 [FeCl3(PPh3)3 ]3+ belongs to an unknown class of iron complexes, and this cation has not been reported. b. ML 2 X3+ MX4L [FeCl3(PPh3)2 ]+ belongs to an unknown class of iron complexes, and this cation has not been reported. Copyright © 2014 Pearson Education, Inc. Chapter 13 Organometallic Chemistry 195 c. MX 2L2+ MX3L d. [FeCl2(PPh 3 )2]+ belongs to an unknown class of iron complexes, and this cation has not been reported. MX 2L2 FeCl 2 (PPh3)2 is a member of a class with less than 1% of known Fe complexes. The crystal structure of FeCl2 (PPh3)2 was reported in 2005 (O. Seewald, U. Flörke, G. Henkel, Acta Cryst., 2005, E61, m1829). 13.48 13.49 13.50 One CO is replaced by 2-butyne. The NMR peaks are due to ethyl CH3 ( = 0.90), ethyl CH2 ( = 1.63), and butyne CH3 ( = 3.16). The = 3.16 peak splits at low temperatures because the two ends of the butyne are not identical; at higher temperatures, they become identical on NMR time scale, perhaps through rotation about the Mo–butyne bond. The single 31P peak indicates identical PEt3 groups, suggesting the isomer shown. The IR indicates that a CO ligand remains on the compound; the molecular weight of the compound shown is 574.2, well within the limits given. O CH3 C Br Mo C Et3 P Br CH3 The analysis fits [(5-C5H5)Fe(–CO)3Fe(5-C5H5)]. The single CO band is consistent with this structure. Using D3h symmetry for the central part of the molecule, a representation based on the bridging carbonyls reduces to A1 (IR-inactive) plus E (IR-active); the observed absorption is for the E vibration. D3h E 2C3 3 C2 A 1 E 3 1 2 0 1 –1 1 1 0 h 3 1 2 2S3 0 1 –1 (5-C5H5) (3-C5H7)Ni There are 5 5 protons, 4 on the two carbons of the second Cp ring that are not bonded to Ni, 2 on the first and third carbons bonded to Ni, and 1 on the center C bonded to Ni. 3v 1 1 0 (x, y) (5) (2) Ni H (1) H H H H 2 H H (4) 2 13.51 Is/Ia = cot (/2) = cot 38° = 1.64 13.52 a. In the 6-C6H6 complex, CO is acting as a stronger π acceptor, as shown in its lower Copyright © 2014 Pearson Education, Inc. PEt 3 C Chapter 13 Organometallic Chemistry 196 energy C–O stretching vibrations. Because the concentration of electrons on Cr is therefore greater in the 6-C6H6 complex, 6-C6H6 is donating more strongly to Cr than is the 6-C4BNH6 ligand, so 6-C4BNH6 is the stronger acceptor. 13.53 b. The longest C–C distance is the bond opposite the B–N bond, as in the resonance structure shown. The nitrogen attracts electrons from its neighboring carbon, which in turn attracts electrons from the next carbon, enhancing the bond strength between these two carbons and resulting in the shortest C–C bond, 1.374 Å. In the π orbitals (see Figure 13.22 for benzene), one of the occupied π orbitals is antibonding with respect to the C–C bond opposite the B–N bond, consistent with this C–C bond being the longest in the molecule. a. Mass of C60: 720 Ir: 193 (most abundant isotope) CO: 28 C9H7: 115 1056 C 60 b. Because the carbonyl stretch decreases by 44 cm–1, there must be a significant decrease in the electron density at Ir. c. The C60 is replaced by PPh3, giving (5-C9H7)Ir(CO)(PPh3). OC C O Cr Spectral data are interpreted in the reference. 13.54 13.55 Ir C O NMe 2 Cu C O PPh3 If CO is liberated from (5-C5H5)Mn(CO)3, it must be replaced by some other ligand if the 18-electron rule is to be maintained. In this case, the new ligand is tetrahydrofuran (THF), a cyclic ether that can act as a sigma donor, and compound Q is (5-C5H5)Mn(CO)2(THF): (5-C5H5)Mn(CO)3 + THF (5-C5H5)Mn(CO)2(THF) + CO Q The NC groups on carbon in H2C(NC)2 can act as donors to transition metals; each has a lone pair on carbon. In the formation of R, the weakly bound THF is replaced by a H2C(NC)2 ligand: H H C H N C C N C H + CpMn(CO)2(THF) Mn C C O O R There are 0.00300 mmol of dimers in the solution. Of the metals in these dimers, one third are Mo and two thirds are W. Copyright © 2014 Pearson Education, Inc. C C 13.56 N N Chapter 13 Organometallic Chemistry 197 The probability that a particular molecule has the formula [CpMo(CO)3]2 = 1/3 × 1/3 = 1/9. Therefore, 1/9 × total moles = 1/9 × 0.00300 mmol = 0.00033 mmol [CpMo(CO)3]2. The probability that a molecule has the formula [CpW(CO)3]2 = 2/3 × 2/3 = 4/9, and 4/9 × 0.00300 mmol = 0.00133 mmol [CpW(CO)3]2. The probability that a molecule has the formula Cp(CO)3Mo–W(CO)3Cp = 2 × 1/3 × 2/3 = 4/9, and 4/9 × 0.00300 mmol = 0.00133 mmol Cp(CO)3Mo–W(CO)3Cp. 13.57 H The IR bands at 1945 and 1811 cm–1 are similar to those of [Ti(CO)4(5-C5H5)]–, suggesting similarites in structure and charge B between this Ti complex and Z. If the other IR bands are for B–H H H stretches and the 1H NMR shows two signals of relative area 3:1, it H is reasonable to suggest that the boron might be present in the BH4– Ti ion, with three of the hydrogens in one environment, the fourth in CO OC another. The peak at 2495 cm–1 is distinctly different in energy C from the peaks at 2132 and 2058 cm–1 and might correspond to a stretch O C O involving the less abundant of the protons. All is consistent with a structure having four CO ligands and one BH4– ligand (which has replaced Z the cyclopentadienyl ligand), and a negative charge. – Symmetry analysis indicates that this complex should have two C–O stretching bands in the infrared (A1 and E in local C4v symmetry) and three B–H bands (two A1 and one E in local C3v symmetry). The higher energy B–H stretch is for the terminal H; the lower energy bands are for the bridging H atoms. 13.58 Elemental analysis: calculated: 36.4% C, 6.10% H by mass. 1 H NMR: peak at 2.02 corresponds to methyl groups on Cp ring (15 protons), peak at –11.00 corresponds to hydrides (which typically have negative chemical shifts) (5 protons); protons are in desired 3:1 ratio. IR: If the environment around the osmium is assigned C4v symmetry, using the method of Chapter 4 gives the representation: C4v E 2 C4 C2 5 1 1 2 v 3 2 d 1 CH3 CH3 CH3 CH3 H H CH 3 Os H H H This reduces to 2A1 + B1 + E. The A1 and E representations are IR-active, for a total of three IRactive vibrational modes, matching the three bands in the spectrum. Copyright © 2014 Pearson Education, Inc. Chapter 13 Organometallic Chemistry 198 2p 2p 13.59 a. 2s 2p 2 p 2s B 13.60 13.61 13.62 BF F b. Because the π* orbitals of BF are concentrated on the boron to an even greater extent than π* orbitals of CO are concentrated on the carbon (because the difference in electronegativity between the atoms in BF is greater than the difference in electronegativity between the atoms in CO), BF would be expected to be a stronger πacceptor ligand. (For a computational comparison of BF, CO, and other diatomic ligands, see U. Radius, F. M. Bickelhaupt, A. W. Ehlers, N. Goldberg, and R. Hoffmann, Inorg. Chem. 1998, 37, 1080.) c. The stronger π-acceptor nature of BF should reduce the concentration of electrons on the Ru atoms in comparison with [(5-C5H5)Ru(CO)2]2. Consequently, the carbonyl ligands in the BF complex would not act as strongly as π-acceptors, and the C–O stretching vibrations should be at higher energies than 1939 and 1971 cm-1 (reported values are 1960 and 2012 cm-1). The product is hexaferrocenylbenzene (Figure 13.32)! In addition to being inherently interesting because of its structure and symmetry, it has potential as a precursor to a variety of derivatives whose properties could be tunable for electronic, catalytic, and other applications. The article provides references to these and other potential applications. OC OC O C O C Fe Fe C O C O CO CO Spectral data are interpreted in the reference. Different software and different parameter settings will generate orbitals with slightly different shapes and energies than those shown in the text, although the results should be similar. The relative energies in Part c in particular may differ from those in the text, because this calculation is sensitive to the methods used. In Part d the dz2 conical nodal surface of the orbital is close to the p orbitals of the C5H5 rings (the dz2 lobes point toward the centers of the rings), so the interaction between this orbital and the rings is weak. Copyright © 2014 Pearson Education, Inc. Chapter 13 Organometallic Chemistry 13.63 Two orbitals that show ethylene as both a donor and acceptor are shown below. H Cl Cl C C H Donor interaction (from π orbital of ethylene): 13.65 H Pt Cl 13.64 199 H Acceptor interaction (involving π* orbital of ethylene): The shapes of the orbitals should be similar to those in Figure 13.8. The eg* orbitals have different shapes because the d orbitals involved (dz2 and dx2–y2) have different shapes. a. See orbitals in Section 13.4.4, page 499. b. px , dxz py , dyz s, pz , dz 2 c. Depending on the parameters used, the upper and lower lobes of the p orbitals should merge in the lowest energy π orbital (to give clouds above and below the plane of the nuclei), and the lobes of two p orbitals in one of the upper π orbitals (derived from the p orbitals in front in the diagram at upper right, above) may also merge. d. The analysis can proceed similarly to the discussion of ferrocene in Section 13.5.2. Copyright © 2014 Pearson Education, Inc. Chapter 13 Organometallic Chemistry 200 13.66 a. 2-Node Group Orbitals z y x 1-Node Group Orbitals 0-Node Group Orbitals 2-Node Group Orbitals: b. 1-Node Group Orbitals: px 0-Node Group Orbitals: py dxz s, d z2 dyz pz The representation , shown below, reduces to the irreducible representations listed. These match the group orbitals and their matching metal orbital assignments in Parts a and b. The valence s orbital matches the A1g representation, so both the s and dz2 orbitals of nickel can interact with the zero-node group orbital on the left. c. D4h A1g B2g A2u B1u Eg Eu d. none dxy E 8 1 1 1 1 2 2 2C4 0 1 –1 1 –1 0 0 C2 0 1 1 1 1 –2 –2 2C2 0 1 –1 –1 1 0 0 2C2 0 1 1 –1 –1 0 0 i 0 1 1 –1 –1 2 –2 2S4 0 1 –1 –1 1 0 0 h 0 1 1 –1 –1 –2 2 2d 4 1 1 1 1 0 0 z2 xy z (xz, yz) (x, y) Comparisons can be made similarly to those for ferrocene in Section 13.5.2. Copyright © 2014 Pearson Education, Inc. 2v 0 1 –1 1 –1 0 0 Chapter 14 Organometallic Reactions and Catalysts 201 CHAPTER 14: ORGANOMETALLIC REACTIONS AND CATALYSIS 14.1 – a. [Mn(CO)5] + H2C=CH–CH2–Cl b. Oxidative addition – H2C=CH–CH2–Mn(CO)5 + Cl CH3 Cl PPh3 + CH3I Ir Ph3P Ir Ph3P C O C O c. Ir Ph3P PPh3 I Cl Ph3P Ir Ph2P H PPh3 PPh3 Two possibilities: methyl migration and addition, or ligand substitution O O C C + PPh3 Fe CH3 PPh3 or Fe CH3 PPh3 C O 14.2 PPh3 Cyclometallation Cl d. Cl C O + CO Fe CH3 C O e. (5-C5H5)Mn(CO)3[C(=O)CH3] (dissociation and methyl migration) f. H3C–Mn(CO)5 + SO2 Mn bonds) a. H3C–Mn(CO)5 + P(CH3)(C6H5)2 H3C(C=O)–Mn(CO)4[P(CH3)(C6H5)2] (methyl migration and phosphine addition) b. [Mn(CO)5] + (5-C5H5)Fe(CO)2Br (nucleophilic displacement) – (5-C5H5)Mn(CO)3(CH3) + CO H3C(SO2)–Mn(CO)5 (1,1 insertion, with C—S and S— (CO)5Mn–Fe(5-C5H5)(CO)2 + Br Copyright © 2014 Pearson Education, Inc. – 202 Chapter 14 Organometallic Reactions and Catalysts H c. OC Ph3P trans-Ir(CO)Cl(PPh3)2 + H2 (oxidative addition) Ir PPh3 H Cl 14.3 d. W(CO)6 + C6H5Li e. Alkyl migration of CH3 to one of the adjacent CO ligands, followed by addition of 13CO. Isomeric products: 1/3 each of the fac enantiomers, 1/3 mer isomer. There should be no 13C in the acyl group. f. Alkyl migration of CH3 to one of the adjacent CO ligands, followed by addition of 13CO. Isomeric products: two mer species, enantiomers if the 13CO is taken into account, identical otherwise. a. cis-(13CO)(CH3CO)Mn(CO)4 Mn(CH3)(CO)5, 25% with no 13CO, 25% with 13CO 13 trans to CH3, 50% with CO cis to CH3. b. – [C6H5COW(CO)5] + Li+ (nucleophilic attack on carbonyl C) C6H5–CH2Mn(CO)5 O O C C h CO + C C O O c. [V(CO)6] + NO d. Cr(CO)6 + 2 Na/NH3 e. Fe(CO)5 + NaC5H5 Na+ + [(5-C5H5)Fe(CO)2]– + 3 CO f. [Fe(CO)4]2– + CH3I [(CH3)Fe(CO)4]– + I– g. [V(CO)5(NO)] + CO CH3 Ph3P Rh PPh3 2 Na+ + [Cr(CO)5]2– + CO CH4 + Ph3P 14.4 Mn Ph2P Rh Ph3P [(C5H5)Fe(CO)3]+ + NaH A (C7H6O2Fe) B (colorless gas) + C (C7H5O2Fe) (purple-brown) A PPh3 A = (5-C5H5)Fe(CO)2H B = H2 (See Figure 13.35, p. 509 for structure of C.) C= [(5-C5H5)Fe(CO)2]2 D (C7H5O2FeI) (brown) D = (5-C5H5)Fe(CO)2I C + I2 D + TlC5H5 E (C12H10O2Fe) + TlI (See Figure 13.35, p. 509 for structure of E.) E = (C5H5)2Fe(CO)2 (one Cp 1 and one Cp 5) E F = (5-C5H5)2Fe, ferrocene F (C10H10Fe) + colorless gas (CO) Copyright © 2014 Pearson Education, Inc. Chapter 14 Organometallic Reactions and Catalysts 14.5 A (C10H12FeO2S) [(5-C5H5)Fe(CO)2]– + ClCH2CH2SCH3 A + heat 203 B A has bands at 1980 and 1940 cm–1, B at 1920 and 1630 cm–1. The sulfur reagent loses Cl– and bonds to the Fe as an alkyl ligand to form A, with two carbonyls. This is an example of nucleophilic displacement of chloride by an organometallic anion. A then rearranges, with the S becoming attached to Fe, and the alkyl migrates to a carbonyl carbon. B contains an ordinary carbonyl and an acyl C=O bond, for the two quite different C–O stretching energies. O C S Fe CH2 C O CH3 O C CH3 Fe CH2 S C O A 14.6 a. CH2 CH2 B Two term rate laws like this could be the result of two parallel associative reactions— the first by solvent, the second by the phosphate—or could result from a dissociative reaction for the first term and an associative reaction for the second. First term, dissociative, k1 [V(CO)5(NO)]: V(CO)4(NO) + CO V(CO)5(NO) PR3 V(CO)4(NO)(PR3) Second term, associative, k2[PR3][V(CO)5(NO)]: PR3 V(CO)5(NO) b. 14.7 V(CO)5(NO)(PR3) V(CO)4(NO)(PR3) + CO V(CO)5[P(OCH3)3](NO) If the NO is a bent, 1-electron donor, it can be an 18electron species ( 5 + 5×2 + 2 + 1 = 18). If the NO is a linear, 3-electron donor, the total is 20 electrons. Co2(CO)8 Co2(CO)7 + CO K1 (fast equilibrium) [Co2(CO)7] = K1[Co2(CO)8]/[CO] Co2(CO)7 + H2 Co2(CO)7H2 k2 (slow) Co2(CO)7H2 + CO 2 HCo(CO)4 (fast) Rate = k2[Co2(CO)7][H2] = k2K1 [Co2(CO)8][H2]/[CO] 14.8 This depends on the cone angle of the phosphine ligands, with the order PPh3 > PBu3 (estimated) > P(OPh)3 > P(OMe)3 from Table 14.1 (p. 543). The PPh3 should dissociate most rapidly and the P(OMe)3 should dissociate least rapidly. 14.9 K for the dissociation reaction is in the order PPh3 > PMePh2 > PEt3 > PMe3, as a result of a combination of decreasing cone angle and increasing negative charge on the phosphorus. Alkyls push more electron density onto P than phenyl rings. Copyright © 2014 Pearson Education, Inc. 204 14.10 Chapter 14 Organometallic Reactions and Catalysts a. Tolman employed Ni(0) complexes of general formula Ni(CO)3 (PX1X 2 X 3 ) where the substituents of the monodentate phosphine were varied. b. The parameter was defined as follows to determine the energy of the Ni(CO)3 (PX1X 2 X 3 ) A1 carbonyl stretching mode, with each phosphine substituent contributing a value. The 2056.1 cm -1 relatively low energy CO absorption is the A1 mode of Ni(CO)3(P t Bu 3 ) , containing the strong -donor P t Bu 3 . 3 CO ( A1 ) 2056.1 cm -1 c. i1 -1 These values range from 0 to 19.6 cm , with the highly electronegative Cl (14.8), F (18.2) and CF3 (19.6) substituents contributing the most towards shifting the A1 carbonyl stretching mode to higher energy. 14.11 d. The debate over the classification of phosphine ligands as -donors and -acceptors is discussed in this reference. Tolman states that while highly electronegative substituents have higher values (and cause higher energy CO absorptions) and more electrondonating substituents have lower values, the relative importance of donation and acceptance for these phosphine-nickel bonds is difficult to assess. a. The conversion of the Ni(II) complex to a Ni(0) complex is shown here: PPh2 PPh2 N Ni H CO CO NH Ni CO PPh2 PPh2 Upfield-shifted metal-hydride 1H resonances are common. In this reaction, the Ni—H resonance (-18 ppm) is replaced by one at 8.62 ppm, consistent with formation of the secondary amine. Two carbonyl absorptions are consistent with the cis carbonyl product shown. The authors speculate that the first step in this reaction is CO coordination at Ni(II) to afford a five coordinate intermediate that presumably brings the amido nitrogen and the hydride into cis positions. Reductive elimination results in N—H bond formation. Initial CO binding renders the Ni(II) center more electrophilic (via -backbonding) and a better oxidizing agent. CO binding also generates a five-coordinate complex with the amido nitrogen and the hydride in closer proximity to facilitate the orbital overlap necessary for reductive elimination. Copyright © 2014 Pearson Education, Inc. Chapter 14 Organometallic Reactions and Catalysts b. 205 The alternate reaction pathway exhibited by the Ni—CH3 complex is: PPh2 N Ni H 3C PPh2 CH3 CO C Ni N PPh2 O PPh2 CO N CH3 PPh2 O CO Ni PPh2 CO The first reaction is proposed as a migratory insertion of CO into the Ni—CH3 bond, affording an acyl complex with a diagnostic 1621 cm -1 carbonyl absorption. Further reaction with CO results in reductive elimination of a N—C 2 bond and formation of a sp neutral amide-functionalized bidentate phosphine ligand. As in Part a, the first step in the reductive elimination reaction is hypothesized as CO coordination at Ni(II) of the intermediate, resulting in a five-coordinate species with the amido nitrogen atom and the acyl ligand carbon atom in close proximity. c. The interesting feature of the Part b scheme is that no N—C sp3 bond formation occurs when CO is added to the Ni—CH3 complex. The authors of the reference primarily attribute this to an electronic effect. The more electron-rich Ni(II) center in the Ni—CH3 complex is less reactive towards reductive elimination compared to the less electron-rich Ni(II) center in the Ni—H complex. Note that these reductive eliminations are proposed to begin via initial coordination of CO, rendering the metal center more electrophilic for subsequent reductive elimination by virtue of -backbonding. In these reductive elimination reactions, Ni(II) formally serves as the oxidizing agent (it goes from Ni(II) to Ni(0)); a less electron-rich Ni(II) would be expected to be a better oxidizing agent. An idea not mentioned in the reference is that activation barriers for reductive eliminations involving H atoms are generally lower than those involving CH 3 groups (leading to faster reductive elimination rates with H atoms). The spherical symmetry of the 1s H orbital better facilitates effective overlap throughout reductive elimination pathways when compared to the sp3 hybrid orbital employed by a methyl group. 14.12 a. This double-cyclometallation very likely occurs via -bond metathesis. A sketch of a possible four-centered transition state (compare to the general -bond metathesis transition state in Figure 14.8) is shown below. This structure presumably leads to toluene expulsion with concomitant orthometallation of the aromatic ring. Ph Ph H C 2 Ph H 2C H2C N b. H Zr N CH2 N N Since Zr(IV) is in its highest oxidation state, and is not electronically predisposed to participate in sequential oxidative addition/reductive elimination steps to affect orthometallation via metal hydride intermediates, a -bond metathesis pathway seems like the best choice. The rate of this intramolecular process would be expected Copyright © 2014 Pearson Education, Inc. 206 Chapter 14 Organometallic Reactions and Catalysts independent of H2 pressure. This pathway includes neither the release of hydrogen gas nor hydride intermediates that could arise from hydrogen gas. The reference does not speculate as to why the rate with R = CH 2 Ph is the slowest. It is possible that the steric bulk of the four aromatic rings in the proposed transition state structure may raise the activation barrier for formation of the four-centered transition state, decreasing the reaction rate. 14.13 If it loses CO followed by migration of CH3 to an adjacent position, all the CO lost should be 12CO, because the CO ligands cis to the CH3C=O will be the ones lost. 14.14 a. CH3 There are two possible products, O both a result of methyl migration C 1 13 O followed by carbonyl addition. The C methyl group can move to either Mn PMe3 CO the 1 position or the 2 position. C The resulting products are different O PMe3 only in the location of the 13CO. b. 14.15 [(C5H5)2Fe2(CO)4] + Na/Hg 2 C O H3C O C A Mn PMe3 PMe3 C C + PhNa A + D A: B: C: D: H Br Fe B B + LiAlH4 a. C The new IR band is from the acyl carbonyl, and should be near 1630 cm–1 (see Problem 14.5). – A + Br2 14.16 O O C 13 O C Fe Fe CO C O CO C O C O A B C Sodium acts as reducing agent. Anion A has lower energy C–O stretches, as expected (see p. 488). Br2 acts as oxidizing agent and is the source of the bromo ligand. The NMR peak at –12 ppm is caused by the hydride ligand. benzene, C6H6 The bromoethoxide anion adds to a carbonyl; the hard oxygen of the anion adds to the hard carbon. Then Br– is lost, leaving a positive carbon. The alkyl tail can then bend around and react with the carbonyl oxygen, giving the compound shown here: O C Br Re(CO)5Br + BrCH2CH2O– OC Re – C C O O O C Br O OC C OCH2CH2Br Re O C OCH2CH2Br C C O O O C Br OC Re O + Br– C C C O O Y Copyright © 2014 Pearson Education, Inc. – O Chapter 14 Organometallic Reactions and Catalysts b. 14.17 207 There are 2 electrons donated from each of the six ligand positions, and Re+ has 6 electrons for a total of 18. In the isomer shown, there are three different carbonyls, and the carbene ligand has two identical carbons, so there are five different magnetic environments for carbon. In the other possible isomer, with Br trans to the alkoxide ligand, there are not five different carbon environments. Finally, Ag+ can remove Br– from the complex. Nitrenium ligands are similar to N-heterocyclic carbene (NHC) ligands (see Figure 14.33(a)) except that they have a nitrogen atom in place of the coordinating carbon; three nitrogens are in a row. R groups that have been studied include those with phosphine arms that can also coordinate to metals. R2 R2 M P M P R N N N + R N N + N Nitrenium ligands have been demonstrated to be weaker sigma donors than NHC ligands on the basis of infrared and bond length evidence, supported by density functional theory calculations; see the reference for details. The phosphine arms that also coordinate to metals give nitrenium ligands pincer characteristics comparable to the ligands discussed in Section 14.1.5 14.18 a. This is the example described in Example 13.4, p. 531. I + PPh3 II II: IR: 2038, 1958, 1906 cm–1 NMR: 7.62, 7.41 muliplets (15), 4.19 multiplet (4) II + PPh3 III III: IR: 1944, 1860 cm–1 NMR: 7.70, 7.32 multiplets (15), 3.39 singlet (2) II has 3 CO ligands in a fac geometry. The NMR shows one PPh3 (15) and the ethylene hydrogens (4). III has only 2 CO ligands in a cis configuration. The ratio of NMR integrated peaks is now 15:2 because there are two PPh3 groups. b. I + Ph2PCH2PPh2 IV IV 2036, 1959, 1914 cm–1, 35.8% C, 2.73% H Replacing two CO’s with the phosphine gives a compound with 25.7% C, 3.1% H. Replacing one CO and the Br with the phosphine gives a compound with 28.9% C, 3.3% H. The only way to get the analysis to work out is to have a single diphos ligand bridging two Re atoms after loss of CO from each: [(Re(CO)3Br(C(O2C2H4))]2{(PPh2)2CH2}, IV, has 36.14% C, 2.46% H. Copyright © 2014 Pearson Education, Inc. 208 Chapter 14 Organometallic Reactions and Catalysts c. Re(CO)5Br + V I + S2CN(CH3)2– V has no metal, has no IR bands between 1700 and 2300 cm–1, has NMR bands at 3.91 (triplet), 3.60 (triplet), 3.57 (singlet), and 3.41 (singlet). Acting as a nucleophile, the dithiocarbamate ion attacks the carbene to form (CH3)2NC(=S)SC2H4O–, which can pick up a proton from trace amounts of water to make the hydroxy compound (CH3)2NC(=S)SC2H4OH, V. The 1500 and 977 cm–1 bands are the N–C and C–S bands, and the NMR shows triplets for the two CH2 units, singlets for the CH3’s. 14.19 The mechanism is the one described in problem 16, with formation of the alkoxide ion by the reaction: O + Br– 14.20 BrCH2CH2O– The reaction proceeds by substitution of Mn(CO)5– for Br–, followed by alkyl migration, addition of Mn(CO)5–, and finally cyclization: Mn(CO)5– + BrCH2CH2CH2Br (CO)5MnCH2CH2CH2Br + Br– Mn(CO)5– O O C C – Mn (CO)5Mn C C O O O O C C O C (CO)5Mn CH2CH2CH2Br Mn O C CH2CH2CH2Br C C O O O O C C (CO)5Mn Mn – O C + Br– C C O O 14.21 The acyl metal carbonyl has a resonance structure with a negative charge on oxygen, shown below, where the proton can add. Protonation is then similar to the alkylation reaction described in Section 13.6.2. – O O C R M(CO)x C + M(CO)x R Copyright © 2014 Pearson Education, Inc. Chapter 14 Organometallic Reactions and Catalysts 14.22 a. Acetaldehyde from ethylene: C2H4 + CO + H2 [PdCl4]2– 209 O H3C C + H2O H The Wacker process (Figure 14.21) with ethylene as the starting alkene will result in acetaldehyde. b. Ethyl propionate from chloroethane: [Co(CO)4]– 14.23 CH2CH2Cl CH3CH2Co(CO)4 CO (CH3CH2CO)Co(CO)4 C2H5OH CH3CH2COOC2H5 + HCo(CO)4 c. Pentanal from 1-butene: The hydroformylation process (Figure 14.18) does this. d. 4-phenylbutanal from an alkene: Again, the hydroformylation process should do this, starting with 4-phenyl-1-propene. e. Wilkinson’s catalyst [ClRh(PPh3)3] (Figure 14.22) should do this. D2 would be added across the least hindered double bond. f. Catalytic deuteration with H3TaCp2 (Figure 14.17) should deuterate the phenyl ring without affecting the methyl hydrogens. Figure 14.19 shows this process of hydroformylation. n-pentanal results if R = C2 H5. Identification of the steps: 1. Dissociation of one CO 2. Addition of the alkene 3. 1,2 insertion 4. Addition of CO 5. Alkyl migration 6. Oxidative addition 7. Reductive elimination of the aldehyde 14.24 Hydroformylation (Figure 14.19) with Rh(CO)2(PPh3)2 as the catalytic species will work, starting with 2-methyl-1-butene: H3C–CH2–C(CH3)=CH2. Copyright © 2014 Pearson Education, Inc. 210 14.25 Chapter 14 Organometallic Reactions and Catalysts a. Direct metathesis would occur as follows: In addition, reactants could undergo self-metathesis, for example + and the products of the direct metathesis could undergo further metathesis to give a variety of products. One of these could be formed as shown here: + b. c. d. This is Hérisson and Chauvin’s classic experiment, with products shown in Figure 14.27. As in Part a, self metathesis can also occur, and the products of direct metathesis can also undergo further metathesis to give a variety of products. + + Copyright © 2014 Pearson Education, Inc. Chapter 14 Organometallic Reactions and Catalysts 14.26 211 The catalyst is an 18-electron species with M = W. The first steps in the catalytic cycle are diagrammed here: W W = (CO)5W R = C6H5 CR2 W W CR2 CR2 + = CR2 W + CR2 = W CR2 W CR2 W ... Polymer 14.27 a. Me Pr + + Me Pr + Me Me Me Pr Me + Pr Me = methyl Pr = n-propyl b. + Pr “Double cross” product + Later metathesis products Pr The pairwise mechanism would call for the dimethyl and dipropyl products to be formed first, followed by the “double cross” product. In the non-pairwise mechanism, the double cross product should form simultaneously with the dimethyl and the dipropyl products. Experimental results showed that the double cross product formed (along with the dimethyl and dipropyl products) at the beginning of the reaction, consistent with the non-pairwise mechanism. Copyright © 2014 Pearson Education, Inc. 212 Chapter 14 Organometallic Reactions and Catalysts H H 14.28 a. C (C6F5)2B B(C6F5)2 H b. Zr (Cp)2 H One possible route: H Cp2Zr + H2 C CH2 Cp2Zr H2C empty site 1, 2-insertion H Cp2Zr CH2CH3 CH2 C2H4 etc. Cp2Zr 1, 2-insertion CH2CH2CH2CH3 Cp2Zr H2C 14.29 a. CH2 Details of the scheme, together with schematic diagrams of the proposed catalytic cycle and related cycles, are presented in the reference. The catalyst precursor is the 1,2dimethoxyethane (DME) adduct of N W(OC(CF3 )2 Me)3 , dubbed 1-DME (the crystal structure is provided in the reference). Proposed steps include: Reaction of 1-DME with 3-hexyne to form metallacycle Dissociation of nitrile, leaving tungsten carbyne complex Reaction with p-methoxybenzonitrile to form second metallacycle Dissociation of product to regenerate 1-DME 1. 2. 3. 4. b. CH2CH3 The first product was considered to be 1-(4-methoxyphenyl)-1-butyne, formed as shown below. However, this compound accounted for only 15 mole percent of the reaction products. The major product, bis(4-methoxyphenyl)acetylene, accounting for 77 mole percent of the products, was attributed to secondary metathesis, also involving metallacycle intermediates. The article discusses in detail alternative cycles for forming bis(4-methoxyphenyl)acetylene. Et N W X3 1-DME + Et C N C C W X3 C Et N C + X3W C Et Et Et + N C OCH3 = N W X3 + Et C C OCH3 1-(4-methoxyphenyl)-1-butyne H3CO C N C X3W C C bis-(4-methoxyphenyl)acetylene Copyright © 2014 Pearson Education, Inc. OCH3 Et OCH3 Chapter 14 Organometallic Reactions and Catalysts 213 14.30 CH3 Fe CH3 CH3 C 2 H4 H2C Fe CH2 H3 C CH2 P(OCH3)3 Fe (H3CO)3P P(OCH3)3 H2C X 14.31 RhCl3 3 H2O + P(o-MePh)3 Rh–Cl = 351 cm–1 I (blue-green) (C42H42P2Cl2Rh) eff = 2.3 B.M., I is trans-RhCl2(PR3)2 The IR is the Rh–Cl asymmetric stretch. The cis isomer would give two IR bands. One unpaired electron (15 electron species, square planar). I + heat II (yellow, diamagnetic) Rh:Cl = 1, = 920 cm–1 II Loss of one Cl and combination of two CH3’s to form a single tridentate phosphine ligand with a π bond. Sixteen electrons around Rh. The double bond in the ligand is perpendicular to the Rh–Cl–P plane because of the size and geometry of the benzene rings. II + SCN– NMR: 6.9–7.5 3.50 2.84 2.40 III Rh(SCN) Area 12 1 3 3 II: Type aromatic doublet of 1:2:1 triplets singlet singlet R2P Rh PR2 Cl III: Same overall structure as II, with SCN substituted for Cl. The singlets are the methyl protons (two separate environments—not obvious from the drawing), the doublet is from the vinylic protons, and the aromatic multiplets are the sum of all the phenyl protons. II + NaCN NMR: 7.64 6.9 - 7.5 2.37 IV (C21H19P, mw = 604) = 965 cm–1 Area 1 12 6 Type singlet aromatic singlet IV is the diphosphine ligand, C42H38P2, shown above in compound II. (CN– substitutes for all the ligands in this reaction.) The IR band is characteristic of trans vinylic hydrogens. All methyl protons are equivalent in the free ligand (singlet), the vinylic protons are a singlet at 7.64, and the phenyl protons are a multiplet at 6.9–7.5. The change in the vinyl hydrogen IR band shows that the ethylene does coordinate to Rh. Coordination reduces the C=C bonding, which also reduces the C–H bending energy (electrons are drawn away from C–H, toward C–Rh). Copyright © 2014 Pearson Education, Inc. 214 Chapter 14 Organometallic Reactions and Catalysts 14.32 a. [MeC C(CH2)2OOC(CH2)]2 = H3CC CCH2CH2OOCCH2CH2COOCH2CH2 C O H 3C C C C C O C O C O C C MeC C(CH2)8COO(CH2)9 C O C C C C CH3 C C CH3 O O b. CCH3 O CMe O H 3C C O C O C O 14.33 Evidence in support of the intermediate shown at the right includes: (1) (2) An infrared band at 2104 cm–1 has been observed. This I band is similar to the higher energy band in the Ir I analogue, [CH3Ir(CO)2I3]–, which has carbonyl bands at 2102 and 2049 cm–1. A second band, expected for [CH3Ir(CO)2I3]–, would be hidden under strong bands of the reactants and products of steps 2 and 3 of the mechanism. – CH3 Rh CO CO I The ratio of the absorbance of the 2104 cm–1 band to the absorbance of a band at 1985 cm–1 of [Rh(CO)2I2]–, the reactant in step 2, is proportional to the concentration of CH3I. This is consistent with what would be expected Copyright © 2014 Pearson Education, Inc. Chapter 14 Organometallic Reactions and Catalysts 215 from the equilibrium constant expression for formation of the intermediate, K [Rh(CO) 2 I2 ] – [CH 3 I] [CH 3 Rh(CO) 2 I3 ] – and is consistent with the steady state approximation for the mechanism. (3) (4) 14.34 The maximum intensity of the 2104 cm–1 band occurs when the product of step 3 is formed most rapidly. When 13CH3I is used, a doublet is observed in the NMR consistent with C–103Rh coupling. Other NMR data also support the proposed structure. 13 – A: [Fe(CO)3(CN)3] B: cis-[Fe(CO)2(CN)4]2 – – The cyanide ion has the capacity to replace ligands such as I and CO. The C–N stretch involves a large change in dipole moment, so this vibration, like that of CO, can be useful in characterizing cyano complexes. Because CN has a slightly smaller reduced mass than CO, C–N stretches typically occur at slightly higher energies than C–O stretches (see Section 13.4.1). In this situation, both A and B have two sets of C–N and C–O stretches, indicating that both complexes have at least two of both ligands. As more cyano ligands are added, the concentration of electrons on Fe increases, and both types of ligands become stronger π acceptors, reducing the energy of the stretching vibrations. The formula of B suggests the possibility of both cis and trans isomers. The observation of two bands in the carbonyl region is consistent with the cis isomer. (See also Problem 13.25 and its reference.) 14.35 2: Mo P Mo C O H CO H The two IR bands are as expected for this dicarbonyl complex. NMR peaks can be assigned as follows: chemical shift 5.28 (relative area 5): Cp; 1.31 (27): nine methyl groups on t-butyl groups on benzene ring; 5.15 (3): protons on benzene ring; 5.46 (2), 4.22 (2) and hidden small peak: 4-C5H6. In the 13C NMR, the resonance at 236.9 ppm can be assigned to the carbonyl carbons. This product is formed via an unusual attack of a hydride on a cyclopentadienyl carbon; for an additional example of this type of attack, see footnote 17 in the reference. Copyright © 2014 Pearson Education, Inc. 216 14.36 Chapter 14 Organometallic Reactions and Catalysts X: B N H H Y: B N H Z: B N TBS Cr OC CO C O H H In heterocycle Z coupling of the N–H proton with the 14N nucleus (which has S = 1) results in a triplet, and coupling for the B–H proton with the 11B nucleus (which has S = 3/2) results in a quartet. (The original spectrum is shown in the reference.) 14.37 See Figure 13.32 for a diagram of this molecular “ferrous wheel”! Copyright © 2014 Pearson Education, Inc. Chapter 15 Parallels Between Main Group and Organometallic Chemistry CHAPTER 15: PARALLELS BETWEEN MAIN GROUP AND ORGANOMETALLIC CHEMISTRY 15.1 a. Mn2(CO)10 + Br2 2 Mn(CO)5Br H C b. – HCCl3 + excess[Co(CO)4] Co(CO)3 (CO)3Co Co (CO)3 c. Co2(CO)8 + (SCN)2 2 NCSCo(CO)4 R C O C d. Co2(CO)8 + C6H5–CC–C6H5 Co O 15.2 15.3 e. Mn2(CO)10 + [(5-C5H5)Fe(CO)2]2 a. Tc(CO)5 b. [Re(CO)4]– CH2 c. [Co(CN)5]3– CH3 d. [CpFe(C6H5)]+ e. [Mn(CO)5]+ CH3+ f. Os2(CO)8 H2C=CH2 C C C O CH4 Many examples are possible. Two for each: CH3 CH3 (6-C6H6)Mn(PPh3)2 [Ni(CS)(PMe3)2Cl2]+ b. CH CH (4-C4H4)Co(CS) (6-C6H6)Rh O C Co C O C O 2 (CO)5Mn–Fe(5-C5H5)(CO)2 CH3 a. R Copyright © 2014 Pearson Education, Inc. 217 Chapter 15 Parallels Between Main Group and Organometallic Chemistry 218 15.4 15.5 c. CH3+ CH3+ d. CH3 – CH3 e. (5-C5H5)Fe(CO)2 (5-C5H5)Fe(CO)2 f. Sn(CH3)2 Sn(CH3)2 Mo(borazine)(PMe3)(CS) – [V(CO)3(en)] – (8-C8H8)Ru(PEt3) [Co(N2)(CO)2(bipy)]+ Co(CO)3Cl2 – [(6-C6H6)Cr(CO)2] (5-C5H5)Ir(CO) – [Tc(CO)4] Many possibilities exist; some of those given here may not be synthetically accessible. a. C2H4 b. P4 c. cyclo-C4H8 d. S8 a. All three complexes are 18 electron species, with the benzene ring of [(C5Me5)Fe(C6H6)]+ replaced by three carbonyl groups or two carbonyls and a phosphine. All have a formal coordination number of six. b. The experimental results are more complex than might be expected. (CO)4Fe=Fe(CO)4 [Ir(CO)3]4 (CO)4Fe Ru(CO)4 (CO)4Ru Fe(CO)4 cyclo-[Fe(CO)4]8 [(C5Me5)Fe(C6H5)]++ H– Fe H H [(5-C5H5)Fe(CO)3]+ + H– 5 + (5-C5H5)Fe(CO)2H – [( -C5H5)Fe(CO)2PPh3] + H OC [(5-C5H5)Fe(CO)2]2 + H2 PPh3 Fe CO (See A. Davison, M. L. H. Green, and G. Wilkinson, J. Chem. Soc., 1961, 3172.) H H Copyright © 2014 Pearson Education, Inc. Chapter 15 Parallels Between Main Group and Organometallic Chemistry 15.6 a. CH2, Fe(CO)4, [Mn(CO)4]–, and PR2 each has two frontier orbitals, each with one electron. Each fragment is two ligands short of the parent polyhedron (octahedron or tetrahedron). b. Fe(CO)4 and CpRh(CO) each has two frontier orbitals, each with one electron. c. [Re(CO)4]– has two frontier orbitals, with one electron in each. R2C is isolobal with two frontier orbitals, each with one electron. 219 15.7 The W(Cp)(CO)2 fragments are isolobal with CR and CPh, each with three orbitals containing one electron each; PtR2 and Cu(C5Me5) are isolobal, each with two orbitals containing one electron each. 15.8 a. If Mn has the lower energy orbital, then the bonding molecular orbital has a greater contribution by the Mn orbital, and the electrons in the orbital are polarized toward Mn. b. The gold orbitals are higher in energy. Rather than matching energies with the lowest of the π orbitals of the Cp ring, the Au orbitals will match better with the higher π orbitals, which have a nodal plane cutting across the ring. a. CH2 and Fe(CO)4 are isolobal, each with two orbitals containing one electron apiece. b. Mn(CO)2(C5Me5) is isolobal with [Mn(CO)5]+, a 16 electron species, which is in turn isolobal with CH2. The Mn–Sn–Mo fragment is similar to allene, C=C=C, in which the double bonds force a linear geometry. 15.9 15.10 [C(AuPPh3)5]+ Structure: slightly distorted trigonal bipyramid, as expected from VSEPR. [C(AuPPh3)6]2+ Structure: slightly distorted octahedron, as expected from VSEPR. The structure of the 7-coordinate [C(AuPPh3)7]3+ has not been reported, but interest in the realm of highly coordinated carbon compounds continues; for a recent update, see H. G. Raubenheimer and H. Schmidbaur, Organometallics, 2012, 31, 2507. In these complexes, each of the AuPPh3 groups can be viewed as having an sp hybrid orbital pointing in toward the carbon. Interactions between these hybrids with the s and p orbitals of carbon give rise to four bonding orbitals in each case. The eight valence electrons available fill these orbitals. The result is the equivalent of four bonds spread over the complex (bond order of 4/5 in [C(AuPPh3)5]+ and 4/6 in [C(AuPPh3)6]2+. Metal-metal bonding is also likely to contribute to the stability of these structures. For more details, see H. Schmidbaur, et. al., Angew. Chem. Int. Ed. Engl., 1989, 28, 463 and Angew. Chem. Int. Ed. Engl., 1988, 27, 1544. 15.11 Carbonyl ligands are significantly stronger π acceptors than cyclopentadienyl. Replacing three CO ligands with Cp on each metal atom, therefore, leaves the metals with greater electron density for involvement in metal–metal bonding, leading to shorter metal–metal bonds in the cobalt complexes. In addition, there is a general decrease in atomic radii from left to right in this row of transition metals, so Co atoms are slightly smaller than Fe atoms (see Table 2.8). Copyright © 2014 Pearson Education, Inc. Chapter 15 Parallels Between Main Group and Organometallic Chemistry 220 15.12 15.13 a. The orbital interactions are similar in the two cases, except that the π orbitals of the P5 ring are lower in energy than those of the C5H5 ring, resulting in a stronger ability of the P5 ring to accept electrons from the metal (see also reference 11 in this chapter). b. Because the P5 ring acts as a strong acceptor, electrons are accepted into antibonding – orbitals in the titanium complexes, leading to longer P–P distances than in P5 (calculated at 2.12 Å). The Ti–P distances are calculated to be shorter in the theoretical structure – 2– [Ti(P5)] than in [Ti(P5)2] , in which ligands compete for electrons in the same d orbitals on the metal. The consequence is that the predicted P–P distance is longer (2.24 Å) in – [Ti(P5)] , where the ligand can act as a stronger acceptor because it is closer to the metal, 2– than in [Ti(P5)2] (calculated: 2.175 Å; experimental in reference 11: 2.154 Å). a. The Mn(CO)5 fragments have a single electron in their HOMO, largely derived from the dz2 (hybridized with the pz orbital) of Mn. Lobes of the HOMOs of two Mn(CO)5 fragments interact in a sigma fashion with the g orbital of C2 (see Figure 5.5 for the approximate shape) derived primarily from the pz orbitals of the C atoms: O C O C OC Mn O C O C C C C O C O b. The empty π* orbitals of C2 can interact with occupied d orbitals of Mn(CO)5 fragments: OC Mn C O C O 15.14 CO C O C O O C O C c. Mn O C O C C C Mn CO C O C O The reference points out other interactions, such as between π (bonding) orbitals of C2 and Mn(CO)5, and discusses the relative energies of the molecular orbitals of this molecule and other molecules having bridging C2 ligands. A staggered configuration is more likely, as predicted by VSEPR. The bonding is similar to that in the triply bonded [Os2Cl8]2–. Copyright © 2014 Pearson Education, Inc. Chapter 15 Parallels Between Main Group and Organometallic Chemistry 15.15 221 End view: matches dx2–y2 15.16 Cotton’s explanation is that removal of an electron effectively changes the oxidation number of Tc, causing the d orbitals in [TcCl8]2– to be smaller than those of [TcCl8]3–. This reduces the orbital overlap, weakening the bonding in spite of the higher formal bond order. The change is small—about 3 pm. (See F. A. Cotton and G. Wilkinson, Advanced Inorganic Chemistry, 5th Ed., Wiley, 1988. p. 1090.) 15.17 As the reference describes, these ions are isostructural and are based on octahedra fused at one mutual face (occupied by three bridging chlorine atoms). The shortening of the Re–Re distance upon reduction is attributed chiefly to two factors. As the complex is reduced, the net positive charge on each metal is reduced (the oxidation state changes from 4 to 3.5), which reduces the metal-metal repulsion. In addition, with the reduction of the metals, the metal d orbitals expand, enabling more effective overlap and stronger bonding. 15.18 This example is similar to the one presented in Figure 15.10. In compound 1 there are six DTolF ligands, each with a charge of 1–. Adding the charges of the bridging hydroxides, the total charge of the ligands in this complex is 8–. Consequently, the total charge on the Mo atoms must be 8+, or a charge of 4+ per Mo2 unit. Each Mo2+ has 4 d electrons, giving 8 d electrons per Mo2. Mo24+, with its 8 d electrons, has an Mo–Mo bond order of 4 (Figure 15.10). In compound 2 there are also six DTolF ligands. The bridging O2– ligands result in a total ligand charge of 10–; consequently, the four molybdenums must carry an average charge of 2.5+, or 5+ per Mo2 unit. There are now just 7 d electrons per Mo2. From Figure 15.10 we can see that Mo25+, with 7 d electrons, corresponds to a bond order of 3.5. The lower Mo–Mo bond order in 2 results in a longer bond. Copyright © 2014 Pearson Education, Inc. Chapter 15 Parallels Between Main Group and Organometallic Chemistry 222 15.19 a. Because the diphenylforamidinate ion has a charge of 1–, the iron atoms have an average oxidation state of 1.5, alternatively viewed as mixed-valent Fe(I) and Fe(II). b. The iron atoms have a total of 7 [Fe(I)] + 6 [Fe(II)] = 13 d electrons. The 3d orbitals interact in sigma (dz2), pi (dxz and dyz), and delta (dxy and dx2–y2) fashions to generate bonding and antibonding orbitals of these types. The calculated order of energy of these orbitals is shown below. Experimental data and supporting calculations suggest that the singly occupied π*, *, , and * orbitals are very close in energy, favoring a high-spin arrangement that, in accordance with Hund’s rule, results in seven electrons with parallel spin. 15.20 The clamping of the nitrogen atoms that bridge the chromium atoms is postulated to enforce short chromium—chromium bonds. The ligand, in effect, assists in pushing the chromium atoms together. Ligand design is critical to minimize inter-ligand substituent repulsion to maximize ligand chelation ability. While Figure 1 of the reference emphasizes steric effects, it is stated that repulsion between filled nitrogen and chromium nonbonding orbitals is maximized (resulting in further metal—metal bond contraction) when chelation is more robust. The aminopyridinates are shown in the same figure to chelate less effectively to the chromium atoms; the nitrogen substituents of the two ligands are directed slightly towards each other, reducing the efficacy of the chelate, and leading to longer chromium—chromium distances. The amidate substituents are directed away from the chromium—chromium bond, resulting in less inter-ligand repulsion, and slightly shorter chromium—chromium bonds. The guanidinates tune the clamping effect further by introducing steric pressure provided by the substituents emerging from the central carbon atom of the chelating ligand. This steric hindrance pushes the substituents of the chelating nitrogen atoms slightly towards the chromium atoms, resulting in the chromium atoms being pushed closer together. The magnetic susceptibility results indicate that the dichromium complex 3 (see reference) features a singlet ground state, with no unpaired electrons. If the chromium atoms are both d5 centers, then these 10 metal valence electrons will fill all five bonding molecular orbitals of the energy level diagram in Figure 15.9 to afford a quintuple bond. 15.21 It is proposed that CH 3C CCH 3 reacts with the quintuply bonded dichromium complex in a manner consistent with 2+2 cycloaddition resulting in a dichromium metallacycle with a carbon—carbon double bond and a chromium—chromium quadruple bond. The crystallographic data support this hypothesis on the basis of a carbon—carbon distance (132.6 pm) similar to that of ethylene (133.9 pm), and a chromium—chromium distance (192.5 pm) longer than that in the starting quintuply bonded complex (180.3 pm), consistent with a fourfold chromium—chromium bond order. Calculated bond orders (via DFT calculations) for the dichromium metallacyle product are 3.43 and 1.98 for the chromium—chromium and carbon—carbon bonds, respectively. Copyright © 2014 Pearson Education, Inc. Chapter 15 Parallels Between Main Group and Organometallic Chemistry 15.22 a. Oh E 6 1 3 2 (s, pz) A1g T1u Eg 8C3 0 1 0 –1 6C2 0 1 –1 0 6C4 2 1 1 0 3C2 2 1 –1 2 i 0 1 –3 2 6S4 0 1 –1 0 6S6 0 1 0 –1 3h 4 1 1 2 223 6d 2 1 1 0 fits either the s or the pz orbitals. b. It can be seen from the table above that = A1g + T1u + Eg. This can also be worked out by the more elaborate methods used in Chapter 4. c. The reducible representation for the px and py orbitals and its components are shown in the following table. Figure 15.13 shows the three T1u and the three T2g group orbitals required (which also include some pz contribution from two of the boron atoms). The T1g and T2u representations are for antibonding orbitals. Oh x,y T1g T1u T2g T2u 15.23 E 12 3 3 3 3 8C3 0 0 0 0 0 6C2 0 –1 –1 1 1 6C4 0 1 1 –1 –1 3C2 –4 –1 –1 –1 –1 i 0 3 –3 3 –3 6S4 0 1 –1 –1 1 6S6 0 0 0 0 0 3h 0 –1 1 –1 1 6d 0 –1 1 1 –1 (Rx, Ry, Rz) (x, y, z) The number of orbitals of each type can be obtained by analogy with the results for B6H62– (Section 15.4.1): 2 valence atomic orbitals of B combine to form: 15 bonding orbitals (2n + 1) consisting of: 8 framework MOs (n + 1) 1 bonding orbital from overlap of sp orbitals 7 bonding orbitals from overlap of p orbitals of B with sp hybrid orbitals or p orbitals of other B atoms 7 B–H bonding orbitals (n) 13 nonbonding or antibonding orbitals 15.24 a. C2B3H7 B5H9 B5H54– nido b. B6H12 B6H66– arachno c. B11H112– closo d. C3B5H7 B8H10 B8H82– closo e. CB10H13 B11H14 B11H114– nido f. B10H142– B10H106– arachno – – Copyright © 2014 Pearson Education, Inc. Chapter 15 Parallels Between Main Group and Organometallic Chemistry 224 15.25 15.26 15.27 15.28 a. SB10H102– B11H132– B11H114– nido b. NCB10H11 B12H14 B12H122– closo c. SiC2B4H10 B7H13 B7H76– arachno d. As2C2B7H9 B11H15 B11H114– nido e. PCB9H11 B11H14 B11H114– nido a. B3H8(Mn(CO)3) B4H8 B4H44– nido b. B4H6(CoCp)2 B6H8 B6H62– closo c. C2B7H11CoCp B9H13CoCp B10H14 B10H104– d. B5H10FeCp B6H10 B6H64– e. C2B9H11Ru(CO)3 B11H13Ru(CO)3 B12H14 B12H122– a. Ni(CO)4: CH4 Ni(CO)3: CH3+ b. [Bi3Ni4(CO)6]x The cluster needs 4n + 2 = 4(7) + 2 = 30 cluster valence electrons for a closo classification (there are 7 Bi and Ni atoms in the core of the cluster). Each Bi atom contributes its 5 valence electrons, and each CO ligand contributes 2, for a total of 3(5) + 6(2) =27. Three additional electrons are needed for a total of 30; consequently, the charge must be x = 3–; the formula is [Bi3Ni4(CO)6]3–. (See reference and citations therein for additional details on electron counting.) c. The cluster valence count starts with 6 CO ligands (12 electrons) [BixNi4(CO)6]2– and a charge of 2– , a total of 14 electrons. If x = 0, the value of 4n + 2 would be 18. Each Bi atom counts an additional 5 electrons, and each additional core atom increases 4n + 2 by 4. At x = 4, the total cluster valence electron count (34) equals 4n + 2 (4 Bi atoms and 4 Ni atoms in core, so n = 8). Consequently, the formula is [Bi4Ni4(CO)6]2– . – – nido nido closo Ni(CO)2: CH22+ Ni(CO): CH3+ B2 H 6 (Cp*RuCO)2 : Cp*RuCO is a 15 electron fragment, 3 electrons short of 18. This fragment is isolobal to the 5-electron fragment BH 2 , 3 short of 8 electrons. The metallaborane B2 H 6 (Cp*RuCO)2 is therefore isolobal with B4 H10 . B4 H10 6H + B4 H 46- (arachno) (Continued on next page) Copyright © 2014 Pearson Education, Inc. Chapter 15 Parallels Between Main Group and Organometallic Chemistry 225 B3H 7 (Cp*RuH)2 : Cp*RuH is a 14 electron fragment, 4 electrons short of 18. This fragment is isolobal to the 4-electron fragment BH, 4 short of 8 electrons. The metallaborane B3H 7 (Cp*RuH)2 is therefore isolobal with B5H 9 . B5H 9 4H + B5H 54- (nido) The reactions of B2 H 6 (Cp*RuCO)2 with Mn 2 (CO)10 and Re 2 (CO)10 , respectively, afford products with dramatically different structures. The Mn 2 (CO)10 reaction yields arachno [(Cp*RuCO)2 B3H 7 ] (left) and the novel bridged borylene complex [3 -BH)(Cp*RuCO)2 ( -H)( -CO){Mn(CO)3}] (right, perspective from above the triply bridging borylene ligand). H H B H H3C B B OC (CO)3 Mn H H H3C H CH3 H3C CH3 Ru CH3 Ru C O H OC CH3 H3C CH3 H3C CH3 H3C CH3 CH3 Ru Ru H3C H3C HB H OC OC CH3 H3C H3C CH3 The reaction of B2 H 6 (Cp*RuCO)2 and Re 2 (CO)10 provides [Cp*Ru(CO)( -H)]2 (below) as the only characterized product. H3C CH3 H 3C OC H3 C CH3 H Ru Ru H CO H C 3 H3C CH3 CH3 Copyright © 2014 Pearson Education, Inc. CH3 226 15.29 Chapter 15 Parallels Between Main Group and Organometallic Chemistry The cluster is below, with each Mo an abbreviation for ( 5 -Cp*Mo) . The presence of molybdenum—molybdenum bonds is postulated to prevent BH capping of some of the cubane faces. The two boron atoms labeled as “1” are five-coordinate and are assigned a 11B{1H} NMR chemical shift of 131.5 ppm. The boron atoms labeled “2” are four-coordinate with bonds to three Mo atoms and one boron atom; the corresponding chemical shift for these two atoms is 94.4 ppm. The three boron atoms labeled “3” are four-coordinate with two bonds each to boron and molybdenum; the chemical shift for these three boron atoms is 58.8 ppm. H 3 B H 3 B Mo 1 HB BH 1 Mo HB 3 2 Mo B H Mo B H 2 This cluster features four 5 -Cp* ligands, and pairs of these are chemically equivalent on the basis of the 1H NMR spectrum. The “top” two Mo atoms feature bonds to five boron atoms and one Mo atom. The “bottom” two Mo atoms feature bonds to four boron atoms and two Mo atoms. These provide two unique chemical environments for the 5 -Cp* ligands. 15.30 Product yields for insertion of the dicarbollide cage into [( 6 -arene)Fe][PF6 ]2 were deemed lower with the pentamethylbenzene and hexamethylbenzene iron complexes as reactants (15-20%, compared to 48-70% with the less substituted benzenes) on the basis of stronger Fe—arene bonds with these more highly methylated cations. A variety of approaches were used to show correlations to the degree of methylation of the benzene. Figure 1 of the reference shows that the diamagnetic shielding of the CH cage hydrogen atoms increases with increasing number of arene methyl groups. The resonances for these CH hydrogen atoms shift upfield as the number of methyl substituents changes from 1 to 6. Similar linear correlations were observed between the 11 B chemical shift of the antipodal dicarbollide cage boron atom (labeled as 12 in Scheme 1) and the 1H chemical shift of the hydrogen atom on this same boron atom; as the degree of methyl substitution increases at the arene ligand, these resonances progressively shift upfield. It is remarkable how sensitive the latter chemical shifts are to modulation of the donor ability of the arene ring since these atoms are separated from the iron atom by a relatively large number of bonds. The Fe2+/Fe3+ redox potential decreases linearly as the number of methyl substituents increases; the complexes become progressively easier to oxidize (the complexes become better reducing agents) as the number of methyl groups increases. Copyright © 2014 Pearson Education, Inc. Chapter 15 Parallels Between Main Group and Organometallic Chemistry 15.31 15.32 227 a. Ge94– has 40 valence electrons = 4n + 4. Its classification is nido. b. InBi32– has 3 + 3(5) + 2 = 20 valence electrons = 4n + 4. Its classification is also nido. c. Bi82+ has 8(5) – 2 = 38 valence electrons = 4n + 6. It is an arachno cluster. The reference notes that significant ion-pairing exists between the potassium ions and the highly negatively charged Zintl ions in K 3E7 (E = P, As, Sb). Addition of cryptand[2.2.2] results in potassium ion complexation, attenuates the ion-pairing interaction, and renders the Zintl ions more effective nucleophiles. The IR spectroscopic data in the carbonyl region for [K(2.2.2)]3[E7Cr(CO)3 ] (below) indicate extensive -backbonding in each case. On the basis of the lowest energy CO stretching frequency, one would tentatively rank the E73 donor ability to the Cr(CO)3 fragment as increasing as Sb < P < As, but these frequencies are sufficiently similar that it is difficult to clearly differentiate the donor ability of these Zintl ions. Salt IR (CO) (cm -1 ) [K(2.2.2)]3[P7Cr(CO)3] 1829, 1738, 1716 [K(2.2.2)]3[As7Cr(CO)3 ] 1824, 1741, 1708 [K(2.2.2)]3[Sb7Cr(CO)3 ] 1823, 1748, 1718 The electronic spectra of [K(2.2.2)]3[E7Cr(CO)3 ] are provided in Figure 5a of the reference. The tails of the absorptions assigned to intra-ligand E73 charge transfer bands decrease in energy from PCr > AsCr > SbCr. Note that the max of these charge transfer bands (Table 4) are nearly identical, although they do exhibit differences in their molar absorptivity constants, , in L L ); AsCr: 365 nm, (11400 ); SbCr: 365 nm, parentheses: PCr: 363 nm, (6600 mol cm mol cm L (16000 ). The assignment of these intra-ligand charge transfer bands in mol cm [K(2.2.2)]3[E7Cr(CO)3 ] is reasonable since the identical trends are exhibited by the Zintl ions E73 themselves; these absorptions are essentially unchanged upon coordination to the Cr(CO)3 fragment. A sketch of [Cr(CO)3 (HSb7 )]2- is below. The hydrogen ion binds to the Sb atom furthest from the Cr atom. 2OC CO CO Cr Sb Sb Sb Sb Sb Sb Sb H Copyright © 2014 Pearson Education, Inc. Chapter 15 Parallels Between Main Group and Organometallic Chemistry 228 15.33 The Lewis structure of [HP7 ]2- is below. It is interesting that this Zintl ion exhibits fluxional behavior in solution that renders all the phosphorus atoms equivalent on the NMR time scale. P P P P H P P P The most likely carbodiimide hydrophosphination mechanism is as follows, with nucleophilic attack at the carbodiimide carbon affording an intermediate amidinate that rapidly abstracts a proton from the Zintl ion cage. Evidence for intramolecular proton transfer is provided by isotopic labeling studies. When [HP7 ]2- is used as a reactant towards carbodiimides in deuterated-solvents, the only amidine-functionalized products obtained feature N—H bonds. When the corresponding reactions with [DP7 ]2- are conducted in protio-solvents, only amidinefunctionalized products with N—D bonds are obtained. These experiments definitively prove the role of [HP7 ]2- as the proton source, and strongly suggest the mechanism below. NR P P P P P P P RN H C NR P P P N H P P nucelophilic attack C R P P intramolecular proton transfer NR P P P P P P 15.34 a. m = 1 (a single polyhedron) n = 5 (each B atom counts) o = 0 (no bridging atoms) p = 2 (2 missing vertices) 8 skeletal electron pairs b. m = 2 (2 polyhedra) n = 10 (each B, C, and Co atom counts) Copyright © 2014 Pearson Education, Inc. P C HN R Chapter 15 Parallels Between Main Group and Organometallic Chemistry 229 o = 1 (1 bridging atom, the cobalt) p = 2 (2 missing vertices: the top part of the molecule is considered as a pentagonal bipyramid with the top vertex missing, and the bottom as an octahedron with the bottom vertex missing) 15 skeletal electron pairs c. m = 2 (2 polyhedra) n = 17 (each B, C, and Fe atom counts) o = 1 (1 bridging atom, Fe) p = 1 (1 missing vertex, the top atom of the incomplete pentagonal bipyramid) 21 skeletal electron pairs 15.35 15.36 15.37 a. C2v b. C5v, D5h c. [Re2Cl8] : D4h; [Os2Cl8] : D4d d. D2h e. C5v 2– 2– f. Cs, C2h g. Te62+: D3h Ge94–: C4v a. Td b. Ih a. The group orbitals are derived primarily from the 3p orbitals of phosphorus, which collectively resemble the five π orbitals for C5H5 in Figure 13.22. Diagrams of these orbitals can be found in the second reference. (See also Z-Z. Liu, W-Q. Tian, J-K. Feng, G. Zhang, and W-Q. Li, J. Phys. Chem. A, 2005, 109, 5645.) Atomic orbitals on transition metals suitable for interaction (assuming metal centered below P5 plane) lowest energy group orbital: s, pz, dz2 1-node degenerate pair: px, dxz, py, dyz 2-node degenerate pair: dxy, dx2–y2 b. The molecular orbitals of P5– are lower in energy than the similar orbitals of C5H5–, giving rise to a generally stronger ability of P5– to π accept. For example, the energy match between 2-node orbitals and metal d orbitals may be closer in P5– complexes than in the case of ferrocene (Figure 13.28), enabling stronger interaction. An example of Copyright © 2014 Pearson Education, Inc. Chapter 15 Parallels Between Main Group and Organometallic Chemistry 230 such an interaction would be between an empty 2-node orbital of P5– and a dx2–y2 orbital of a metal: dx2–y2 c. The reference provides an energy level diagram of the molecular orbitals of [(5-C5H5)2Ti]2–. In analyzing the orbitals it is important to see how the π orbitals of P5– ligands match up with d orbitals on Ti and how the resulting shapes illustrate how the lobes of these interacting orbitals merge (in the bonding orbitals) and how nodes are formed between them (in the antibonding orbitals). 15.38 The three types of interaction, , π, and , should be evident in the orbitals generated. In addition to the bonding interactions (see Figure 15.8) there should be matching antibonding interactions, with a nodal plane bisecting the metal–metal bond. The relative energies of the molecular orbitals should be similar to those on the right side of Figure 15.9, depending on the level of sophistication of the calculations used. 15.39 In addition to the , π, and interactions between d orbitals, shown in Figure 15.8, a interaction could occur between s orbitals of the two atoms: (In a chromium atom the valence 4p orbitals are empty, so interactions between them need not be considered.) The extent of the calculated interactions between the atomic orbitals is likely to differ significantly. Consequently, even though there may be six possible orbital interactions, the strength of such bonding is calculated to be significantly weaker than would be expected in a true “sextuple” bond. It is suggested that, for this exercise, the Cr–Cr distance be set at 166 pm, the equilibrium distance reported in the references. 15.40 a. The A1g orbital should have a very large, nearly spherical lobe in the center of the cluster and six smaller, also nearly spherical lobes, on the outside of the borons (centered on the hydrogens). b. The T1u orbitals should each have two regions where lobes of p orbitals on four B atoms merge, plus additional lobes, nearly spherical, centered on opposite hydrogens. c. The T2g orbitals should show how the lobes of adjacent p orbitals merge. The result should have four lobes, somewhat similar in appearance overall to a d orbital. Copyright © 2014 Pearson Education, Inc. Chapter 15 Parallels Between Main Group and Organometallic Chemistry 15.41 15.42 231 The orbitals are naturally much more complex, with 167 orbitals and 270 electrons. The T1u and T2g orbitals shown are from a Scigress Extended Hückel calculation. a. The A1g orbital should be similar to that for B6H62– in Problem 15.31, with one large lobe in the center and smaller d orbital lobes on each of the Ru atoms. It is the HOMO, antibonding in symmetry, and the only orbital near this energy that involves carbon orbitals. b. The T1u orbitals should also be similar to the T1u orbitals of B6H62–. There should be two large lobes, each derived primarily from d orbitals on four Ru atoms and a p orbital of the central carbon, plus lobes on opposite Ru atoms (the other Ru atoms). There may also be fragment lobes of d orbitals on the Ru atoms. c. The T2g orbitals are not directly involved in bonding with the carbon, but they do strengthen the cluster framework. The atomic orbitals should interact in sets of four, similarly to the T2g orbitals in B6H62–, but some of the distinctive features of d orbitals should also be observable in the ruthenium cluster. a. As in ferrocene, important interactions should be observed between 1-node group orbitals on the (in this case, Ge5) ligands and dxz and dyz orbitals on cobalt (see example for ferrocene above Figure 13.27). With the Ge5 ligands held in closer proximity by bonds between the top and bottom rings, interactions should also be found between 2-node group orbitals and the dxy and dx2–y2 orbitals of Co and between 1-node group orbitals and px and py orbitals of Co. b. In contrast to the cyclopentadienyl rings in ferrocene, primarily sigma bonding should be observed within the Ge5 rings; the reported Ge–Ge distance in the rings is within the range of single bonds. c. In addition to the comparisons mentioned above, it would be informative to see if in the [CoGe10]3– cluster the interactions between the rings and the dz2 and pz orbitals of Co are weak (because the lobes of these orbitals point toward the center of the rings) and to look for interactions involving the “doughnut” of the dz2. If the software generates energies of the molecular orbitals, additional comparisons can be made between the energies of the cluster orbitals and those of ferrocene shown in Figure 13.28. Copyright © 2014 Pearson Education, Inc.