— 11.1.}; .1. .2. .§:,...,,.Efi+mfi . .. . y hr$ .1. $3.» I: a v3 . .. . .n. in: ‘ 4%“... as. .n!’ . 5‘ .m .9 5 5 L..." as n 5.7”... 3.31 :12: p. i 52...... 9~’{I'?'plln£ I. 1).. . .93.! l. a . :‘mew. ha. .P..3.I‘o: .lt! . 5‘ . v» .. t... n r!) “fir... .: .. i I ;K £1155fi st! .. I t. 3 :1 .1. a... (1.72.... N. (\rlr ii... at .14.. hark unfit! .5. 511““..ch ; .- Dal}. . an. 1.170.. . 9.6 . . txft,.:..so"¥a 9&1. I {nfi hurd- .. i5-..a:.§.«wm$-UM¥1 .l...§.!..st:)snfl‘d.lz.ls# “MS all..rur..r..15. . .531... arkg3uxtfi.§hu _ $.4..9.;finunh£§§xf 1.1.. 1...»?! , {zill1v1fl flint}: v . to. f, I 80““... 2 . 2. .93....xmtkunfimfful .31; .1. .MH. 1.311.! ..E1axf.£¥iN-| :9 iRWfi-‘flrhv t u. input; .0!!! $0.1 i. .e .2...» 21...! has»)! 44.21....) I: [Ab-ior\ - .E. 1.1.5.3.. , .2 95.21 . 14‘s.; 23).??on la . V €2.53: 1...: $1.7. 1.11 l.‘;illl ‘ .ullttly: K x 7‘ I‘llt‘ltl .nuv.“ Iv..- 1.x. Ii 5;")?! 90. ..¢1\lftst. THE§S 4' ,0: O LIBRARY Michlgan State University This is to certify that the dissertation entitled TOPOCHEMICAL ASSEMBLY OF COVALENT MATERIALS USING DIHYDROGEN BONDING presented by Radu Custelcean has been accepted towards fulfillment of the requirements for Ph . D. degree in Chemistry f 0/». / Major profesch Date 7 Aujgét LOCO MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE JUL 0 2 2002 11m elem-gasp.“ TOPOCHEMICAL ASSEMBLY OF COVALENT MATERIALS USING DIHYDROGEN BONDING By Radu Custelcean A DISSERTATION Subnfinedto Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHYLOSOPHY Department of Chemistry 2000 ABSTRACT TOPOCHEMICAL ASSEMBLY OF COVALENT MATERIALS USING DIHYDROGEN BONDING By Radu Custelcean This study explores the potential of the unconventional hydrogen bonds (dihydrogen bonds) between hydridic hydrogens in X-BH3' (X = H, CN) and traditional AH (A = O, N) proton donors to serve as preorganizing interactions for the topochemical assembly of covalent materials. Such weak A-H---H-B interactions can indeed be used to organize and hold a structure’s form while it is more firmly fastened together by A-B bond formation, transferring thus the initial order from the starting crystal to the newly formed covalent frame. This strategy makes dihydrogen bonding a potentially powerful tool for the rational assembly of crystalline covalent solids with controlled structures and properties, a class of compounds generally accessible only by empirical methods or serendipitous discoveries. Evidence of such crystal lattice control in the transformation O-H---H-B —) O-B + H2 was first obtained in studies of the solid-state structures and reactivities of triethanolamine (TEA) complexes with various metal borohydrides (MBH3X-TEA: M = Na, X = H; M = Na, X = CN; M = Li, X = H). Thus, solid-State decomposition of N aBH4-TEA, whose crystal structure shows multiple dihydrogen bonds between the BH.{ ions and the OH groups from TEA, is topochemical, leading to a polymeric trialkoxyborohydride, in sharp contrast to the N aBH4 and hydride-free polymeric borate disproportionation products obtained from decomposition in solution or melt. With its less basic hydridic sites, the dihydrogen-bonded N aNCBH3-TEA complex could not be decomposed in the solid state, losing H2 only at ca. 100 °C above its melting point. Thus, in addition to close H-H contacts, the relative acidity/basicity of the proton/hydride partners is also critical for the reactivity of the dihydrogen-bonded systems. The Hn-H interactions were conveniently tuned by substituting Li+ for Na” in the NaBH4-TEA complex, which led to shorter H-H contacts and enhanced solid-state reactivity, as a result of stronger complexation by Li+ and consequently higher acidity of the OH sites. Mechanistic investigations using in Situ solid-State llB NMR, optical microscopy, FT-IR, and XRD, found that, like most solid-state reactions, the topochemical decomposition of LiBH4-TEA is heterogeneous, with nucleation of the product phase from the parent crystal. Kinetic analysis and HID exchange experiments established that proton transfer along the H---H bond, at the reactant/product interface, is the rate limiting step, with activation parameters AH" and AS" of 20.1 :i: 2.4 kcal/mol and -16.8 i 6.2 cu, comparable with the analogous values found for the aqueous hydrolysis of BH4‘ in neutral water. Finally, the concept of crystallinity transfer from dihydrogen- to covalent-bonded crystals was demonstrated by the crystal-to-crystal decomposition of NaBH4-THEC (THEC = N,N’,N’ ’,N’ "-tetrakis-(2-hydroxyethyl)cyclen), in which self-assembly into dihydrogen-bonded closed-loop large globular dimers minimized the unit cell Shrinkage accompanying the O-H---HB to 0-3 conversion. Further elaboration of these dihydrogen- bonded building blocks offers real prospects for the rational assembly of extended H-c-H bonded, and ultimately covalent solids with controlled architectures. TABLE OF CONTENTS LIST OF TABLES .............................................................................. v LIST OF FIGURES ............................................................................. vi 1. Dihydrogen Bonding: A Literature Review ........................................... l 1 . 1. Introduction ................................................................................. 1 1.2. Structural and Energetic Characterization .............................................. 2 A. Dihydrogen Bonding in the Main Group Hydrides ...................... . .......... 2 B. Dihydrogen Bonding in Transition Metal Hydrides ................................ 14 1.3. Self-Assembly of Extended Dihydrogen-Bonded Systems ........................... 25 1.4. Dynamics of Dihydrogen Bonds ......................................................... 29 2. Organic Solid-State Reactions: Mechanisms and Concepts ........................ 46 2.1. Topochemical Polymerizations and Polycondensations ............................... 51 2.2. Hydrogen Bonding Assisted Topochemical Reactions ................................. 55 2.3. Kinetics of Solid—State Decompositions ................................................. 60 3. Topochemical Assembly of Covalent Materials Using Dihydrogen Bonding 64 3.1. Results and Discussion ..................................................................... 67 3.2. Conclusions .................................................................................. 1 19 3.3. Experimental Section ....................................................................... 122 BIBLYOGRAPHY ............................................................................. 145 iv LIST OF TABLES Table 3.1. NMR data for BHx(OCH3)4-x', 47, and 48 ...................................................... 85 Table 3.2. IR data for LiBH4o TEA (47), LiBD4-TEA (47a), LiBH4(50%D)-TEA (47b), and their corresponding solid state decomposition products, 48, 48a, and 48b. LiBH4 is included for comparison .................................................................................................. 87 Table 3.3. Rate constants k for the solid state decomposition of 47 , calculated from the Avrami-Erofeev law for nucleation and two-dimensional growth in the specified conversion ranges, together with the correlation coefficients of the linear regression analysis R2 ....................................................................................................................... 94 Table 3.4. Crystal data and structure refinement for N APA H3BCN (44) ..................... 125 Table 3.5. Crystal data and structure refinement for N aBH4-TEA (45) ......................... 126 Table 3.6. Crystal data and structure refinement for N aN CBH3'TEA (46) ................... 127 Table 3.7. Crystal data and structure refinement for LiBH4-TEA (47) .......................... 128 Table 3.8. Crystal data and structure refinement for NaBI-I4-THEC (49) ...................... 129 Table 3.9. Crystal data and structure refinement for the decomposition product of NaBH4-THEC in DMSO ................................................................................................ 130 Table 3.10. Crystal data and structure refinement for N aNCBH3'THEC (50) .............. 131 Table 3.11. Crystal data and structure refinement for NaBH4-THPC (51) .................... 132 Table 3.12. Crystal data and structure refinement for 9-(1’,4’,7’-triS-hydroxyethyl- cyclen)methylanthracene (52) ........................................................................................ 133 Table 3.13. Crystal data and structure refinement for N aBI-I4-[9-( l ’,4’,7’-tris- hydroxyethylcyclen)methylanthracene] -EtzO (53) ....................................................... 1 34 LIST OF FIGURES Figure 1.1. O-H---H-B dihydrogen bonds found in N3BH4'2H20 in the solid-state ........ 3 Figure 1.2. C2 isomer of the N H3BH3 dimer .................................................................... 4 Figure 1.3. C2}. isomer of the NH3BH3 dimer .................................................................. 5 Figure 1.4. The shortest N-Hu-H-B contacts found in NH3BH3 in the solid-state .......... 6 Figure 1.5. Dihydrogen-bonded dimers found in (CH3)2NH-BH2-N(CH3)2-BH3 in the solid-State ......................................................................................................................... 6 Figure 1.6. C-H---H-B close contacts in 7-10 ................................................................. 9 Figure 1.7. Calculated structure of the NH3A1H3 dimer ................................................ 10 Figure 1.8. Calculated structure of (NH2A1H2)3 ............................................................ 10 Figure 1.9. Calculated structure of the (NH2A1H2)3 dimer ............................................ 11 Figure 1.10. Self-assembly of cyclotrigallazane in the solid-state ................................ 12 Figure 1.11. Calculated structure of the (NHzGaH2)3 dimer ......................................... 13 Figure 1.12. Calculated structure for dihydrogen-bonded complex 26a; < FH---H-Mo = 127°, < F—H---HMo = 174° ............................................................................................. 23 Figure 1.13. Generic representation of C-H---H-M dihydrogen bonds in complexes of transition metal hydrides with R3-x(Ph)xP ligands ......................................................... 24 Figure 1.14 Self-assembly of dihydrogen-bonded complex 28 in one-dimensional chains ............................................................................................................................. 26 Figure 1.15. Self-assembly of dihydrogen-bonded complexes 29-[K(1,10-diaza-18- crown-6)] in one-dimensional chains ............................................................................ 27 vi Figure 1.16. Self-assembly of dihydrogen-bonded complexes 29a,b-[K(THF)(1-aza-18- crown-6)] in one-dimensional chains ............................................................................... 27 Figure 1.17. Self-assembly of dihydrogen-bonded complexes 29b-[K(THF)(18-crown- 6)] in one-dimensional chains ......................................................................................... 28 Figure 1.18. Self-assembly of guanidinium borohydride in one-dimensional extended tapes ................................................................................................................................ 28 Figure 3.1. X-Ray Crystal Structure of NAPA H3BCN Showing the H---H contacts in A ...................................................................................................................................... 69 Figure 3.2. X-Ray crystal structure of NaBH4-TEA: (a) coordination of Na+; (b) dihydrogen bonds connecting the chains, with H-H contact distances in A .............. 72 Figure 3.3. Progress of N aBH4-TEA decomposition monitored by X-ray powder diffraction: (a) initial complex; (b) high-temperature polymorph; (c) final decomposed material ............................................................................................................................ 75 Figure 3.4. X-Ray crystal structure of N aNCBH3-TEA: (a) coordination of N a+; (b) dihydrogen bonds connecting the chains, with H-H contact distances in A .............. 79 Figure 3.5. X-Ray crystal structure of LiBI-h-TEA showing the H~--H close contacts in A ...................................................................................................................................... 81 Figure 3.6. X-Ray powder diffraction pattern of LiBH4-TEA (a) and its solid-state decomposition product (b) ............................................................................................... 83 Figure 3.7. Calculated structure of LiBH(OCH3)3 with the interatomic distances in A.. 89 Figure 3.8. Typical microscopic view (transmitted light) of LiBH4-TEA solid-state decomposition: (a) initial crystal, (b) 10 min at 110 °C, (c) final decomposed crystal; Scale bar: 100 um ............................................................................................................ 91 vii Figure 3.9. Conversion (0t) vs. time curves for the solid-state decomposition of 47: O = 105°C;CI=110°C;A=115°C;X=120°C ................................................................. 92 “2 Figure 3.10. Plots of the Avrami-Erofeev law, [-ln(l-a)] against time, for the solid state decomposition of 47: O = 105 °C; C] = 110 °C; A = 115 °C; x = 120 °C .............. 93 Figure 3.11. Arrhenius plot for the solid-state decomposition of 47 at 105-120 °C ....... 95 Figure 3.12. Variation of activation energy with conversion for the solid-state decomposition of 47 , obtained using the isoconversional method .................................. 95 Figure 3.13. Proposed mechanism for the first O-H---H-B to 0-8 topochemical conversion in 47 ............................................................................................................... 97 Figure 3.14. X-Ray crystal structure of NaBH4-THEC illustrating the self-assembly in dihydrogen-bonded dimers .............................................................................................. 99 Figure 3.15. Crystal packing in NaBH4oTHEC .............................................................. 100 Figure 3.16. Polarizing microscopic view of NaBH4~THEC crystal-to-crystal decomposition: a) initial crystal; b) after 20 h at 160 °C. Scale bar = 100 um ............... 101 Figure 3.17. Powder XRD patterns for: a) dihydrogen-bonded complex 49; b) solid-state decomposition product; 0) product of decomposition in the melt .................................. 102 Figure 3.18. Evolution of unit cell axes (a) and volume (b) during the solid-state decomposition of NaBH4-THEC at 120 °C, monitored by powder XRD ....................... 103 Figure 3.19. Molecular model of the product from the solid-state decomposition of 49; selected bond lengths [A] and angles [°]: B-O(1) 1.458, B-O(2) 1.482, B-O(3) 1.497, B- O(4) 1.474, O(1)-B-O(2) 112.5, O(l)-B-O(3) 109.1, O(l)-B-O(4) 114.6, B-O(1)—C 123.7, B-O(2)-C 122.3, B-O(3)-C 125.9, B-O(4)-C 125.6, N(l)-Na 3.003, N(2)-Na 2.714, N(3)- Na 2.435, N(4)-Na 2.565, O(2)-Na 2.670, O(3)-Na 2.276, O(4)-Na 2.228 .................... 105 viii Figure 3.20. X-Ray crystal structure of the major conformation found in the covalent dimer obtained by decomposing 49 in DMSO; selected bond lengths [A] and angles [°]: B-O(l) 1.502(2), B-O(2) 1.511(2), B-O(3) 1.450(2), B-O(5) 1.457(2), O(1)-B-O(5) 111.6(1), O(2)-B-O(5) 109.8(1), O(3)-B-O(5) 111.6(1), B-O(1)-C 121.7(1), B-O(2)-C 113.2(1), B-O(5)-C 122.8(1), N(1)-Na 2.519(1), N(2)-Na 2.540(1), N(3)-Na 2.565(1), N(4)-Na 2.682(1), O(1)-Na 2.371(1), O(2)-Na 2.796(1), O(3)-Na 2.391(1), O(4)-H 082(2), O(4)H---O(2) 1.92(3), O(4)---O(2) 2.735(4), O(4)-H-O(2) 175.1(3) ................. 106 Figure 3.21. Conversion (0t) vs time curve for the solid-state decomposition of NaBI-h-THEC at 160 °C, obtained using in situ "B MAS NMR ................................... 108 Figure 3.22. Lattice conversion (VO'VNO‘Vf) vs time curve for the solid-state decomposition of N aBH4-THEC at 120 °C, obtained using powder XRD ..................... 108 Figure 3.23. X-Ray crystal structure of NaNCBH3-THEC ............................................ 110 Figure 3.24. or-Helix formed by NaNCBHyTHEC in the solid-state ............................ 111 Figure 3.25. X-Ray crystal structure of NaBH4-THPC; dihydrogen bonding parameters (A, °): O(4)H---HB(1) (lower) 2.106, O(4)H---HB(1) (upper) 2.266, O(3)H---HB(1) 2.005, O(7)H---HB(2) 1.927, O(8)H---HB(2) 1.676, O(2)H---HB(2) 1.904, O(4)H-H-B(1) (lower) 96.1, O(4)H-H-B(l) (upper) 88.7, O(3)H-H-B(1) 113.8, O(7)H-H-B(2) 110.7, O(8)H-H- B(2) 120.2, O(2)H-H-B(2) 133.3, O(4)-H-HB(1) (lower) 162.5, O(4)-H-HB(1) (upper) 143.9, O(3)-H-HB(1) 153.4, O(7)-H-HB(2) 155.4, O(8)-H-HB(2) 153.0, O(2)-H—HB(2) 169.3 ................................................................................................................................ 113 Figure 3.26. X-Ray crystal structure of 9-(1’,4’,7’-triS-hydroxyethyl- cyclen)methylanthracene ................................................................................................ l 15 ix Figure 3.27. X-Ray crystal structure of 53; normalized dihydrogen bonding parameters (A, °): O(1)H~-~HB 1.692, O(2)H---HB 1.661, O(3)H---HB 1.730, O(1)H-~H-B 114.5, O(2)H---H-B 117.8, O(3)H---H-B 116.1, O(1)-H---HB 166.8, O(2)-H-~-HB 155.1, O(3)- H---HB 154.6 ................................................................................................................... 116 Figure 3.28. Prospective multidentate ligands for the construction of extended dihydrogen-bonded and covalent networks .................................................................... 118 l. Dihydrogen Bonding: A Literature Review 1.1. Introduction Hydrogen bondingl occupies a prominent position in modern chemistry. It is fundamental for molecular recognition and supramolecular synthesis, and holds a central role in biology.2 Typically, this noncovalent interaction occurs between the positively charged hydrogen of an A—H (A = O, N, halogen, C) proton donor, and the lone pair of an electronegative element, or the It electrons of a multiple bond or aromatic ring, representing the proton acceptor. Recently, an unusual type of hydrogen bonding, in which a o M-H bond (where M is less electronegative than H) acts as the electron donor, has attracted considerable attention.3 With strength and directionality comparable with those found in conventional hydrogen bonding, this proton-hydride interaction, or dihydrogen bonding, can influence the structure, reactivity, and selectivity in solution and solid state, thus finding potential utilities in catalysis, crystal engineering, and materials chemistry. This review summarizes the emergence and development of this topic, starting with the early observations of H---H bonding, continuing with its comprehensive structural and energetic description, and concluding with the implications of this interaction in supramolecular synthesis, as well as its influence on reactivity and selectivity in both fluid and condensed phases. 1.2. Structural and Energetic Characterization A. Dihydrogen Bonding in the Main Group Hydrides. Chemists had intuitively thought about proton-hydride interactions long before these associations were formally categorized as hydrogen bonds. In 1934, Zachariasen and Mooney reported the crystal structure of ammonium hypophosphite (NI-14+H2P02’),4 and noted that “the hydrogen atoms of the hypophosphite group behave toward ammonium as if they were H' ions".5 Thirty years later, Burg suggested the presence of N-H---H3B interactions comparable to hydrogen bonds in liquid (CH3)2NH-BH3, based on the perturbation of the N-H and B-H bands in the IR spectrum.6 Titov et a1. explained the enhanced chemical reactivity of aminoboranes toward H2 loss by the “close spatial arrangement of the oppositely-charged hydrogen atoms”.7 However, the first to recognize this interaction as a true hydrogen bond were Brown and coworkers in the late 1960’s.8 Based on a thorough analysis of the IR Spectra of the boron coordination compounds of the type L-BH3 (L = Me3N, Et3N, Py, Et3P) and Me3N-BH2X (X = Cl, Br, I) in the presence of proton donors such as MeOH, PhOH and p-F-C6H4-OH in CC14, they proposed the formation of a novel type of hydrogen bond in which the BH3 and BHz groups acted as proton acceptors, despite their lack of lone pairs or 1: electrons. They measured the strengths of these interactions by variable temperature IR spectroscopy, finding association energies in the range of 1.7-3.5 kcal/mol, comparable with moderately strong conventional hydrogen bonds.9 Similarly, they inferred the occurrence of intermolecular N H---HB hydrogen bonding in MCzNH°BH3 and (RNH-BH2)3 (R = Pr, Bu), from the temperature and concentration dependence of their N-H Stretching absorptions in CC14.8 In a recent study, Epstein and coworkers confirmed the ability of boron hydrides to act as proton acceptors in hydrogen bonds.10 They Studied the interaction of neutral NEthH3 and P(OEt)3BH3 as well as ionic Bu4N+BH4' with different alcohols as proton donors, by IR and NMR spectroscopy in CHzClz, C6H14 and C6D12, and concluded that the properties of these unconventional OH---HB interactions are similar to those found in classical hydrogen bonds. Their association energies were found to increase proportionally with the proton donors’ acidities, being situated in the range 1.1-3.6 and 2.5-6.5 kcal/mol, for the neutral and ionic boron hydrides, respectively. Theoretical calculations (RHF/6-3 16) confirmed the attractive nature of these proton-hydride interactions. The solution studies, however, cannot unambiguously establish whether these unusual interactions involve the boron atom, the hydridic hydrogen, or the BH group as a whole. The X-ray and neutron crystal structures of NaBH4-2HZO were recently determined by Jackson et al., in order to probe the existence of O-H---H-B dihydrogen bonding in the solid state, and provide a detailed structural description of it (Figure 1.1).ll H_. H ’0 H l O H B\ .uHr ‘H H/ H +3 O + H LnM:::OR ——-> LnM<—-|_I_' 'OR Scheme 1.1 14 The first unequivocal evidences of dihydrogen bonding involving a transition metal hydride came independently from the groups of Crabtree and Morris in 1994. The X-ray crystal structure of 14 determined by Crabtree et al. showed an unexpected tautomerization of the amide group from the ligand into the iminol form, facilitating thus the interaction between the hydridic Ir-H and the OH proton.27 While the exact H-H distance could not be determined in the solid state due to the failure to locate the Ir-H hydrogen, 1H NMR T1 relaxation time measurements in solution gave a HmH contact distance of about 1.8 A. The interaction seems to have some covalent character, as demonstrated by the observed coupling of 3 Hz between Ir-H and O-H hydrogens. + _ —] SbF6 (x = PPha) An even higher coupling of 5.5-5.6 Hz was found in 15 (Y = H, Cl, Br, I), in which the HmH contact distance was estimated around 1.7 A (Y = Cl), based on T1 relaxation time measurements.28 a mail? .0 Y—lr-H’ I'l'l\PPh3 15 15 These H---H interactions are fairly strong, competing favorably with the conventional Y---H hydrogen bonds, as demonstrated by the predominance of the O-H---H-Ir dihydrogen-bonded rotamers over the opposite O-HmY-Ir hydrogen-bonded ones. A quantitative estimation of the strength of these unconventional hydrogen bonds involving iridium hydride complexes was elegantly made by Crabtree’s group in the analogous 2- aminopyridine complexes 16 (Y = H, F, Cl, Br, 1, CN, CO). They measured the rotation barrier for the NH; group, which represents the sum of the HuH interaction energy and the intrinsic rotation barrier around the C-N bond in the ligand. The strongest N-H---H-Ir hydrogen bond (5.0 kcal/mol) was found for the case where Y = H, while the trans Y ligand tends to weaken the interaction in the order: F > C1 > Br > I > CN > CO > H, which can be rationalized by their decreasing electronegativity in the same order. Morris and coworkers demonstrated the presence of Hull interactions in the iridium hydride complex 17 by X-ray diffraction in the solid-state, and NMR spectroscopy in solution.29 The crystal structure of 17 indicated a close contact between the pyridinium protons and the Ir—H hydridic hydrogens, which unfortunately could not be located precisely from the electron density difference maps. Nevertheless, their lH NMR data provided clear evidence for N-H---H-Ir hydrogen bonding in CD2C12, with a 16 H-H contact distance of about 1.75 A, calculated from the observed T1 relaxation times of the protonic and hydridic hydrogens. S .i’ L H _ L/IIr\H” \ /N-H--~H l7 (L=PCY3) 18 (L=PH3) Theoretical calculations by Hoffmann et al.223 on the model complex 18 confirmed the attractive H~--H interaction, and concluded that its nature is mostly electrostatic. Interestingly, when THF was used as a solvent, the dihydrogen bonds were switched off in 17, presumably by the formation of conventional N—HmO hydrogen bonds with the solvent. Bifurcated N -H---H(Ir)---H—N dihydrogen bonds were also detected by Morris et al. in 19 and 20 by X-ray crystallography and NMR spectroscopy, and their H-H contact distances were estimated around 1.80 and 1.86 A, respectively.30 l7 They also reported an interesting bifurcated Ir-H---H(N)---F-B interaction in complex 21, in which the N-H proton is shared by a hydridic Ir-H hydrogen and a conventional B-F electron donor from the BF4' counterion.31 PPhg " Ir " N— PPha/H\H~H ' / F. / BF3 21 The occurrence of intermolecular X-H---H-M dihydrogen bonds was first documented by Crabtree and his collaborators, with the neutron diffraction crystal structure of the rhenium polyhydride complex 22, which exhibits a three-center interaction between two Re-H hydridic hydrogens and the N -H proton from an indole molecule of crystallization.32 The geometrical parameters for the two HnH contacts, i.e. H-H distances of 1.73 and 2.21 A, and strongly bent NH---H-Re angles of 119 and 97°, respectively, fall in the range expected for dihydrogen bonding. The strength of the interaction was estimated around 4.3 kcal/mol, from the shift of the N-H stretching band in the solid-state IR spectrum of 22, relative to free indole. / LLH\ //L¥ H H HFI 22 18 Theoretical calculations using the DFI‘ method on a [ReH5(PH3)3]-NH3 model confirmed the attractive nature of the interaction, and predicted a similar three-center H-bond with H--H distances of 1.92 and 2.48 A, and an interaction energy of 8 kcal/mol in the gas phase. The intermolecular N-H---H-Re hydrogen bonding appears to be general, as demonstrated by the X-ray crystal structure of the analogous complex 23, in which the role of the proton donor is played by an imidazole molecule.32b’ 33 The H-H distances could not be satisfactorily determined due to the failure to locate the N-H hydrogen from the disordered imidazole. However, the strength of the interaction (5.3 kcal/mol, estimated by IR spectroscopy) is greater than in the similar complex 22, as expected considering the higher acidity of imidazole relative to indole. / N H H\ /N__.// H H NJ H 23 Other weak acids such as 2,4,6-Me3C6HzOH, 2-tBu-6-MeC6H30H, pyrrole, PhNHPh, PhNHBn, PhNHMe, also associate with ReH5(PPh3)3, as shown by the IR of the thin films obtained by evaporation of CH2C12 solutions containing a 1:1 mixture of the polyhydride and the proton donor.34 The ~AH° of the interactions, evaluated from the observed shifts of the NH or OH bands, were found to generally correlate with the acidities of the proton donors, and vary between 3.0 and 5.6 kcal/mol. Replacing the hydridic partner with the WH4(PMePh2)4 complex resulted in slightly weaker dihydrogen l9 bonds of 1.1-5.2 kcal/mol. That these interactions involve primarily the hydridic hydrogen and not the d2 non-bonding electron pairs of the Re or W metals was demonstrated by the analogous d0 complex ReH7(dppe) (dppe = thPCHZCHzPth), that also associated with the same proton donors, with -AHO in the range 1.3-4.7 kcal/mol. In solution, the N -H---H-Re interactions were studied by Crabtree et al. using UV- vis spectroscopy.35 Thus, the ReH5(PPh3)2L (L = pyridine, 4-picoline, 4- dimethylaminopyridine, and 4-carbomethoxypyridine) hydrogen bond acceptors interact with indole as the hydrogen bond donor, showing free energies of association, -AG, of 3.8-5.0 kcal/mol. In a thorough analysis, Epstein, Berke, and coworkers surveyed dihydrogen bonding in solution in the tungsten hydride-alcohol complexes 24.36 Using IR and NMR spectroscopies, they ruled out hydrogen bonding to the CO or NO groups, and proved the exclusive formation of the unconventional O-H---H-W interactions. As expected, the strengths of these dihydrogen bonds increase with the donator abilities of the ligand L (PMe3 > PEt3 > P(OiPr)3 > PPh3) and are directly proportional to the acidities of the proton donors phenol (PhOH), hexafluoro-2-propanol (HFIP), and perfluoro-2-methyl-2- propanol (PFI‘B) (PFI‘B > HFIP > PhOH). Their magnitude was estimated in the range 4.1-6.9 kcal/mol, from the observed shifts in the corresponding Von bands, as well as from the variation of the association constants K with temperature. In addition, NMR experiments (6 shifts, NOE, and T1 relaxation times), all supported the formation of O- H---H-W dihydrogen bonds, with H-H contact distances as short as 1.77 A in the case of HFIP. A linear OH---H-W orientation was arguably suggested for these interactions, in 20 sharp contrast to the previously established propensity of dihydrogen bonds for a strongly bent geometry. CO \ 0N7w'L-H-«H—on 0‘3 L L = PMe3, PEtg, P(iPr0)3. PPh3 R = Ph, CH(CF3)2, C(CF3)3 24 In an analogous series of rhenium hydride complexes (25), the same two research groups proved the occurrence of intermolecular O—H---H-Re dihydrogen bonding in solution.” 38 When PFI'B was used as proton donor, interaction energies between 4.5 and 6.1 kcal/mol were calculated from the observed v0" shifts in the IR spectra in hexane. In toluene, however, the AH values, derived from variable temperature N MR spectroscopy, are smaller by about 3 kcal/mol, apparently due to competitive O-H-o-n interactions with the solvent.38 The H---H contact distances calculated from T1 relaxation times range between 1.78 and 1.94 A. The phosphine ligand appears to have an important influence over the regioselectivity of the H-bonding formation. Thus, while in 25a interaction with one of the hydridic hydrogens is preferred, the NO group competes more and more effectively for the proton donor as the bulk of L increases, to the point where only O-H---ON hydrogen bonds are observed for 25¢. However, DFI‘ calculations on a ReH2(CO)(NO)(PH3)2.H20 model indicated that the H---H interaction is energetically preferred by about 3.0-3.5 kcal/mol.37 Also, a stronger interaction was predicted with the Re-H hydride trans to the NO group (dim = 1.49 A) compared to the Re-H trans to CO 21 (dim = 1.79 A), and confirmed experimentally by the high regioselectivity displayed by both PFI'B and HFIP alcohols toward the former, as shown by NMR spectroscopy.38 L l v” ON—/—Rle——H CO L 25 L = PMe3 (a), PEt3 (b), PiPr3 (c) The experimental results obtained by Epstein and Berke on the intermolecular dihydrogen bonding in solution were complemented by the theoretical work of Scheiner et al. on the Mo and w hydride complexes 26.39 Their HF/3-21G and DFT (BBLYP, BLYP, B3PW91) calculations confirmed that the HmH interactions are favored over conventional hydrogen bonding involving the N 0 group. L M = M0, W , l .~‘CO L= PH3, NH3 '- 7'\|"—H\.\ L’ = NO, Cl, H CO L “R R = F, OH, H20+ 26 The dihydrogen bonds in 26 become stronger and shorter with the increases in the donating ability of the cis-ligand or the acidity of the proton donor, consistent with experiment. However, the strongly acidic H3O+ induces complete proton transfer, resulting in the formation of an nz-Hz dihydrogen complex. The H-H-M angles are 22 strongly bent in all the optimized structures, as illustrated in Figure 1.12 for the representative complex 26a. For this dihydrogen-bonded system, a 2.6 kcal/mol destabilization energy was calculated for a linear F-HmH-Mo orientation. PH3 (co 0N7Mo‘é—H 00 l PH3 H F 263 Figure 1.12. Calculated structure for dihydrogen-bonded complex 26a; < FHmH-Mo = 127°, < F—HmHMo = 174°. As with the boron hydrides, C-H sites may participate as protonic partners with transition metal hydrides. Recent X-ray structural studies and CSD surveys confirmed the existence of intra-40 as well as intermolecular“ C-HmH-M close contacts in transition metal hydrides. A large number of these examples were observed in complexes containing R3-X(Ph)xP (x = 1-3) ligands, in which one or more ortho C-H bonds point toward the M-H hydridic hydrogens (Figure l.13).4°°°’d While the HmH distances and M- HmH angles in these complexes were found to fall essentially in the same range as observed for the more “conventional” dihydrogen bonds involving N—H or O-H proton donors, the C-HmH angles usually tend to be smaller, due to the inherent constraints imposed by the chelation.4oa However, caution is advisable in interpreting some of these 23 C-HmH-M short contacts, as steric compression by bulky ligands or packing forces may make a significant contribution to the observed HmH close proximities.40b Figure 1.13. Generic representation of C-HmH-M dihydrogen bonds in complexes of transition metal hydrides with R3-X(Ph)xP ligands. The comprehensive analysis of the manganese hydride complex 27 carried out by Brammer and collaborators provided convincing evidence for an intramolecular C-HmH- Mn dihydrogen bond.42 Their combined low-temperature neutron and X-ray diffraction study revealed a short intramolecular C-HmH-Mn contact of 2.10 A, with HmH-Mn and C-HmH angles of 126.5 and 129.0°, respectively, and an essentially coplanar relative orientation of the Mn—H and OH bonds ((1 Mn—HmH-C = 07°). The experimental atomic charges found for the Mn-H hydridic and ortho CH protonic hydrogens, of —0.40 and + 0.32, clearly indicate an attractive electrostatic interaction, whose magnitude was calculated to be 5.7 kcal/mol. Moreover, topological analysis of the charge density using the “atoms in molecules” theory unequivocally supported the existence of a moderately strong intramolecular C-HmH-Mn hydrogen bond. 24 CO I __00 00—Mrlm\'—H\ ‘H .. Pco 0 Ph 27 1.3. Self-Assembly of Extended Dihydrogen-Bonded Systems. In the previous section it was demonstrated that dihydrogen bonding is a significant interaction, with energetic, electronic, and spectral characteristics, as well as directionality, comparable with those found in conventional hydrogen bonding. A direct consequence of this similarity is that dihydrogen bonds, like traditional H-bonds, could find potential utility in crystal engineering and supramolecular synthesis. That these HmH noncovalent interactions are indeed capable of controlling crystal packing was already suggested by the crystal structure of cyclotrigallazane (12), which self-assembles into an extended dihydrogen-bonded tut-network in the solid state (Figure 1.10), as demonstrated by Gladfelter’s group.2| The osmium polyhydride complex [K( 1-aza-18-crown-6)][mer- OsH3(CO)(iPr3P)2] (28) synthesized by Morris et al. also forms polymeric one- dimensional chains in the solid-state (Figure 1.14).43 The N-HmH-Os dihydrogen bonds are however complemented by COmK+ interactions in the assembly of the chains. 25 Mm . . o ( x fifPiPr3 ‘ ‘~K’+ _____ _N/H---H—osI—co-~ 0" _____ K )iPrgP/H D0 n 28 Figure 1.14 Self-assembly of dihydrogen-bonded complex 28 in one-dimensional chains. When 1,10-diaza-18-crown-6 was used to complex K+, similar chains held together exclusively by N-HmH-M dihydrogen bonds (Figure 1.15) could be assembled from the osmium, ruthenium, and iridium anionic polyhydrides [MHX+3(iPr3P)2]' (M = Os (3), Ru (b), x = 2; M = Ir (c), x = 1) (29).“ 45 The intermolecular HmH distances in these complexes were estimated from their X-ray crystal structures to be in the range 1.8-1.9 A. The NH bands in their solid-state IR spectra are broadened relative to the free diaza- crown ether, and shifted to lower numbers by 96, 107, and 132 cm'l, for the Os, Ru and Ir complexes, respectively, in accord with the increasing basicity (and hydridicity) of these polyhydrides in the same order. The observed shifts correspond to N-HmH-M interaction energies of about 3 kcal/mol.4S Weaker C-HmH-M hydrogen bonds can also be involved in the supramolecular association of the polyhydrides 29a,b with K(THF)(1-aza-18- crown-6), in which alternating NHmH-M-HmHN and CHmH-M-HmHC units lead to the formation of zigzag chains, as illustrated in Figure 1.16.45 The NHmHOs and NHmHRu 26 separations were estimated at 1.7 A, while the weaker CHmHOs and CHmHRu interactions exhibit longer H-H contact distances of 2.2 and 2.1 A, respectively. /—\ /—\ 0. .0 Q. .0 { ‘ H PiPr3 ( V — H—N— ,Kf -N—H -H—M‘—H---H—N---—--‘K;+--- - —H - x /' x O \O O \__/ \_/ 29 M=OsH,RuH,lr Figure 1.15. Self-assembly of dihydrogen-bonded complexes 29-[K( l ,10-diaza-18- crown-6)] in one-dimensional chains. 0. ‘ .PIPr ( 03 H s ..... .K? ------ o M = Os, Ru ingP H H Figure 1.16. Self-assembly of dihydrogen-bonded complexes 29a,b-[K(THF)( l-aza-18- crown-6)] in one-dimensional chains. 27 Finally, a similar polymeric chain can be assembled exclusively by weak C-HmH-M interactions (Figure 1.17), as illustrated by the crystal structure of 29b-[K(THF)(18- crown-6)].45 In this case, the observed C-H---H-Ru contact distance is 2.2 A. o‘ \ \ ‘. (3 +1 i4 [ll ..uPiPl'3 ( V \Ru'—H\ 4 H iPrapfi/ \H “H O- ---- (K; ------ O x Fad” ) \__/O Figure 1.17. Self-assembly of dihydrogen-bonded complexes 29b-[K(THF)(18-crown- 6)] in one-dimensional chains. As part of early endeavors in the Jackson group directed toward the structural characterization of dihydrogen-bonded systems involving anionic borohydrides, the crystal structure of guanidinium borohydride was explored.46 In the solid-state, this salt is organized into extended tapes, in which the alternating BH4' and C(NH2)3+ ions are connected by multipoint dihydrogen bonds, as illustrated in Figure 1.18. a- .-—-P1 fl--- - H\N’H \3/ “Hm/H” Ike/w“ -’ \ .5 \ \ |+ ,H‘,'H‘ I H ”H ~H\ C\ H ““‘s C” l I H H H H Figure 1.18. Self-assembly of guanidinium borohydride in one-dimensional extended tapes. 28 1.4. Dynamics of Dihydrogen Bonds Numerous studies have now established that dihydrogen bonding is an important and general interaction involving element-hydride 0' bonds, and its geometrical and energetic features have been described in great detail. The significance of these unusual hydrogen bonds extends, however, beyond their fundamental aspects. With their substantial strength and directionality, they can be used to control reactivity and selectivity of chemical reactions, in the same time finding a place alongside conventional hydrogen bonding in the supramolecular chemists’ arsenal of noncovalent interactions. However, what makes dihydrogen bonding particularly interesting is the special reactivity conferred by its peculiar nature. It has been recently demonstrated that HmH bonds have a role in the formation of dihydrogen Ill-H2 complexes and the reverse heterolytic splitting of H2, as well as o-bond metathesis (Scheme 1.2). M—H + H—A —-+ M—H---H—A : HT H M~-I A' ‘——2 M—A +H2 Scheme 1.2 The Ir-Hb bond in 30 has been found to be activated by dihydrogen bonding for a number of reactions.47 Thus, the hydridic and protonic hydrogens l-la and Hb involved in the HmH interaction can interchange relatively easily whereas the non-interacting Hc is 29 exchanged much more slowly with H8 and Hb. The AH‘ for the Ha/Hb exchange has been estimated by variable temperature NMR spectroscopy at around 14-16 kcal/mol and found to go down as the R group becomes more electron-withdrawing, consistent with a mechanism involving proton transfer from the OH group to the Ir-Ha bond, to give an 772- H2 intermediate complex (Scheme 1.3). Rotation of the H2 ligand in this complex and transfer of the proton back to the oxygen completes the exchange. When the reaction was performed in the presence of benzonitrile, the H2 ligand could be displaced by PhCN, in a rate-limiting step. 30 L = PPha R = Me, n-Bu, p-tolyl, Ph, p-FCsH4. 3.4-CeH3F2 Scheme 1.3 Alternatively, H2 elimination from 30 by heating in a sealed tube at 80 °C yielded the chelate complex 30a, in a o-bond metathesis reaction. The initial complex 30 could be recovered by exposure to H2 in CH2Cl2 at room temperature, via the isomer 30b (Scheme 1.4). 30 Scheme 1.4 Similarly, hydrogen scrambling between Ha and H, is facilitated by dihydrogen bonding in complexes 31, apparently via an nz-H2 intermediate.47 H2 loss at room temperature was also observed, with the formation of 31a in a first order reaction, with measured activation parameters AH‘ and AS" of 14 i 2 kcal/mol and —32 :t 6 eu, respectively (Scheme 1.5). The highly negative activation entropy suggests an associative process with a highly ordered transition state. The first direct observation of a dynamic equilibrium between a HmH bonded system and an If—H2 complex resulting from proton transfer along a dihydrogen bond was made by Chaudret and coworkers, using NMR spectroscopy.48 Thus, in the presence of phenol, the ruthenium hydride complex RuI-I2(dppm)2 (32) exists as a mixture of dihydrogen-bonded 31 cis and trans isomers in benzene or toluene solutions. The trans isomer is also involved in a dynamic equilibrium with the dihydrogen complex 32a, which lies 17 kcal/mol lower in enthalpy than 32-PhOH (Scheme 1.6). It was proposed that the reversibility of the process originates in the strong dihydrogen bonding between 32 and phenol. In the presence of the more acidic hexafluoroisopropanol, the corresponding dihydrogen complex 32a further reacts by H2 loss, to ultimately give 32c via 32b. / I \ R -H2 N N’ —* PEP-.2 | ,Ha Ph3P Hd—lr-Hfi _ l\ Hd l Hc PPh3 31 31a R = H, Ph Scheme 1.5 DFT calculations by Scheiner et al. on a HOH--.H2Ru(PH2CH2PH2)2 model yielded AE and Ali“ values of -10.7 and 10.0 kcal/mol, respectively, in qualitative agreement with experiment.49 However, when the stronger proton donor HF was used in the calculations, no F—H---H-Ru adduct could be identified, and the system evolved directly toward the dihydrogen complex, which in this case lies 23.8 kcal/mol lower in energy than the separated HF and ruthenium hydride complex. 32 H-OR / I H I P- EPH" | ..IP/ >ph I ”.H "H—OR "'Ru "" _ a "" Fl /P/| \PZ \p/ |u\H I H | /\/P—- trans-32 \ ,. \l HTH I 7 I 0" I (P """" Ru """ P; QR —>'H2 2PM“; ""P/ . u.. / /| \P\ /P/| \PZ H l I H | 32a 32b 1+ROH / p... >p... I NOR “Ru" >P/I \OR P...— \ 32c Scheme 1.6 33 Using in situ IR and NMR spectroscopy, Epstein et al. studied the proton transfer in the dihydrogen-bonded complexes between (triphos)Re(CO)2H (33) and phenol, tetrafluoroboric acid (HBF4-0Me2), chloroacetic acid (ClCH2CO2H), hexafluoro-2- propanol (HFIP) or perfluoro-2-methyl—2-propanol (PFTB), as proton donors, at 200-260 K (Scheme 1.7).50 The 172-H2 complexes 33a were found again to be thermodynamically more stable than their HmH bonded precursors. Higher temperatures induced H2 loss with the formation of the covalent products 33b. Scheme 1.7 In the case of H2Re(CO)(NO)(PMe3)2 + CF3COOH, the dihydrogen-bonded complex coexists in equilibrium with the corresponding nz-H2 complex, which loses H2 upon heating.5 l Similarly, the dihydrogen-bonded adduct 34 (CF3)2CHOH---HW(CO)2(NO)(PMe3)2 was found to undergo proton transfer in the rate- limiting step with the formation of an unstable nz-H2 complex, which subsequently eliminates H2 to form a W-OR covalent product. The proton transfer in the dihydrogen-bonded complex (CF3)3COH---HRu(Cp)(CO)(PCy3) was also studied by Epstein et al. using in situ IR spectroscopy, and a relatively high barrier of 15 kcal/mol was found for this process.5 I To model this reaction, Scheiner et al. used the HRu(Cp)(CO)(PH3) ruthenium hydride model, which was allowed to interact with H3O+, CF3OH or H2O, representing strong, moderate, and weak proton donors, respectively.52 While in the first case spontaneous transfer of proton with the formation of a corresponding hydrated 02-H2 complex was observed, the other two weaker acids did not transfer the proton at all, suggesting that the activation barrier for this process is largely determined by the proton donor ability of the acidic partner. The critical role of the proton donor acidity has also been recognized recently by Lau et al., who concluded that strongly acidic conditions give nz-H2 complexes, while weakly acidic conditions favor dihydrogen-bonded species.53 In the ruthenium polyhydride complex 34, Chaudret et al. noted a substantial increase of the H-H coupling Jab upon formation of dihydrogen bonding with various proton donors in toluene, which was tentatively explained by the decrease of the electron density on Ru, caused by the partial charge transfer from the metal hydride to the hydrogen bond donor.54 When the CDC12F/CDF3 (2:1) solvent system was used instead, proton transfer with the formation of the (Cp*)(PCy3)RuH4 complex was observed at low temperatures, as a result of the unusual property of the Freon mixture to strongly increase its dielectric constant upon cooling, assisting thus the protonation of 37 .5 5 Interestingly, 35 the reactants can be mixed at room temperature, but the proton transfer only occurs when the temperature is sufficiently lowered to induce an adequate increase in the dielectric constant of the solvent. The same behavior was observed in CD2Cl2, a solvent which is also known to have a strong temperature-dependent dielectric constant. CP\ ..Hb Hui—Ha 34 According to Lau et al., intramolecular N-H~--H-Ru dihydrogen bonds also appear to mediate proton transfer and subsequent formation of N-Ru bonds in 35, as illustrated in Scheme 1.8.56 HID exchange of both protonic and hydridic hydrogen atoms with D20 strongly suggest the existence of nz-H2 intermediate species in equilibrium with the HmH bonded complexes. H2 loss with the formation of a Ru-N bonded chelate structure is facile in 35b, and the reverse Ru-N bond hydrogenolysis can be done at 60 °C under 60 atm (Scheme 1.8). This system was found to catalyze the reduction of CO2 to formic acid, although with low yields.56 The heterolytic splitting of the H2 ligand is believed to be a crucial step in the proposed mechanism, which is depicted in Scheme 1.9. The only NMR detectable metal-containing species throughout the reaction is 35, suggesting that the insertion of CO2 into the Ru-H bond is the rate-determining step, as also supported by recent theoretical calculations on a similar system.57 36 \ \\'R till Phi l‘k ,N\ Me thP“ ;u\ H N ...Me Me PPh2 LPth Me 35 n = 2 (a) n = 3 (b) + H2 H ' 2 14- @4042»: th P“ L/U\ ”,Me LPth Me Scheme 1.8 n = 3 ©(CH2M 1+ ©(CH230 1 + H2 “I Ru ,.. ,. Phi / \N N'IMe Phi ;UI\+’H N\"M° Pth Me PPh2 M9 -HCOOH " ? (CH2)n ©(CH2)n thp‘v/Ru\o H’N\'"Me + C02 thP‘“ /U\H ---‘H NimMe Agh )‘H Me Lppth \Me 2 36 35 (n = 2 (a); 3 (b)) Scheme 1.9 37 While the proposed metal formate intermediate 36 could not be detected, the analogous dithioformate complex 37 was easily prepared from 35b and excess CS2 (Scheme 1.10), and its identity was unambiguously established by IR and NMR spectroscopies.56 9) (CH2\)3 032 ©(CH223 thP‘“';u\H____ ,Nf‘ 'Me thP“"/Ru \s H,N'+{"Me th 35b 37 Scheme 1.10 Hydrogen exchange in the structurally related complex 38 was studied by Chaudret et a1. using 1H N MR spectroscopy, and the activation energy for this process was determined to be around 11 kcal/mol.58 However, extensive DFT calculations suggested that the mechanism for the exchange does not involve any proton transfer within the N-H---H-Ru dihydrogen bond. ©(CH222 w°RU N3” Ph P‘ , 'Me 3 / \H-u-H \ Ph3P Me 38 A very interesting dihydrogen-bonded system with its hydrogen exchange dynamics has been recently described by Jalon et al.59 They reported a three-center py2H---H-Ru intramolecular interaction in 39, in which fast scrambling between the 38 hydridic and protonic hydrogen atoms occurs most probably via an "2_H2 complex intermediate (Scheme 1.11). An activation energy of about 13.6 kcal/mol was determined for this process, using variable temperature 1H NMR spectroscopy. Moreover, this system proved to be a very active catalyst for DVHZ exchange. Thus, when a solution of 39 in CD3OD was exposed to a dihydrogen atmosphere (1 atm) at room temperature, more than 90% of Hg was exchanged for D2 in about half an hour. Scheme 1.11 A similar exchange was reported by Morris et al. in the ruthenium polyhydride complex 29b-[K(l,lO-diaza-l8-crown-6)] (Scheme 1.12).” Upon exposure to D2 gas at l 39 atm and room temperature for 5 min, the intensities of the NH and RuH 1H NMR (THF- (13) signals were depleted by 100% and 90%, respectively. For comparison, only 13% decrease in the hydride resonance was observed after 10 days when the much less acidic l8-crown-6 ether was used for complexation, implying efficient activation of the M-H bonds by N-H---H-Ru dihydrogen bonding. The exchange is also significantly slower if the ruthenium hydride is replaced by the less basic analogous osmium hydride (29a). While the conjugate acid of 29b, the known RuH2(H2)2(iPr3)2 dihydrogen complex, could reasonably play the role of the intermediate in the exchange process depicted in Scheme 1.12, its involvement in this transformation was ruled out by control experiments. PiPr3 ( PiPra ( H ’ .H 1(wa DZ,THF,5min> D" I “D _ VN D H’| \H \LOD—H _H2 0.. \D k ) PiPr3 PiPr3 o Scheme 1.12 Intramolecular N-H---H-Re interactions can affect hydride fluxionality in ReH5(PPh3)2L (L = N-acetyl-2-aminopyridine) (40).60 Thus, the free energy of activation for the tumstile rotation involving the H1, H4, and H5 atoms in this complex is 0.7 kcal/mol smaller than for the analogous complex with the NHAc group in the para position of the pyridine ring. Stabilization of the transition state by strong N-HmH-Re dihydrogen bonding, which is only possible in the ortho-NHAC isomer, seems to be 40 responsible for the observed difference. However, this effect is partly offset by non- bonding repulsive interactions between the two HmH hydrogen atoms, which are forced to approach to 1.49 A in the transition state, according to theoretical calculations.61 40 Dihydrogen bonding can have important consequences on the selectivity and stereochemistry of reactions in solution. Thus, the formation of directional and strong HmH interactions can differentially stabilize one particular transition state among two or more possibilities, ultimately controlling the product distribution or stereochemical outcome. An illustration of this concept is the selective imination of the nl-aldehyde complex 41 with ortho- vs. para—aminophenol, carried out by Crabtree et al.62 In a competitive experiment using an equimolecular mixture of 2-aminophenol, 4-aminophenol and 41, a 4.2:1 ratio of the resulting products 41a and 41b was obtained (Scheme 1.13), which was calculated to correspond to a k5/k6 value of 6. This outcome appears to be the result of O-H---H-Ir dihydrogen bonding stabilization of 413 (and presumably of the TS leading to it). The two isomers do not interconvert, indicating that the observed product distribution is dictated by kinetic not thermodynamic control. While the observed ratio between the two rate constants is translated into a AAGat value of 1.1 41 kcal/mol, the 4.5 kcal/mol estimated HmH interaction energy in 41a is substantially larger, the difference being apparently offset by the unfavorable chelate ring conformation required for efficient H-bonding in the transition state. Scheme 1.13 Dihydrogen bonding can also direct the regiochemistry of ligand attachment to transition metal clusters, as demonstrated by Aime and coworkers.‘53 Thus, reaction of the electronically unsaturated osmium cluster 42 with EtNHz or EtzNH yields exclusively the syn product 423, stabilized by an intramolecular N-H---H-Os interaction, which would not be possible in the anti isomer (Scheme 1.14). Notably, with Et3N, which lacks the acidic hydrogens required for dihydrogen bond formation, no reaction occurs. 42 - - - - "' . 0' . '4 o o - ‘9 N-—-R H (00) (gszosmok L co (I) 40s(00)4\ 3 \H/OS(CO)3 —> ( )3 S\H/05(CO)3 42 423 L = EtNH 2, Et2NH Scheme 1.14 The syn isomer is also preponderantly formed when 42 reacts with NH3. Subsequent treatment with acetaldehyde or acetone in chloroform leads to the exclusive formation of the dihydrogen-bonded imino-derivatives 42b with a syn configuration (Scheme 1.15).“ 65 The intramolecular N-H-uH-Os interaction is disrupted in more polar, hydrogen bonding solvents such as methanol or acetone. In this case, the interconversion between the syn and anti isomers was demonstrated by variable temperature NMR spectrosc0py, which suggests that the observed regioselectivity is a result of thermodynamic, rather than kinetic control. 33““ ...... H\ /C\R WOO) 1. NH3 T” N 4 2. CH C(R)=O Os(CO)4 | (CO)3054 > 3 _ 4 \H/Os(CO)3 CHCI3 r (CO)aOs\H/—\-05(co)3 42 42b H = H, CH 3 Scheme 1.15 43 The ruthenium hydride 43, whose crystal structure shows an intramolecular N- H---H-Ru short contact, was found to efficiently catalyze the asymmetric hydrogenation of ketones to chiral alcohols.66 However, the specific contribution of the H--H interaction to the high enantioselectivity observed in these reductions was not analyzed. 'Irs Ph,, N - \ L R NI Ph :_. \ . H H Work in our group by Gatling established the ability of O—H---H-B dihydrogen bonds to direct the borohydride reduction of ketones to alcohols.67 Thus, reductions of 2- hydroxycyclobutanone or 2-hydroxycyclopentanone with tetrabutylammonium borohydride in the non hydrogen bonding solvents CH2C12, ClCHzCH2Cl, or o- dichlorobenzene are accelerated about 150 times relative to the reductions of the corresponding unsubstituted cycloalkanones, and yield almost exclusively trans diols after workup (Scheme 1.16). These effects are greatly reduced in the presence of competing hydrogen bonding alcohols or anions like F, Cl', or Br'. Capping of the OH with a trimethylsilyl group also shuts off both the stereodirection and the rate acceleration. AMl semiempirical calculations predicted a 3.5 kcal/mol preference for the hydride delivery from the OH substituted face of 2-hydroxycyclobutanone. This value is in reasonable agreement with the experimental findings, despite the crude level of theory and the absence of counterions or solvent in the model. 1H, /B\H O H HO H H (CH2)n§—O C 202 (CH2)n2>< \I OH n = 1 , 2 Scheme 1.16 In the solid-state, dihydrogen bonding can control the transformation of cyclotrigallazane (12) into nanocrystalline gallium nitride, according to Gladfelter and coworkers.68 Initial loss of H2 at 150 °C resulted in an amorphous GaN phase, which upon annealing at 600 °C led to the metastable crystalline cubic gallium nitride, as a 1:1 mixture with the thermodynamically favored hexagonal GaN. The crystallization in the cubic system appears to be dictated by the initial crystal packing in 12, consisting of N- H---H-Ga dihydrogen-bonded chains (Figure 1.10), which can be considered essentially “hydrogenated” cubic GaN. For comparison, decomposition of cyclotrigallazane in thin films obtained by vapor deposition, a process that presumably disrupts the dihydrogen- bonded network, yields exclusively hexagonal GaN. It is remarkable that despite the huge contraction of the unit cell accompanying the conversion of 12 into cubic GaN, the reaction still maintains partial topochemical character. 45 2. Organic Solid-State Reactions: Mechanisms and Concepts Organic solid-state reactions have been reported to occur since the beginning of organic chemistry. The 1828 Wohler’s famous transformation of ammonium cyanate into urea was proven to take place both in solution69 and solid-state.7O However, it was not until relatively recently that this discipline has developed as a mature field, with the invaluable assistance of physical methods like X-ray crystallography, optical and electronic microscopies, thermal analysis, and solid-state NMR spectroscopy. A fundamental difference between the reactivity in fluid and solid phases is that while the former is dominated by the electronic properties of molecules, reactivity in solids is a balance between packing and electronic effects. In the solid-state, the reactants are usually locked in a fixed orientation with relatively low mobility, and consequently the intrinsic reactivity of the molecules is often less important than their spatial arrangement relative to the neighboring molecules in the crystal. Reactions under such geometrical control can therefore be highly selective, leading sometimes to products and materials otherwise inaccessible.71 The first to recognize this powerful concept was G. M. J. Schmidt, who in the 1960’s articulated the “topochemical” principle, of lattice control over the course of solid-state reactions and stereochemistry of the products.72 Thus, in their seminal study of photochemical dimerizations of trans-cinnamic acid derivatives, Schmidt and coworkers found correlations between their orientations in the starting crystals, and their solid-state reactivity, as well as stereochemistry of the resulting dimers (Scheme 2.1). 46 X—Ph "P h—— X X—Ph X-Ph [I"3.8-4.2A —_” X—Ph .i-‘COR "0 OR Y'IYPC X = H, OH, CI, N02, CH3, OCH3, OC2H5, OP!“ R = OH, oer-:3, OCsHs, NH2, CSHS Scheme 2.1 47 The monomers may have three major types of crystal packings. In the or arrangement, the molecules are anti-parallel, with 36-42 A between the double bonds, and yield almost exclusively centrosimmetric dimers. In contrast, dimers with a plane of symmetry are obtained by the irradiation of the B-type crystals, in which the double bonds are oriented in a parallel fashion, approximately 38-42 A apart. Finally, the y—type polymorphs, with distances between the closest double bonds greater than 4.8 A, are photochemically inert in the solid-state. This study made Schmidt soon realize the imperative of understanding the rules governing the packing of molecules in crystals, which led to the emergence of crystal engineering.72 According to Schmidt, for a reaction to be topochemical, it has to proceed with a minimum of atomic and molecular movement; otherwise, disorganization of the crystalline medium due to excessive molecular motions may lead to loss of lattice control over the reaction. Later, Cohen introduced a new concept, that of the “reaction cavity”, defined as the space occupied by the atoms directly involved in the solid-state reaction.73 The topochemical principle can be thus reformulated to state that those reactions which proceed under lattice control occur with minimal distortion of the reaction cavity, and that formation or removal of any empty space within the cavity should be energetically unfavorable, since they imply substantial changes in attractive and repulsive forces. Gavezzotti quantified this theory by proposing precise methods for molecular volume calculation, and concluded that a prerequisite for crystal reactivity is the availability of free space around the reaction site.74 Topochemical control was also noticed in the solid state decomposition of diacyl peroxides which photolyze to a pair of C02 molecules and a pair of free radicals (Scheme 48 2.2).75 While in solution or melt the fate of the radicals is complex and variable and mixtures of products are obtained, photolysis of single crystals give radical-radical coupling products in high yields. O—O C02 __—_—’ . Rflo chi? R—/ (:02 \—R Scheme 2.2 At the first inspection one might conclude that the topochemical least-motion principle is obeyed since radical coupling predominates in the solid state over disproportionation, which requires that one radical abstract 3 hydrogen atom from the more distant second carbon in the other radical. However, low temperature EPR studies of the generated radical pair suggest that the reaction does not follow a least-motion path, as the reactive carbons initially move apart before they finally combine. This behavior appears to be caused by the two C02 molecules resulting from photolysis, which cannot be accommodated in the reaction cavity. The pressure created by the forming C02 was estimated around 20 kbar from its asymmetric stretching mode, which is sensitive to pressure. The conclusion is that in such gas evolving systems, where there is no possibility that the product molecules would fit into the reaction cavity, the developed 49 stress could control the course of the reaction. Recently, a unified theoretical formulation of solid-state reactivity, that can be applied to thermal-, light-, and shock-induced reactions as well has been proposed, in which the reaction cavity and steric compression concepts have been quantified.76 There are two main mechanisms that may operate in organic solid-state transformations.77 The first one, the homogeneous mechanism, involves conversion of "A" molecules of the starting material in the initial crystal lattice, to "B" molecules of the product, with the B molecules remaining in the same region of the lattice initially occupied by the A molecules. A single crystal of the reactant will thus pass through a continuous series of solid solutions, first of B in A and later of A in B, before becoming a single crystal of the product. Minimal structural change is required on going from A to B so that the crystal lattices of reactant and product, which should have very close geometrical parameters, can accommodate each other in the process of conversion. These severe requirements make candidates for the homogeneous mechanism extremely rare. In most of the solid-state reactions the product is structurally incompatible with the reactant lattice, and therefore a heterogeneous mechanism operates, with initiation of reaction at nucleation sites frequently located at crystal imperfections or surfaces, and subsequent growth of the product phase, which usually does not retain the three-dimensional periodicity of the initial crystal, and appears amorphous or polycrystalline. Even under these conditions, the process can be topochemical; a great number of solid-state reactions have shown that there is no strong correlation between the chemical specificity and the long-range crystallographic order of the product phase.78 50 Another term frequently used by solid-state chemists is “topotaxy”, which is concerned with the relationship between the three-dimensional crystallographic orientation of reactants and products.79 In a “topotactic” transformation, the lattice parameters, crystal system, and space group of the product are closely related to those of the reactant. The solid-state process does not necessarily have to be homogeneous, as nucleation of the crystalline product at the surface of the initial crystal can lead to oriented growth under the influence 6f the surface tension, resulting in coincidence between certain crystallographic axes of reactant and product phases. It thus follows that while topochemical control often leads to topotaxy, the later phenomenon may be completely independent from the occurrence of a topochemical process. A potential problem in solid-state organic chemistry is that the intermblecular forces in organic crystals are predominantly of van der Waals or hydrogen bonding type, which are significantly weaker than typical covalent bonds. As a consequence, melting or sublimation may occur very often in thermally activated reactions, before any detectable chemical transformation. Another possible complication is the formation of liquid products. It is therefore imperative that these possibilities be ruled out any time the exertion of topochemical control is invoked. 2.1. Topochemical Polymerizations and Polycondensations In ordinary solid-state polymerizations there is essentially no lattice control over the polymer chain growth, apparently due to the considerable movement that the monomers must undergo to add to the end of the growing chain.80 This is the result of the 51 significant shrinkage that occurs when going from van der Waals distances between molecules to much shorter covalent bonds, leading to a growing gap between the chain end and the next reacting monomer. This means that ultimately the monomer can add to the polymer only by breaking away from the lattice, resulting in the loss of stereoregularity and crystallinity. Hirshfeld and Schmidt argued that this effect could be avoided in crystals of carefully designed bifunctional monomers.81 It was predicted that in suitable structures, the monomer would attach itself to its neighbor via a rotation about its center of mass, so that the polymerization would not require diffusive motion, and lattice control would be maintained throughout the transformation. Realization of this model has indeed been achieved in divinyl arenes82 (Scheme 2.3) and diacetylenes79 (Scheme 2.4), which can photopolymerize topochemically, leading to highly stereoregular, sometimes crystalline polymers. Scheme 2.3 52 C$C \C \C\\ \\\ \ C\ x R / C— R / FL R— C CS ‘0 C\c\ *’ \\C\ \ R / C— R / R\C R—C\ §\C\ C \ C\\ “ \C\R ‘c R // Scheme 2.4 These topochemical polymerizations may proceed either homogeneously or heterogeneously.79 In homogeneous cases the polymer is formed as a solid solution, which grows from points randomly distributed throughout the parent crystal. The product is thus isomorphous with the monomer crystal. This mechanism usually operates in the polymerization of diacetylenes, and in some cases, singleocrystal to single-crystal transformations can be achieved. In the case of [2+2] photopolymerizations, heterogeneous mechanisms dominate; the reaction starts preferentially at defect sites of the monomer crystal, with nucleation of the polymer phase. The parent crystal typically shatters into a polycrystalline aggregate, or, sometimes, an amorphous phase. Once the crystal is broken, polymerization continues into the crystallites, usually starting from the edges. Comparison of the monomer and polymer crystals showed that in many cases the space group does not change and the modification of the unit cell volume during 53 polymerization is very small, implying topotactic control. Moreover, when the initial diolefin crystallized in a chiral space group, an absolute asymmetric synthesis with the formation of an optically active polymer could be achieved.83 An interesting solid-state thermal polycondensation occurs in salts of halogenoacetic acids, leading to polyglycolide, the simplest possible polyester (Scheme 2.5).84 x—CH2-000'*M ———>1/nE——CHz-eoo-—]In + MX x = Cl, Br, I; M = Na, K, Rb, NH4, Ag Scheme 2.5 Interestingly, the metal halide byproducts are deposited as small cubic crystallites in the polyglycolide matrix, implying long-range diffusion of the formed M+X' ion pairs in the crystals. The salts can be subsequently washed out with water, to leave a highly microporous polymer. Crystal structure analysis in the case of silver chloroacetate indicated that this polycondensation is topochemical, and is driven by the very short Ag- Cl interatomic distance of 2.903 A, which is only 0.128 A longer than in silver chloride.85 The observed interatomic distance between the reacting carbon and oxygen atoms of 3.25 A is also within the range expected for high solid-state reactivity. 54 2.2. Hydrogen Bonding Assisted Topochemical Reactions The topochemical principle is a very powerful concept that can be applied to the assembly of both small molecules and extended covalent networks with controlled topologies. However, the major disadvantage of this approach is the lack of solid-state reactivity in many systems due to misalignment of the reactive functionalities in their crystals. This obstacle can, however, be eliminated by successful application of crystal engineering in designing solid-state reactive molecular solids. Schmidt and Leiserowitz noted as early as 1969 that the probability of mB-unsaturated amides to crystallize in photodimerizable structures was very high, due to favorable hydrogen-bonded, predictable packings.86 Alternatively, they could control crystal packing by introduction of stackable dichlorophenyl units, or by co—crystallization with mercuric chloride, which were shown to be capable of imposing 4 A periodic arrangements.72 More recently, Feldman and coworkers used the predictable hydrogen bonding dimer motif of the carboxylic acid functionality to properly align olefins for photodimerization (Scheme 2.6).87 This example is a nice illustration of using hydrogen bonds as preorganizing interactions, which can be subsequently reinforced by robust covalent linkages. While in this case a discrete dimer was assembled, extension of this strategy to higher-dimensional systems by appropriate design can be easily imagined.88 A similar approach was used by Matsumoto et al. for the topochemical photopolymerization of benzylammonium salts of butadiene-1,4-dicarboxylic acid (Scheme 2.7).89 55 x” 'o c 1 002-)? 2 —-\ - —/ 002- X+ — _ / n cogx” NH3 NH3 Cl C C Scheme 2.7 56 Hydrogen bonding between the N H3+ and COO' groups enforce the required orientation of the butadiene units for solid-state polymerization in this case (Scheme 2.8). NH3+ o - — O\c/—\:/C\0' / \ + H *HaN \ NH3 F? O -O\/(/:/:\.:/C\O' 0 +H N NH3+ 9 3 O -O\/(’:/_\\:/C‘O' // \ O +H3N \ Scheme 2.8 The reliable self-assembly of urea derivatives into hydrogen-bonded a—networks was exploited by Lauher and Fowler to preorganize diacetylenes for solid-state topochemical photopolymerization.90 Thus, by co-crystallization of a pyridyl-substituted diacetylene with dicarboxylurea, an anticipated B-network could be assembled, in which the diacetylene units were forcibly aligned into the desired solid-state reactive arrangement (Scheme 2.9). Using a similar strategy, they were recently able to align a triacetylene for solid-state polymerization, which led to an unprecedented 1,6-polytriacetylene via a single-crystal to single-crystal reaction (Scheme 2.10).91 57 O H H O \ \ é , N O H H 0 o O H I'I IRA H o o ' O \0 \I/ 0’ , . o l H N N H / O 0 I I , \ . o H H o N / \ " ‘ t‘ I} \ \ O/ , N O H 1‘". O H’0 N/U\N O‘H O I | I ~‘N ’ // O \ \ O H’ '7‘ l‘ I N // VN. , 'fi fi' 0 o H H o ’ ,0 NA 0. /N H \H/\‘I 1:1/Y H~.\N/ // H ? E '7' O H o H H 0 SY0 ‘0 Y o/ o 0 Scheme 2.9 58 com--- N - / H’xt QN \ \\ \ ¢N H/\©N H020 // © (H) C: 7 H020 l hv /~COQH- H OVI // m o // I") ré ICOZH AJVN/ Ho2 c] K H 0' H020 Scheme 2.10 59 Besides hydrogen bonding, other noncovalent associations such as phenyl- perfluorophenyl stacking interactions have also been explored as preorganizing elements in the solid-state topochemical polymerization of diacetylenes92 and diolefins.93 2.3. Kinetics of Solid-State Decompositions In this section the focus will be mainly on organic solids which decompose to yield solid and gaseous products?4 as also expected from dihydrogen-bonded systems upon heating in the solid-state. This can be represented as: Asolid '9 Bsolid + Cgas Decomposition may be initiated by heat, light, or ionizing radiations. The stability toward thermal decomposition may be affected by the previous history of the sample, presence of impurities, aging, sample size, or the nature and pressure of gas atmosphere above the solid.95 The rate of the process can be most conveniently followed by measuring the pressure developed in a constant volume system or by using a thermobalance to determine the weight loss under isothermal conditions, although XRD or spectroscopic methods can also be used.94b The fractional decomposition 0t vs. time curves are typically sigmoid in shape, indicating an autocatalytic reaction. This pattern is almost universally associated with the initiation of reaction at specific sites followed by the growth of nuclei, with the reaction mostly confined to the product/reactant interface. The kinetics is therefore controlled by the number of nuclei present and the total area of the expanding interface. After the inflection point, the growing nuclei start to coalesce, slowing the decomposition rate as the interfacial area decreases, until the reaction eventually st0ps. 6O Different kinetic models with their corresponding equations have been elaborated to account for various solid-state decomposition mechanisms.94 Among them, the most common are the Avrami-Erofeev equation (1) and the phase boundary model (2): [-ln(1-0t)]""=kt; it: 1-4 (1) 1-(1-ot)“"=kt; n= 13 (2) The first one corresponds to a nucleation and growth mechanism, while the second is associated with an inward advancement of the reaction interface from the crystal’s edges. The exponent n in Eq. 1 is given by B + X, where B is the number of steps involved in nucleus formation (typically B = O or 1, the former corresponding to spontaneous nucleation), and 2. is the dimensionality of the nuclei growth.94b For some decompositions the acceleratory region can be described by an exponential expression (3).94a (X = Ce,“ (3) Due to molecular volume change, the formation of product molecules induces strain in the crystal, which produces cracks forming fresh surfaces on which decomposition can occur. The reaction will therefore spread down these crevices into the crystal. Decomposition on these surfaces produces additional strains and hence further cracking and a type of chain branching process thus develops.94c This mechanism can be mathematically described using the Prout-Tompkins equation (4): ln [oz/(1-oc)] = kt (4) It should be pointed out, however, that very often there is no single model that can acceptably describe the whole conversion range, as different mechanisms may operate for different stages of decomposition. 61 A substantial autocatalytic effect has been also observed for some single-phase solid-state polymerizations, which could be explained quantitatively by a strain- dependent rate of chain initiation and propagation model.96 Thus, it was demonstrated that the conversion—time curves display a more pronounced sigmoid character as the differences in unit cell parameters between the monomers and polymers increase. The rate constants obtained from solid-state reactions usually obey the Arrhenius equation. However, special care must be exercised when interpreting the obtained activation parameters, since decompositions of solids are complex processes involving not only chemical steps such as breaking or formation of bonds, but also physical transformations such as phase transitions, diffusion and desorption of gaseous products, and heat transfer. Therefore, chemically specific techniques may be necessary, to obtain mechanistic details at the molecular level. Another important issue is whether the Arrhenius equation has any physical meaning in the solid-state, as its theoretical foundation has been established in the context of collision theory in the gas phase. However, even if the Maxwell-Boltzmann energy distribution is not applicable to the immobilized constituents of a solid, it was demonstrated that energy-distribution functions of similar form (Fermi-Dirac or Bose-Einstein) also give rise to an Arrhenius— type equation.97 Thus, its application to solid-state transformations is justifiable not only in terms of a useful empirical parametrization, but is also supported by a rigorous theoretical foundation. An alternative approach for the kinetic analysis of solid-state reactions is the isoconversional method, applied under isothermal or nonisothermal conditions.98 This strategy allows the estimation of the activation energy without assuming a particular 62 reaction model, and is particularly convenient for analysis of nonisothermal data such as that obtained from DSC and TGA experiments. Under isothermal regime, the activation energy at a particular conversion, Ea, can be evaluated using Equation 5:98b - In to“ = ln[A/g(0t)] - Eel/RT, (5) While in some cases it can unmask the complexity of the solid-state process studied, the major disadvantage of this method, however, is that the obtained activation energy varies significantly with the extent of reaction, which complicates the interpretation of the kinetic data. 63 3. Topochemical Assembly of Covalent Materials Using Dihydrogen Bonding As illustrated in Chapter 1, many studies regarding the dynamics of dihydrogen- bonded systems in solution have demonstrated that proton transfer from the acidic AH partners to the transition metal hydrides MH, along the H---H bonds, generally leads to 172-H2 non-classical complexes, which subsequently eliminate hydrogen upon heating, with the formation of covalent M-A bonds (Scheme 1.2). An analogous process appears to occur in the case of the borohydride anion. In aqueous solutions, BH4' is very likely dihydrogen-bonded to H20, as suggested by the crystal structure of NaBHn-ZHZO,ll as well as theoretical and experimental studies by Epstein et al.10 Under neutral or acidic conditions, borohydrides undergo hydrolysis to boric acid (B(OH)3), for which the established mechanism involves slow proton transfer resulting in a BH5 intermediate, followed by fast H2 loss and B-0 bond formation (Scheme 31).” Activation parameters AH“ and AS“ of 20.6 i 1 kcal/mol and —22.3 i 3 en were measured for the neutral hydrolysis, while under acidic conditions the corresponding obtained values were 8.0 i 1 kcal/mol and —3 _+.. 3 eu, respectively.100 The structure of BH5, as deduced by theoretical calculations, is best described as an almost planar BH3 molecule, loosely coordinated by 10] H2. H - - BH4i-—Hzo s'°‘” Has—H OH fi+ HaB-OH + H2 Scheme 3.1 Theoretical work by Elguero et al. indicates that H2 generation from dihydrogen-bonded borohydrides can also be induced by the internal forces within a crystal.‘02 All these premises, together with the established ability of borohydrides to self-assemble into extended dihydrogen-bonded networks, suggest that A-Hn-H-B dihydrogen bonds could be employed in topochemical assembly of covalent materials. Such weak HmH interactions, in principle, may be used to organize and hold a structure’s form while it is more firmly fastened together by A-B bond formation, transferring thus the initial order from the starting crystal to the newly formed covalent frame. This strategy makes dihydrogen bonding a potentially powerful tool for rational assembly of new covalent crystalline materials with controlled structures and properties. In the topochemical transformation outlined above, the initial H-bond arrangement should detemiine the final covalent structure. However, when going from the A-H---H-B interaction to a covalent A-B bond there is approximately a 1.5-2 A change in distance between A and B. This is a considerable shrinkage, and if it were cumulative in the reaction direction, it would lead to an increasing gap between the growing covalent network and the unconverted dihydrogen-bonded molecular units as the decomposition advanced, with eventual loss of lattice control over the reaction. Even under these circumstances, however, the transformation may still be topochemical, as demonstrated by the solid-state conversion of cyclotrigallazane into nanocrystalline gallium nitride reported by Gladfelter et al.“, which is accompanied by 62.9% shrinkage of the unit cell. The price to pay was the initial loss of crystallinity and the consequent requirement for high annealing temperatures to restore it. Although this thermal treatment 65 had no detrimental effect upon the robust GaN product, more delicate structures would not tolerate such high temperatures, limiting the general applicability of this approach. A low temperature procedure for topochemical dihydrogen to covalent bonding transformations would allow the extension of this strategy into the structurally more diverse domain of organic materials. Like many solid-state processes, this reaction includes two threats to the crystalline order: (a) geometry change upon bond reorganization, and (b) gas generation and diffusion within the lattice. Clearly, careful design of the starting dihydrogen-bonded networks is necessary in order to meet these challenges. Success, however, would mean that the well-developed tools of molecular synthesis could now be applied to the rational construction of crystalline covalent solids with desired structures and functions. This work involves two strategies to address this problem: (a) design of cations to form closed loops in coordination with hydride-bearing anions, in which case the lattice distortion would not be cumulative, and (b) selection of globular cations large enough that their close packing determines the lattice parameters, with the hydridic anions fitting into the interstitial holes, in which bond formation via flexible arms would induce 103 A dividend of this latter strategy is that minimal change in the unit cell (Scheme 3.2). the anticipated looseness of the lattice should allow the released H2 to diffuse readily through and out of the crystal. 66 (a) A y. I I M —m Scheme 3.2 3.1. Results and Discussion N-[2-(6-Aminopyridyl)]-Acetamidine (NAPA) Cyanoborohydride: A Closed Loop Dihydrogen-Bonded Structure by Design.'03 NAPA H3BCN (44) was synthesized (Scheme 3.3) to explore the prospect of convergent coordination in closed loops, which appeared as a likely possibility due to the bent geometry of the aminopyridyl-acetamidine unit. The cyanoborohydride anion was chosen as the hydridic partner because of its enhanced stability toward acidic substrates relative to BHa‘,IOIC which was decomposed instantaneously in solution by the NAPA cation. The white crystalline compound melts at 119-120 °C with decomposition and gas evolution. The 1H, '3 C, and 11B NMR, as well as IR spectra confirmed its structure. 67 ., / OCH / \ I + H3C_< + 3 - —_’ £1 —] H2 NH2 NH2 CI H2 N NH ct Wk CH3 NaNCBHa -NaC| CH30H NCBH3 H2N’J\ H3 Scheme 3.3 Figure 3.1 shows the X-ray crystal structure for 44. There are two independent centrosymmetric (NAPA H3BCN)2 dimers in the unit cell, each of them exhibiting close H-H contacts: 1.98, 2.12, 2.26 A, and 2.04, 2.09, 2.31 A, respectively. Applying the corrections for the N -H and B-H bonds that appear too short from X-ray compared to typical literature values, these H-H contacts become 01-015 A shorter. This is well below the 2.4 A van der Waals contact radius, implying strong and specific interactions. The NH---H-B angles are strongly bent, ranging between 91.6° and 126.3°, with an average of 108.2°, while the N-H-~HB angles are larger (range 148.1 - l75.6°; average 158.5°), as expected for dihydrogen bonds. It thus appears that despite the less negative charge on the hydridic hydrogens (-0. 18 vs. -0.27 by Mulliken population analysis for MP2/6-311++G** wavefunctions), the NCBH3' ion is capable of strong dihydrogen bonding, comparable to that of the BH4’ ion in this case, owing to the increased acidity of the proton donor partner. More important, the structure of this very first attempt showed two independent occurrences of the intended closed loop packing, with its potential for topochemical control. 68 N(5A)€:’:‘ i A A C(BA “sf- f .L, B(A)\3 Figure 3.1. X-Ray Crystal Structure of NAPA H3BCN showing the HmH contacts in A. 69 Decomposition in both solid and solution unfortunately led to complex mixtures (Scheme 3.4) via the unwanted reduction of the amidine group by the cyanoborohydride, which accounts for the majority of the decomposition products identified by MS, lH, '3 C and 1'B NMR. Although the decomposition product ratios were slightly different in solid state and solution, we could not attribute that result to topochemical control, since the solid state reaction occurred with partial liquefaction due to the low melting point of some of the products, relative to the temperature required for decomposition (~65 °C). / / H \N I NH solid-state, 65 deg N11 + 2 or THF refl. H2 \N NH2 NCBngb * CH3 44 / / / \ I . \ I \ I H2N NH—CHZCH3 H2 NH CN N N NH2 1 AN+IB\NJK l I H H NCBHg 443 Scheme 3.4 70 Despite the lack of observable topochemical control, this system demonstrated the ability of NCBH3' to form dihydrogen bonds and showed that the closed loop coordination is a viable packing arrangement that can be readily designed into dihydrogen-bonded systems. Additionally, another critical design feature for the next dihydrogen-bonded systems became apparent: their melting point, which should be sufficiently high to allow the decomposition to be carried out in the solid. Topochemical Control by Dihydrogen Bonding. Structure and Reactivity of NaBH4tTEA (45).103 The complex of NaBH4 with triethanolamine (TEA) was synthesized as a candidate for the globular cation strategy. I‘\ The complex precipitated from a stirred mixture of NaBH4 and TEA in THF as a white crystalline compound, which melts at 107-108 °C with decomposition and gas evolution. The X-ray powder diffraction pattern of the complex is unique and confirms the absence of NaBH4. The 11B and 23 Na solid state MAS NMR chemical shifts are -47.9 and -9.9 ppm respectively, 2.5 and 6.1 ppm downfield relative to sodium borohydride, indicating different environments for the BH.{ and Na+ ions. However, the B-H stretching vibration frequency for 45 (2288 cm") is almost identical with the corresponding value in N aBH4 (229lcm"). The X-ray crystal structure is shown in Figure 3.2. 71 (a) I' ‘t I. ' 7 1 I / l 1‘ / I»). A, _ five/'2' C(23)“ - Has)?" U ."I (b) Figure 3.2. X-Ray crystal structure of NaBHa-TEA: (a) coordination of Na+; (b) dihydrogen bonds connecting the chains, with H-H contact distances in A. 72 Each Na+ is hexacoordinated by the N and two 0 atoms from a TEA molecule and by three 0 atoms from the two neighboring TEA molecules, forming TEA-Na+ linear chains (Figure 3.2a). Two 0 atoms from each TEA molecule are shared by two adjacent Na+ cations, but the third oxygen only coordinates the next Na+ in the chain. Thus, 45 is unlike the previously reported structures of TEA complexes where each cation was complexed by all four heteroatoms from a given TEA molecule.104 Multiple dihydrogen bonds connect the chains via the BH4’ anions, giving rise to extended two-dimensional layers (Figure 3.2b). Each BHa' H-bonds with two OH groups in one chain and one OH group from the next chain in a total of 5 dihydrogen bonds. The H—H contacts from the three different OH sites are 1.94, 1.93 and 2.12 A, and 2.16 and 2.13 A in distance respectively, the latter two being bifurcated H-bonds. Again, typical corrections for O-H and B-H bond lengths lead to H-H distances which are 0.15-0.2 A shorter. The OHmH-B angles vary between 91.8 and 111.70 (average 98.0°) while the O-H-HHB angles range between 143.9 and 166.7° (average 157.0°). There is no conventional H-bonding; instead all the hydroxylic protons point to the interchain space to form dihydrogen bonds with the borohydrides, demonstrating the importance of these interactions in defining the crystal packing. However, an important driving force for the formation of 45 must be the complexation of Na” by TEA, since no complex was formed with (CH3)4NBH4 or KBHa. Solid-state decomposition of NaBH4-TEA at 82 °C under Ar or open atmosphere for 30-58 days resulted in a loss of 3 moles of H2 for each mole of 45, as indicated by the H“ content of the initial complex and final decomposed material. Attempts to use higher decomposition temperatures induced melting. The rate of decomposition seems to be affected by different factors such as sample size, humidity, or the nature or pressure of 73 gas atmosphere above the solid, which is characteristic for many solid-state reactions.95 Thus, decomposition is slower under vacuum than under Ar, and it is considerably faster in open atmosphere, probably due to the presence of humidity. The resulting white solid is insoluble in common organic solvents and it does not melt up to 300 oC, suggesting a polymeric structure. Its llB solid state MAS NMR shows a single peak at 8 = -7.1 ppm. X-Ray powder diffraction of the same material exhibits only a semi-amorphous phase with a broad peak at 20 = 94°, but no peaks corresponding to NaBH4 or to the initial complex (Figure 3.3c). Furthermore, the 1‘B NMR taken after hydrolysis of this material in neutral D20 showed only traces of BH4’, indicating the virtual absence of this boron species in the solid-state decomposition product. These data, correlated with the hydride content of the decomposed material (one H' left), suggest a trialkoxyborohydride structure for the decomposed material. The B-H stretching frequency is shifted 5 cm'1 relative to the initial complex (from 2288 to 2293 cm"'). This is not a dramatic change, but it is not surprising since the variation of VB“ in the BHX(OR)4-,[ series is generally 1’05 The loss of 3 moles of H2 is in agreement with the crystal structure of 45, where smal each BH4‘ is H-bonded to three -OH groups, two from one chain and one from the next chain. This fact suggests a topochemical relationship between the starting and final materials. Additionally, the two OH from the same chain in 45 belong to different adjacent TEA molecules. A topochemical reaction should therefore lead to a two- dimensional, extended covalent structure, and indeed, the peak at 29 = 9.4 in the powder XRD (Figure 3.3c) corresponds to a d spacing of 9.41 A in a layered structure. This value is very close to the 9.24 A inter-layer distance in the initial NaBHa-TEA. 74 l_ I I V ‘ ; I. I I. I -I- I b p : ‘EJ‘j I. "“ C hllLllllllnlllJlell O 1 O 20 30 40 50 60 20/° Figure 3.3. Progress of NaBHa-TEA decomposition monitored by X-ray powder diffraction: (a) initial complex; (b) high-temperature polymorph; (c) final decomposed material. 75 Monitoring the solid-state decomposition by X-ray powder diffraction (Figure 3.3) and H' analysis revealed a slow (~ 2 weeks), reversible phase transition of the initial complex, prior to any H2 loss. The IR spectrum for this high temperature polymorph shows the B-H stretching vibration at 2291 cm‘1 and the “13 solid state MAS NMR spectrum shows a chemical shift at -51.4 ppm. Structure elucidation of this high temperature crystalline intermediate, which somewhat complicates the topochemical relationship between the initial and final structures, would provide useful insight into the intimate mechanism of conversion from the dihydrogen bonding to the covalent networks. Unfortunately, only the low temperature polymorph of 45 has been obtained in recrystallization attempts to date. An indication for topochemical control is that different reaction products or stereoselectivity are observed in the solid state than in solution or melt. Therefore, we also studied decomposition of 45 in DMSO solution and in the melt. A solution of 45 in DMSO was stirred and heated under Ar at 110 °C for 55 hours. Like the solid-state decomposition product, the resulting precipitate is insoluble in common organic solvents and does not show any melting below 300 °C. However, H’ analysis and IR showed that this compound has virtually no hydridic hydrogen left. Its llB solid state MAS NMR chemical shift is slightly different from the solid-state decomposition material (6 = -5.0 ppm). No other product of decomposition was found by 11B NMR of the mixture resulting from a parallel reaction run in DMSO-d6, BH4' being the only boron species left in solution. On the other hand, when the solid-state decomposition product was stirred in DMSO under the same conditions, no change was observed in the hydridic content or IR spectrum, implying again different products for the solid and solution decompositions. 76 This experiment demonstrated that unlike in the solid state, where the reaction stops at trialkoxyborohydride, decomposition of 45 in solution resulted in complete alcoholysis of BHa' to B(OR)4' as expected for a reaction in which the BHX(OR)4-,{ intermediates, which are more reactive than the starting borohydride, are mobile and therefore susceptible to disproportionation. '06 Similarly, the reaction in the melt at 130 0C under Ar for 2 h resulted in complete alcoholysis of half of the BH4' from the initial complex, NaBH4 being the byproduct of decomposition as indicated by X-ray powder diffraction and HB NMR. It is remarkable that despite the considerably higher temperature of decomposition in the melt, only 2 moles of H2 were lost compared to 3 in solid state. The decomposition process in the solid state thus differs significantly from those in solution or melt. These results demonstrate that the particular packing of the molecules in the original crystal, dominated by H---H interactions, induced the reaction to take place under topochemical control, leading to an otherwise inaccessible poly-trialkoxyborohydride structure. However, this product showed poor crystallinity, and the reaction times were exasperatingly long due to the relatively low temperature required for decomposition in order to avoid melting. Influence of the Relative Acidity/Basicity of the Proton/Hydride Partners Upon Solid-State Reactivity of Dihydrogen-Bonded Systems: Structure and Reactivity of NaNCBH3~TEA (4t5).'03 NaNCBH3-TEA complex was synthesized in order to explore the influence of varying the hydridic donor partner upon the structure and solid—state reactivity of dihydrogen-bonded systems. Due to the smaller negative charge on the hydridic hydrogens, the dihydrogen bonds were expected to make smaller 77 contributions in defining the solid-state structure and reactivity, compared to the borohydride analog. The complex was crystallized by slow evaporation of a 1:1 triethanolamine-NaNCBH3 mixture in isopropanol. The X-ray structure is shown in Figure 3.4. The Na+ cations are heptacoordinated by the four heteroatoms from one TEA molecule, the nitrogen from the cyanoborohydride and two oxygens from neighboring TEAS which bridge adjacent N a” cations to form extended chains (Figure 3.4a), crosslinked by dihydrogen bonds. One H from each CNBH3' hydrogen bonds to a hydroxylic proton from another chain (Figure 3.4b) with an uncorrected H-H contact distance of 2.16 A (OH---H-B angle = lOO.6°; o-H---HB angle = 159.6°), forming a three- dimensional network. There are also O(1)-H---O(2) intra-chain conventional H-bonds connecting neighboring TEA molecules. It appears that dihydrogen bonds are less important in the NaNCBH3-TEA crystal packing than in the borohydride analog, with its shorter H---H contacts and complete absence of conventional H-bonding. This difference is also reflected in the significantly different solid-state reactivity of the two complexes. While the borohydride complex melted at 107-108 °C with decomposition, the cyanoborohydride analog melts at 85-86 °C, and it takes approximately 100 more degrees to finally start decomposing. Thus, close HmH contacts do not automatically confer solid-state reactivity; the relative acidity and basicity of the protonic and hydridic partners are also critical. Precise estimates of the basicities of BH4‘ and BH3CN' are unavailable, but the bimolecular rate constants for their H3O+-catalyzed hydrolysis are ~106 and 10'2 l/mol x sec respectively which may be translated into a pKB difference of 2 8 units.'07 78 (a) Figure 3.4. X-Ray crystal structure of NaNCBH3-TEA: (a) coordination of N a+; (b) dihydrogen bonds connecting the chains, with H-H contact distances in A. 79 Tuning Dihydrogen Bonds: Enhanced Solid-State Reactivity in LiBHa-TEA (47).108 The simplest way to tune the solid-state reactivity in 45 appeared to be the substitution of the Li+ cation for Na“, which should result in stronger complexation by TEA, thus making the OH sites more acidic, and consequently more reactive. The LiBHa- TEA complex (47) was obtained by slow evaporation of a 1:1 TEA-LiBH4 mixture in 2- propanol, as a white crystalline compound, which does not melt up to 300 °C. The X-ray crystal structure of 47 is presented in Figure 3.5. Each Li+ is pentacoordinated by the N and three 0 atoms from a TEA molecule and by an 0 atom from a different TEA molecule, forming (Li+TEA)2 dimers, in contrast to the Na analogue, where (Na‘TEA)n chains were present. Multiple dihydrogen bonds connect the dimers via BH4' pairs, giving rise to extended ribbons. The dihydrogen bonding network is asymmetrical: one BHa' H- bonds with three OH groups (0(2’), 0(1) and 0(3A)) in a total of 6 dihydrogen bonds, while the second BH4' only H-bonds with two of the remaining OH groups (0(2A’) and 0(3)) in a total of 4 dihydrogen bonds. The 0(1A)H groups are not involved in any dihydrogen bonding; instead they form conventional H-bonds with 0(l)s from the neighboring ribbon, creating thus overall two-dimensional layers. The H-H contact distances range between 1.69 and 2.32 A. The typical normalization of 0-H and B-H bonds to 0.96 and 1.21 A respectively leads to the more realistic H-H distances of 1.62 - 2.28 A. All hydrogen atoms were unambiguously located from the difference Fourier map, making these distances sufficiently reliable. The 0H---H-B angles range between 75.8 and 106.1° (average 95.0°), while the 0-H---HB angles are larger, ranging between 130.0 and 170.7° (average 150.8°). 80 2.30" IONA) \_. Figure 3.5. X-Ray crystal structure of LiBHa-TEA showing the H---H close contacts in A. 81 The shortest dihydrogen bonds are 1.69 and 1.76 A in distance, which after the normalization of the 0-H and B-H bonds become 1.62 and 1.67 A respectively. These are the shortest H-H distances reported so far for dihydrogen bonds. Notably, both OH groups involved in these two dihydrogen bonds contain bridging 0 atoms coordinating two Li+, which undoubtedly results in increased acidity of their corresponding protons, and thus enhanced hydrogen bonding ability. All the other OH groups complex only one Li+, leading thus to H-H separations that are significantly longer. The very short H-H distances and the expected acidity increase of the OH groups due to their complexation of Li“ predicted enhanced solid-state reactivity for this system. Indeed, when heated approximately 1 h at 120 °C under Ar, the complex completely decomposed, as indicated by X-ray powder diffraction (Figure 3.6), HB solid-state MAS NMR and IR spectra of the resulted material. Only about 24 h were necessary for decomposition at 82 °C, the temperature used for the solid-state decomposition of NaBHa -TEA, which took about 6 weeks. Three moles of H2 per mole of 47 were lost, as indicated by the H' content of the initial complex and final decomposed material. The resulting solid is insoluble in common organic solvents and does not melt up to 300 °C, indicating a polymeric nature. Its llB solid state MAS NMR exhibits a single peak at 8 -4.6 ppm, which, together with the H' content, suggests a trialkoxyborohydride structure for the decomposed material. The v3-“ of 2247 cm'1 for the same material is also significantly shifted from the 2290 cm‘] in the initial complex. No phase transition was observed in this case when the decomposition was monitored by X-ray powder diffraction. 82 lljl4lllllllllll_lJ_lllllll 0 1 0 20 30 4O 50 60 29l° Figure 3.6. X-Ray powder diffraction pattern of LiBH4-TEA (a) and its solid-state decomposition product (b). The loss of three moles of H2 points to a topochemical relationship between the final decomposed material and the initial complex, where two BH.{ and six OH groups are in close proximity due to the crystal packing imposed by the dihydrogen bonding network. Unlike in the NaBH4-TEA case where two-dimensional covalent layers were likely to have formed, a topochemical reaction in the present example is expected to yield a one-dimensional covalent structure (48), according to the crystal structure of 47 (Scheme 3.5). However, due to the modest crystallinity of the final material (Figure 3.6b), a definitive knowledge of its structure remains elusive. Using lower temperatures for decomposition (down to 65 °C) did not improve the crystallinity of this material. 83 E LI-. / . I 2) x-,\' ‘~ s o ‘ Q 9 2H0 t E W ”’4 N~~-L- ,IJ- —~.,, It Ir H2 Q a H\ -/H HM. B \\H O\ X / l 1 BH BH H ,,,,,,, H r (I) ,,,,, “0‘0 0 ‘ um. :' <5 9’» N“-l'.i a“ ' "I ‘1 E. ’ E/fll \\ 3(2' ‘ €;Zl.-i‘~~:j|'z'{ N. " d:o\ /OJ 0.’ ‘9" 3,? BH‘BH H131 - I I 48 H g : ’Hlu. .MH I H 47 Scheme 3.5 In direct contrast, decomposition of 47 in DMSO at 120 °C under Ar for 4 days yielded a material with no hydridic hydrogen left as shown by its H' content and IR spectrum. Its llB solid state MAS NMR chemical shift (-5.9 ppm) is slightly different from the corresponding value in the solid-state decomposition product. In a parallel reaction run in DMSO-d6, unreacted BHa' and completely alcoholyzed B(0R)4‘ were the only boron species found by HB N MR in solution throughout the reaction. Such behavior is expected for a reaction in which the BHX(0R)4-,( intermediates are more reactive than the starting BH4' and susceptible to disproportionation.106 84 It is important to rule out such disproportionation in the solid-state. As depicted in Scheme 3.5, pairs of trialkoxyborohydrides are presumably present in 48. Disproportionation into dialkoxyborohydn'de and borate is therefore conceivable in principle. For instance, in LiBH(0CH3)3, slow disproportionation into LiBH2(0CH3)2 and LiB(0CH3)4 was observed in THF.105 a’ 109 Table 3.1 presents the experimental llB NMR chemical shifts for the BHX(0CH3)4.,{ (x = 0-4) anions, for comparison with the experimental data found for 48. Unfortunately, no N MR data in the literature referring to BH2(0CH3)2' were found. Therefore the 11B NMR chemical shifts for the same series were calculated at the RHF/6-3 10* level (Table 3.1), finding good agreement between the experimental103 ’ 108’ ”0 and calculated values for the known BHX(0CH3)4-,( anions. Table 3.1. NMR data for BHX(0CH3)4-X', 47, and 48. llB NMR 8“ (ppm) 7Li N MR 8" (ppm) calculated experimental BH4' -507 49.7 (solid)c 1.5 (solid Linn.)d 49.5 (solid)d BH3(0CH3)' -26.4 -25.1 (THF)f - BH2(0CH3)2' -12.8 - - BH(0CH3)3' -10.8 -8.6 (THF)f - B(OCH3)4' -l4.9 -152 (CH30H)° - -13.0 (THF)f 47 - 48.0 (solid)d 2.8 (solid)d 48 initial - -39 (solid) 3.1 (solid) annealed - -2.6 (solid) 1.0 (solid) a relative to B(0CH3)3. b relative to LiCl. ° reference 103. d reference 108. ° reference 1103. freference llOb. Compared to these values, however, our decomposed borohydrides exhibit 8 values that are consistently shifted downfield. For example, decomposition of 47 or of its Na 85 analogue in DMSO yielded polymeric borates which showed llB solid state MAS NMR peaks at -5.2 and -4.3 ppm," respectively, significantly different from the corresponding values in B(0CH3)4'. However, for the solid state decomposition product 48 and its Na analogue, 8 values of -3.9 and -6.4 ppm,‘ respectively, were observed, in better agreement with the corresponding experimental value of -8.6 ppm in BH(0CH3)3'. The only possibilities that can be ruled out unambiguously based on the available data are the presence of BH3(0R)' or 8114' in 48, which would display distinctive chemical shifts around -26 or -50 ppm, respectively. The chemical shifts for the other alkoxyborohydride species are too close to each other for a safe conclusion to be drawn. It would be very difficult to distinguish, especially in the solid state, between a BH(0R)3' structure and the 1:1 mixture of BH2(0R)2' and B(0R)4' that would result from disproportionation. An experiment that would possibly differentiate between the two possibilities is to decompose a sample of 47 that contains 50% deuterated borohydride. If any disproportionation occurred during decomposition, mixed BHD(0R)2' species would result, which should display IR bands significantly different from those of BH2(0R)2' or BD2(0R)2'. Ab initio calculations at the RHF/6-3IG* level predict the BH and BD stretching frequencies in BHD(0R)2' to be shifted by -10 and -53 cm], respectively, relative to the corresponding values in dialkoxyborohydride or -borodeuteride, respectively. The reference LiBDaoTEA (47a) was first synthesized. The solid-state decomposition product resulting from this material (48a) exhibits the V31) at 48 cm'| lower than the starting compound 47a (Table 3.2). This change is comparable with the corresponding shift of -43 cm’1 accompanying the transformation of 47 to 48. The 50% ' Corrected by +0.7 ppm for adjustment to the B(0CH3)3 reference.1 '0“ 86 deuterated complex (47b) was then obtained, starting from a 1:1 mixture of LiBH4 and LiBD4 in THF, by precipitation with TEA. It exhibits VB" and Van values that are very similar to the corresponding values in 47 and 47a (Table 3.2). The IR spectrum of its solid-state decomposition product (48b) is virtually identical with the spectrum obtained by addition of the spectra corresponding to 48 and 483 (Table 3.2). This result suggests that no disproportionation occurs during the solid-state decomposition of 47, as a result of the isolation and reduced mobility of the borohydride units in 47 and 48. Table 3.2. IR data for LiBH4- TBA (47), LiBDaoTEA (47a), LiBH.t(50%D)-TEA (47b), and their corresponding solid state decomposition products, 48, 48a, and 48b. LiBl-I4 is included for comparison. -1 IR VBH (BD) (cm ) 47 2231, 2290, 2369 48 initial 2171, 2247, 2301 annealed 2340, 2368 473 1652, 1704 488 initial 1656 annealed 1615a 47b 1649, 1703, 1752 2226, 2292, 2384 48b initial 1656 2169, 2247, 2291 annealed 1613a 2340, 2369 LiBH4 2219, 2301, 2378 a uncertain value due to the low intensity of the absorption and overlap with other vibration modes. 87 Thus, the decomposition process in the solid-state differs significantly from that in solution, demonstrating again the ability of dihydrogen bonding to exert topochemical control in the 0H---HB to O-B conversion. Attempts to enhance the crystallinity of 48 by annealing at 120 °C under Ar resulted in a decrease of the XRD peaks’ intensities, until they completely disappeared after approximately 2 weeks, or 4 days if heated as 3 DMSO suspension. While the hydridic content remained unchanged (one H'), the 11B and 7Li MAS NMR chemical shifts moved 1.3 and 2.1 ppm downfield and upfield, respectively (Table 3.1). These changes point up the metastable nature of this topochemically controlled product. Moreover, the appearance of the BH stretching region in the IR spectrum changed considerably (Table 3.2) and apparently irreversibly, as the BH stretching absorptions did not revert to the original values even after 8 months at room temperature. A Van shift to shorter wavelengths in metal borohydrides generally indicates a stronger BH‘---W interaction,l '1 and therefore a similar situation could be present in 48. The observed shifts in the solid-state HB and 7Li MAS NMR spectra are in agreement with this supposition. 0n the other hand, no disproportionation seems to have occurred, since again the IR spectrum of annealed 48b is basically a summation of the corresponding spectra for 48 and 483 (Table 3.2). The result of a topochemical reaction, the structure of 48 was very likely controlled by kinetic factors, and its subsequent annealing presumably allowed structural relaxation into a thermodynamically more stable arrangement in which the oppositely charged BH(0R)3’ and Li+ units are closer in space. The profound modifications in the IR spectrum of 48 and the relatively small downfield shift in its HB NMR spectrum are in good qualitative agreement with a BH(0CH3)3'-~Li+ model, whose 88 Optimized Cs symmetrical structure at the RHF/6-3IG* level is shown in Figure 3.7. Its VBH frequency (scaled by 0.97 based on the ratio between the theoretical and experimental values of the strongest absorption in BI-Lt’) is estimated at 2418 cm’l, comparable with the 2368 cm'1 value found in 48 after annealing. The 11B NMR 8 value of -10.7 ppm calculated for this model structure is virtually identical with the corresponding value in BH(0CH3)3', predicting that the above mentioned interaction with Li+ should have minimal influence over the “B NMR chemical shift. H301 H CH3 Olfi;O-" do-Li = 1.785 Hap—o“ ,: dB-Li = 2.346 Figure 3.7. Calculated structure of LiBH(0CH3)3 with the interatomic distances in A. The fresh or annealed 48 is inert to ethanol, phenol, or acetone. However, it reacts with Pd(Ac0)2 in THF at room temperature to yield an amorphous hygroscopic green powder, soluble in DMSO, and presumably containing Pdo, as indicated by deposition of Pd black upon H20 hydrolysis. IR, 1H and 11B NMR spectroscopies showed the consumption of the BH, and attachment of the acetate group to the B (See Experimental). This material might be a very effective catalyst for various transformations, considering the presumable high dispersion of the Pd centers on the polymeric borate support. Mechanistic Study of the Topochemical Dihydrogen to Covalent Bonding Transformation in LiBH4~TEA.112 There are a few fundamental issues to be considered regarding the mechanism of the topochemical decomposition of 47 in particular, and of dihydrogen-bonded complexes in general. As for many other solid-state reactions, 89 conventional concepts like concentration, reaction order, or molecularity have little applicability here. More relevant in this context are the appearance and morphological evolution of the product phase, as well as its compatibility and "solubility" in the reactant phase (See Chapter 2). As thermally initiated solid state reactions that yield both solid and gaseous products, decompositions of 47 (and other dihydrogen-bonded systems), are complex processes involving not only chemical steps such as breaking and formation of bonds, but also physical transformations like destruction of the initial lattice, reactant/product solid solution formation (with possible separation of the product phase), diffusion and desorption of H2, and heat transfer. The variations in the crystals’ morphology are of particular interest as they can provide critical information about the mechanism of decomposition.1 '3 The mechanism of this solid state decomposition at the molecular level is particularly relevant for the successful design of the next generation of dihydrogen-bonded systems, and also for a better understanding of such fundamental processes as proton transfer or covalent bond formation, extensively studied in the gas phase and solution, but considerably less in the solid state. Figure 3.8 presents typical optical micrographs showing crystals of 47 at different stages of decomposition at 110 °C. The initial transparent crystals gradually become opaque as the reaction progresses toward completion, suggesting the separation of a new phase. There is no visible reaction front advancing through the crystal; instead, the process appears to start randomly and proceed uniformly in the crystal bulk. The size and morphology of the crystals remained virtually unchanged during decomposition, practically eliminating the possibility of melting. 90 Figure 3.8. Typical microscopic view (transmitted light) of LiBH4-TEA solid-state decomposition: (3) initial crystal, (b) 10 min at 110 °C, (c) final decomposed crystal; Scale bar: 100 um. 91 When decomposition of 47 was monitored in situ by HB solid-state MAS NMR spectroscopy, no other boron species than the initial BH.{ and the final trialkoxyborohydride could be detected, probably because of the increased reactivity and the short lifetime of the intermediate mono- and dialkoxyborohydrides. Integration of the two well-separated peaks proved to be a suitable way to measure the extent of the solid- state decomposition, and thus to study the kinetics of this topochemical reaction independent of other physical transformations that occur during the process. The reaction was studied at temperatures between 105 and 120 °C, using LiBH4-TEA samples from the same batch for each experiment to eliminate any possible error introduced by sample variation, as significant differences in decomposition rates was noticed among various preparations, a feature characteristic of many solid-state reactions.95 The resulting conversion-time curves (Figure 3.9) have the typical sigmoid shape characteristic for most solid-state decompositions of the type Asond —) Bsohd + Cgas (See Section 2.3).94 0.9 0.8 +- x A ‘0300‘ Q T xx“x°‘. DOD 0.7+ g? ‘ X A 0.6» and? 1‘ “ 0.5+ xta? “f" a 0.4-“ x Q nu. 0.3‘”x{‘~ 02» a? . x“ 0.1}? 04: i i a 0 200 400 600 800 t(min) Figure 3.9. Conversion (0t) vs. time curves for the solid-state decomposition of 47 : O = 105°C;D=110°C;A=115°C;X= 120°C. 92 The agreement of our data to various solid-state decomposition models (Section 2.3)94 were compared, and the best match was found for a nucleation and two—dimensional ”2 against time for different growth mechanism. The corresponding plots of [- ln(l - (1)] temperatures studied are presented (Figure 3.10), together with the obtained rate constants k, the correlation coefficients of the linear regression analysis R2, and the conversion ranges for which the Avrami-Erofeev law is obeyed (Table 3.3). 1.2 3 0.8 ‘r ,(‘x “A“ n D C 06 _0_ Xx A on D 5 I “‘ an O. . —'-: 0.4 -- x a . °° ll xz?‘ u n .0 ° 8 0.2 ~~ , 0° 70 0 h 1r . 0 50 100 150 200 250 t (min) 112 Figure 3.10. Plots of the Avrami-Erofeev law, [-ln( l-a)] against time, for the solid state decomposition of 47: O = 105 °C; CI = 110 °C; A = 115 °C; x = 120 °C. It could be speculated that the two-dimensional expansion of nuclei might originate in the crystal structure of 47, consisting of one-dimensional dihydrogen-bonded ribbons linked by conventional H-bonds in overall extended layers. Once decomposition has started, it is more likely it will propagate within the same layer, where the H-bonding network is disrupted, weakening thus the compactness of the crystalline environment. By comparison, a similar mechanism, but with a three-dimensional growth of nuclei (n = 3), gave a significantly worse match (av. R2 = 0.9852 vs. 0.9984 for n=2). 93 Table 3.3. Rate constants k for the solid state decomposition of 47, calculated from the Avrami-Erofeev law for nucleation and two-dimensional growth in the specified conversion ranges, together with the correlation coefficients of the linear regression analysis R2. Temp (°C) k (min’1)x104 Conversion Range R2 105 37 i 1.4 0.284-0.681 0.9980 110 56 i 2.6 0.090-0.480 0.9988 115 66 :1: 2.8 0.224-0.606 0.9979 120 111 i 0.9 0064-0531 0.9990 The measured rate constants for the solid-state decomposition of 47 at temperatures between 105 and 120 °C obey the Arrhenius equation (R2 = 0.9755), as illustrated by the linear dependence of log k against l/T (Figure 3.11). At temperatures above 120 °C, the solid-state decomposition of 47 becomes faster than predicted by the Arrhenius equation. A possible explanation is that the dissipation of the resulting heat becomes too slow compared to decomposition rate, eventually leading to the autoacceleration of reaction. An activation energy of 21.0 :1: 2.4 kcal/mol for the solid- state decomposition of 47 can be estimated from our data. Similarly, using the Eyring equation, the activation parameters for decomposition, AH“2 and AS“, were calculated to be 20.1 :1: 2.4 kcal/mol and -16.8 d: 6.2 e.u., respectively. 94 5.5 —~ 541. O -lnk 4.5 «~ . 4 i # 0.0025 0.00255 0.0026 0.00265 in (K") Figure 3.11. Arrhenius plot for the solid-state decomposition of 47 at 105-120 °C. When the isoconversional method98 (Eq. 5, Section 2.3) was used, comparable activation energies within experimental errors were obtained (Figure 3.12). .a: Efiffilifigm :T I I L I l L 0 05 0'2 ole 0!4 0!5 0!6 0.7 (I Figure 3.12. Variation of activation energy with conversion for the solid-state decomposition of 47, obtained using the isoconversional method. However, analyses starting with t0,,- initial times set at various extents of reaction 0t, yielded different Ea values. For instance, when ta,- was set for CL = 0.1, the resulting 95 activation energies varied between 14.9 and 24.1 kcal/mol. In comparison, the Avrami- Erofeev (n = 2) model analysis did not show much variation. Since the kinetic measurements were done by in situ 11B NMR spectroscopy which allowed direct monitoring of the appearance of the final trialkoxyborohydride product, independent of other physical processes such as nucleation, phase separation, or diffusion and desorption of H2, and considering the fact that no phase transition occurred prior to decomposition, as shown by powder X-ray diffraction, it is reasonable to assume that the kinetic parameters found in this study are directly related to the chemical transformations responsible for decomposition.114 Moreover, the present activation parameters are comparable with the activation enthalpy of 20.6 i 1 kcal/mol and activation entropy of -22.3 i 3 en. found for the hydrolysis of ER; in neutral water, and associated with the rate limiting proton transfer step.loo However, the mechanism could be different in the solid-state, with the H2 evolution slowed down by the crystal constraints, possibly becoming the rate-determining step. To address this question, we studied the solid state decomposition of LiBH4-N(CH2CH2OD)3, the analogue of 47, with the OH groups deuterated. H/D exchange between Eli; and OD groups of the TEA is expected for a scheme involving fast, reversible proton transfer followed by slow H2 loss. However, no HID exchange was observed at any stage of decomposition, under various conditions, as indicated by the IR or 1H NMR of the partly or fully decomposed deuterated complex. This experiment suggests that as in solution, the proton transfer is slow compared to H2 loss and B-0 covalent bond formation (Figure 3.13). It should be stressed here that these results are interpreted using a stepwise mechanism for the solid state decomposition based on analogy with the aqueous hydrolysis of BH4' in neutral 96 water, and theoretical calculations supporting the existence of the BH5 intermediate'o' In solution, a one step reaction with a four center transition state could be unambiguously eliminated based on kinetic isotope effects and hydrolysis experiments in the presence of trimethylammonium ion.l '5 In our solid-state system, on the other hand, the possibility of a concerted mechanism cannot be ruled out. However, considering the similarity of the reacting partners and the activation parameters found in this study and in solution, it appears the same mechanism seems likely to apply for both. H\ BH 3 BH ,H 5 H N f? / WW N~--.:l‘|_'. slow : Elu“2i,-I+ fast N~-~-:l'_l+ /,;’| ‘/.;’\| ‘/,;’\l O O O O O O \ I \ I \ I HH HH HH Figure 3.13. Proposed mechanism for the first O-H---H-B to O-B topochemical conversion in 47. To all appearances, the measured activation parameters in this study correspond to rate-limiting proton transfer at the interface between the crystalline dihydrogen-bonded system 47 and the newly formed covalent product 48. The reacting partners in this region are presumably more flexible than in the perfectly ordered environment of the undisturbed crystal. The somewhat lower activation entropy found for the solid-state process, compared to the similar reaction in water, could be attributed to the preorganization imposed by the dihydrogen bonding network in the crystal. Despite the 97 heterogeneous nature of decomposition, the topochemical information is nonetheless transferred through the interface to the newly formed covalent product. Toward Crystalline Covalent Solids: Crystal-to-Crystal Dihydrogen to Covalent Bonding Transformation in NaBI-L-THEC (49).l '6 The N aBH4-THEC (THEC = NNJV”,N”’-tetrakis—(2-hydroxyethyl)cyclen) complex exemplifies a system that combines both the closed loop and the globular cation strategies, leading as a result for the first time to a crystalline covalent product, in a crystal-to-crystal thermal decomposition. The complex precipitates from a 1:1 mixture of N aBH4 and THEC in 2-propanol, and its crystal structure is shown in Figure 3.14. The THEC ligand adopts a conformation with all hydroxyethyl arms oriented toward the same face of the azacrown ring, as observed in previously reported metal complexes containing this ligand.117 The nitrogen and oxygen atoms are thus organized in a pseudo-cubic geometry, encapsulating the octacoordinated Na cation. The N a+THEC units have thus the appropriate geometry for the formation of the desired closed-looped dihydrogen-bonded assembly. 98 ,3; 012AA) ‘92 ,W0(1AA)" .."- .', j. m i i ‘3 011A) ' ‘j "" Figure 3.14. X-Ray crystal structure of N aBIheTHEC illustrating the self-assembly in dihydrogen-bonded dimers. Indeed, these cations self assemble in D2 symmetrical dimers, held together by four conventional O(2)-H---O(1) hydrogen bonds, complemented by four orthogonal O(1)- HmH-B proton-hydride interactions involving the borohydride anions (Figure 3. 14). The observed H---H distance is 2.00 Along, and the OH---H-B and O-H---HB angles display values of 115 and 151°, respectively. There is no hydrogen bonding interconnecting the dimers, which are packed in two-dimensional layers (Figure 3.15). The layers are stacked in a parallel fashion, creating thus one—dimensional channels along the (1 axis, in which the BH4' anions reside. 99 Figure 3.15. Crystal packing in NaBH4-THEC. The association in closed loops, as well as the packing mode of the large globular {Na2(TI-IEC)2}+2 cations, appear to fulfill the two geometrical requirements that were predicted to favor the transfer of crystallinity from dihydrogen- to covalent—bonded networks. Solid—state decomposition of 49, induced by heating under an inert atmosphere at 160 °C for 20 h, resulted in complete elimination of H2, as indicated by the FT-IR and HB solid state MAS-NMR spectra of the resulting material. Thermogravimetric analysis showed a remarkably sharp 1.88 % weight loss between 168.0 and 168.8 °C, which corresponds to 3.63 mol H2. The solid-state decomposition product exhibits high crystallinity, as indicated by microscopic examination, which showed good transparency to polarized light for the decomposed crystals (Figure 3.16). The reaction appears to be crystallographically homogeneous, in direct contrast to the LiBH4-TEA dihydrogen- bonded system, where the covalent product nucleated as a separate phase. Figure 3.16. Polarizing microscopic view of NaBH4-THEC crystal-to—crystal decomposition: a) initial crystal; b) after 20 h at 160 °C. Scale bar = 100 um. That the solid state decomposition of 49 is a single-phase transformation is also supported by powder XRD analysis, which shows a gradual shift of the diffraction pattern as the reaction progresses. As depicted in Figure 3.17, the final covalent product exhibits high crystallinity, in agreement with the microscopic observations. For comparison, an amorphous solid results when 49 is decomposed in the melt at 200 °C for 10 min (Figure 3.17c). 101 3) b) C) ”W" 2 l l TMWTWH 7 0 1 30 40 50 20 20/0 —————> Figure 3.17. Powder XRD patterns for: a) dihydrogen-bonded complex 49; b) solid-state decomposition product; c) product of decomposition in the melt. The unit cell parameters for the decomposition product, calculated using the powder XRD data, are: a = 8.600(34) A, b = 13.86404) A, c = 16.895(27) A, and v: 2014(11) A3. They correspond to 3.3, 2.5, 3.5, and 9.0 % shrinkage, respectively, relative to 49. For this comparison, the following values derived from powder XRD at room temperature were used for 49: a = 8.894(7), b = l4.214(13), c = 17.509(17), V= 2214(3). The evolution of unit cell parameters during the solid-state decomposition of 49 at 120 °C could be monitored by powder XRD (Figure 3.18). 102 V/A‘ _4p_ 50 4. 50 t/h t/h Figure 3.18. Evolution of unit cell axes (a) and volume (b) during the solid-state decomposition of NaBH4-THEC at 120 °C, monitored by powder XRD. 103 As shown in Figure 3.18, the initial crystal lattice smoothly transforms into the corresponding lattice of the product, indicating a crystal-to-crystal process. Despite sustained efforts, the detailed structure of the decomposed crystal has remained elusive to date. The relatively big change in the unit cell accompanying the solid-state decomposition ultimately results in deterioration of the single crystals’ quality, precluding their X-ray structural investigation. Also, the unexpected insolubility of the product in common organic solvents does not allow its recrystallization. Considering the crystal packing of 49, with no interdimer hydrogen bonding present, formation of discrete molecular cages is expected from the solid-state decomposition. Crosslinking between dimers would require substantial intermolecular rearrangements, which is highly unlikely in a crystallographically homogeneous, lattice-controlled transformation. Figure 3.19 presents a possible structure for the putative molecular cages, optimized with the DFT method at the B3LYP/6-3 10* level.1 '8 No unusual values for the bond lengths and angles were observed in the resulting C2 symmetrical structure, suggesting minimal strain in these covalent dimers. When 49 was decomposed in DMSO by heating at 120 °C for 48 h, a crystalline product precipitated in high yield. X-ray structural analysis of the resulting crystals revealed a centrosymmetric molecular cage very similar to that expected for the solid- state reaction (Figure 3.20). However, one of the hydroxyethyl arms in each of the two THEC ligands in the dimer does not bind the boron atom, nor does it complex the Na cation, being replaced by an OH group, presumably from water present in the solvent. Except for the Na, the four N atoms, and C(4) and C(6), all the atoms in the structure are disordered over two sites, with 90.1 and 9.9 % occupancies, respectively. 104 Figure 3.19. Molecular model of the product from the solid-state decomposition of 49; selected bond lengths [A] and angles [°]: B-O(1) 1.458, B—O(2) 1.482, B-O(3) 1.497, B- O(4) 1.474, O(1)-B-O(2) 112.5, O(1)-B-O(3) 109.1, O(1)-B-O(4) 114.6, B-O(1)-C 123.7, B-O(2)-C 122.3, B-O(3)-C 125.9, B-O(4)-C 125.6, N(l)-Na 3.003, N(2)-Na 2.714, N(3)- Na 2.435, N(4)-Na 2.565, O(2)-Na 2.670, O(3)-N a 2.276, C(4)-Na 2.228. 105 ‘nnt, .. AD N12A1 ’30» 'v Figure 3.20. X-Ray crystal structure of the major conformation found in the covalent dimer obtained by decomposing 49 in DMSO; selected bond lengths [A] and angles [°]: B-O(1) 1.502(2), B-O(2) 1.511(2), B—O(3) 1.450(2), B-O(5) 1.457(2), O(1)-B-O(5) 111.6(1), O(2)-B-O(5) 109.8(1), O(3)-B-O(5) 111.6(1), B-O(1)-C 121.7(1), B-O(2)-C 113.2(1), B-O(5)-C 122.8(1), N(1)-Na 2.519(1), N(2)-Na 2.540(1), N(3)-Na 2.565(1), N(4)-Na 2.682(1), O(1)-Na 2.371(1), O(2)-Na 2.796(1), O(3)-Na 2.391(1), C(4)-H 082(2), O(4)H---O(2) 1.92(3), O(4)---O(2) 2.735(4), O(4)-H-O(2) 175.1(3). 106 Like the solid-state decomposition product, the compound obtained from DMSO is surprisingly insoluble in common organic solvents, despite its relatively small size, overall neutral charge, and lack of intermolecular hydrogen bonding. In situ llB MAS NMR spectroscopy proved again to be a convenient method for monitoring the solid-state dihydrogen to covalent bonding transformation and measuring the kinetics of decomposition. As in the case of the heterogeneous transformation in LiBH4-TEA, no boron species other than the initial borohydride and the final borate could be observed. The extent of reaction at 160 °C could be easily measured by integration of the two peaks, and the conversion-time curve is depicted in Figure 3.21. After an initiation period, the reaction is accelerated up to approximately 75% conversion, after which time it is slowed down significantly. No common kinetic model for solid-state decompositions was found to fit the existing data. For comparison, the extent of lattice conversion vs time was plotted, using the data from the solid-state decomposition of 49 at 120 °C, monitored by powder XRD. As illustrated in Figure 3.22, the obtained curve is qualitatively comparable with the one obtained from the llB NMR measurements. The enhanced reaction rate at the intermediate stages of decomposition might originate in the increased strain caused by the enforced coexistence of the initial and final crystal lattices. On the other hand, powder XRD examinations at different extents of decomposition showed a gradual deterioration of crystallinity as the reaction progressed, followed by significant recovery of the diffraction pattern at the late stages of reaction. This increased lattice looseness at intermediate conversions appears to facilitate the chemical reaction relative to the initial and final stages, when the reacting molecules are constrained in a more rigid, unperturbed crystalline lattice. 107 1 00” 0.8 _. o ’ .0 0.6 ~~ v’ 5 04 ¢ . 1r 0” 0 2 —— O ,o o 0 e i 0 100 200 300 t (min) Figure 3.21. Conversion (or) vs time curve for the solid-state decomposition of NaBH4-THEC at 160 °C, obtained using in situ “B MAS NMR. 1 I 0.8 . . $- - - ' g: 0.6 7 Z >.c 0.4 « ' ' > 0.2. _ o a . . 0 50 100 150 Nb) Figure 3.22. Lattice conversion (VO‘V/Vo-Vf) vs time curve for the solid-state decomposition of NaBHa-THEC at 120 °C, obtained using powder XRD. 108 In summary, the first crystal-to-crystal dihydrogen to covalent bonding transformation was carried out. This outcome appears to be the result of the judicious engineering of the starting dihydrogen-bonded system, which was designed to satisfy both the closed loop and globular cation geometrical criteria. Structural Variations from the NaBI-h-THEC System. The N aBHa-THEC complex demonstrated that careful design of dihydrogen-bonded crystals permits transfer of crystallinity to the covalent products resulting from their solid-state decomposition. The next step is to extend this concept to the construction of higher-dimensional crystalline covalent systems, a class of compounds to which few purposeful synthetic paths exist. Further elaboration of the dihydrogen-bonded building blocks present in 49, which proved to possess the required geometrical and solid-state reactivity prerequisites, appears to be the most potentially successful approach toward our objective. However, an important and necessary condition is that the dihydrogen-bonded dimeric motif observed in 49 be robust enough to survive substantial modifications in molecular structure. This section presents some preliminary results from the systematic investigation of the relationship between molecular structure and supramolecular organization in metal borohydride complexes with THEC-type ligands. We seek an understanding of the factors that govern the crystal packing in these dihydrogen-bonded systems, and possibly to control them for the crystal engineering of the desired extended dihydrogen-, and ultimately covalent-bonded networks. We first explored the role of the metal borohydride in determining the self- assembly into dihydrogen-bonded dimers. The LiBH4 and KBH4 complexes with THEC were therefore synthesized to investigate the effect of changing the metal cation upon 109 crystal structure. Unfortunately, the modest crystallinity of the former, and the propensity of the latter to form polycrystalline aggregates instead of single-crystals, precluded their detailed structural elucidation. The NaCNBH3~TI~IEC (50) complex was then synthesized to explore the contribution of the hydridic partner in the formation of the closed-looped dimers. The white crystalline compound melts at 230-234 °C with decomposition and gas evolution. However, its crystal structure, which is depicted in Figure 3.23, reveals the absence of any O-HmH-B close contacts, the cyanoborohydride anion participating with the N atom via conventional H-bonding instead. Figure 3.23. X-Ray crystal structure of NaNCBHyTHEC. 110 Figure 3.24. a-Helix formed by N aNCBH3-THEC in the solid-state. lll Notably, instead of forming dimers, in this case the Na-THEC units self-assemble into extended a-helices, via bifurcated O-H---N---H-O hydrogen bonds involving the CN groups (Figure 3.24). The helices are further associated in two-dimensional layers by O- H---O interactions. It thus appears that the O-H---H-B dihydrogen bonds in 49 are essential for the formation of the {N a2(THEC)2}+2(BH4')2 dimers, as their absence in 50 resulted in a completely different packing. NaBHa-THPC (51) (THPC = 1,4,7,10-tetrakis((S)-2-hydroxypropyl)-1,4,7,10- tetraazacyclododecane) was synthesized to explore the effect of subtle molecular modifications in the ligand upon crystal packing. It was also hoped that the introduction of the Me groups would improve the solubility of the covalent dimers (if they are formed) in organic solvents. H,’ Me Me e\<\NH 473$ H OH Na H0 3H4- $7? 3%. 51 The complex precipitates as a white solid upon mixing NaBHa and THPC in 2-propanol. Its crystal structure is presented in Figure 3.25. The most notable feature is the formation of hydrogen-bonded dimers, which, however, display an S-shaped topology, as opposed to the O-shaped one in 49. The most probable cause of this difference is the steric repulsion that would result between the Me groups, in an O-shaped configuration in 51. 112 Figure 3.25. X-Ray crystal structure of NaBHa-THPC; dihydrogen bonding parameters (A, °): O(4)H-~-HB(1) (lower) 2.106, O(4)H---HB(1) (upper) 2.266, O(3)H---HB(1) 2.005, O(7)H---HB(2) 1.927, O(8)H---HB(2) 1.676, O(2)H~-HB(2) 1.904, O(4)H-H-B(1) (lower) 96.1, O(4)H-H-B(1) (upper) 88.7, O(3)H-H-B(1) 113.8, O(7)H-H-B(2) 110.7, O(8)H-H- B(2) 120.2, O(2)H-H-B(2) 133.3, O(4)-H-HB(1) (lower) 162.5, O(4)-H-HB(1) (upper) 143.9, O(3)-H-HB(1) 153.4, O(7)-H-HB(2) 155.4, O(8)-H-HB(2) 153.0, 0(2)-H-HB(2) 169.3. 113 The two THPC units in the dimer are held together by O(1)H---O(6) and O(5)H---O(2) conventional hydrogen bonds, as well as by O(7)H-~HB(2)H~~-HO(2) and O(8)H---HB(2)H---HO(2) dihydrogen bonds. The second borohydride anion interacts with the remaining O(3)H and O(4)H from one THPC ligand. As in 49, no hydrogen bonding interconnects the dimers. The solid-state decomposition of 51 was unfortunately precluded by partial liquefaction during heating, apparently due to the low melting point of the product(s). More profound structural variations in the THEC supramolecular building block were brought about by replacing one of the hydroxyethyl arms with a 9-methylanthracene group. The intention was to test the ability of the resulting tripodal ligand to maintain the association in dimers. Recurrence of this supramolecular motif could ultimately open the way to extended networks, by simple attachment of the tris-hydroxycyclen units to multidentate spacers, using the remaining N atoms. Furthermore, monitoring the orientation of the anthracene chromophores via fluorescence spectroscopy might offer an additional handle on the evolution of order during the solid-state decomposition. The 9-(1’,4’,7’-tris-hydroxyethylcyclen)methylanthracene (52) ligand was synthesized from 9-(1’,4’,7’,10’-tetraazacyclododecyl)methylanthracenel '9 and ethylene oxide (Scheme 3.6), and its structure was confirmed by X-ray crystallography (Figure 3.26). Upon mixing with NaBHa in 2-propanol, the NaBH4-52 complex (53) was formed, which precipitated as a white crystalline solid. Elemental analysis and 1H NMR indicated the crystallization with 1 mol of solvent. The inclusion of 2-propanol was also supported by TGA, which showed a 10.35% weight loss, in good agreement with the 10.14% theoretical value. Single crystals, however, were grown from CH3CN/Et2O. 114 A O ”OH O 0 ° ‘ ‘ O N N E * E N 1:] EtOH, H20 N :1 \__/ \_\ HO OH 52 Scheme 3.6 Figure 3.26. X—Ray crystal structure of 9-(1’,4’,7’-tris-hydroxyethyl- cyclen)methylanthracene. 115 Figure 3.27. X-Ray crystal structure of 53; normalized dihydrogen bonding parameters (A, °): O(1)H---HB 1.692, O(2)H---HB 1.661, O(3)H~--HB 1.730, O(1)H---H-B 114.5, O(2)H--~H-B 117.8, O(3)H-~H-B 116.1, O(1)-H--~I-IB 166.8, O(2)-HmHB 155.1, O(3)- H---HB 154.6. 116 X-Ray structural examination of crystals of 53 obtained from acetonitrile-ether revealed the inclusion of 1 mol of disordered 320. The crystal structure (EtzO was omitted) is depicted in Figure 3.27, and shows the self—assembly in dihydrogen-bonded centrosymmetric dimers. However, compared with the previous structures, no O-H---O interaction is present in this case; instead, the BH4' anions are intercalated between the two ligands, forming three dihydrogen bonds with the OH groups. No conclusive results from the solid-state decomposition of 53 have been obtained to date, due to the insufficient quantity of its ether clathrate available. In summary, NaBH4 and various hydroxyethylcyclen ligands persistently self- assemble in dihydrogen-bonded dimers in the solid-state, as demonstrated by the crystal structures of 49, 51, and 53. However, the exact topology of the dimers varies from one case to another, in response to structural modifications in the ligand. Nevertheless, this consistent recurrence of the general dimeric motif makes these dihydrogen-bonded units sufficiently reliable building blocks for supramolecular synthesis. Their incorporation into more elaborate structures, as suggested in Figure 3.28 would offer real prospects for the construction of extended dihydrogen-bonded, and ultimately covalent solids, with controlled architectures. 117 Figure 3.28. Prospective multidentate ligands for the construction of extended dihydrogen-bonded and covalent networks. 118 3.2. Conclusions 0 Dihydrogen bonds can play an important role in defining solid-state supramolecular structures, comparable to conventional hydrogen bonds, as demonstrated by the crystal structures of all H---H bonded compounds studied, where this interaction significantly influenced the arrangement of molecules in crystals. Thus, with its capacity to control crystal packing, dihydrogen bonding, like traditional H—bonding, finds its place among the tools that chemists can use for supramolecular synthesis. The additional feature that makes this unconventional H-bonding particularly interesting is the ability to react, trading the weak HmH interactions for strong covalent bonds. As a direct consequence of this character, dihydrogen bonding has the potential to induce topochemical control, providing access to new materials otherwise not achievable. 0 The NaBH4-TEA complex has demonstrated the topochemical control concept. Its crystal structure shows multiple dihydrogen bonds between 3114' and the three OH groups from TEA. Loss of 3 moles of H2 in the solid-state leads to a polymeric trialkoxyborohydride structure, whereas solution or melt decompositions yield NaBH4 and a polymeric borate with no hydridic hydrogen left. Remarkably, more H2 is lost in the solid-state decomposition than in the melt, despite the much higher reaction temperature in the latter (130 °C vs 82 °C). However, the product of the solid-state decomposition showed poor crystallinity and the reaction times were exasperatingly long due to the relatively low temperature required for decomposition in order to avoid melting. 119 o In addition to close H---H contacts, the relative acidity/basicity of the proton- hydride pairs significantly affect the solid-state reactivity of the dihydrogen-bonded systems. This was demonstrated by changing the H’ donor from BI-L{ to CNBH3‘ in the TEA complexes, which resulted in substantial decreases in dihydrogen bonding and solid-state reactivity. o Fine-tuning of the dihydrogen bonds was possible in the TEA complexes by replacing the Li” for Na+, which resulted in exceptionally short HmH contacts and enhanced solid-state reactivity. Decomposition of LiBI-IarTEA is also topochemical, apparently leading to a metastable one-dimensional polymeric trialkoxyborohydride structure, in direct contrast to decomposition in DMSO solution, which yielded a polymeric hydride free borate and unconverted LiBI-I4, the disproportionation products. For comparison, no such disproportionation occurs during the solid-state decomposition, as demonstrated by H/D isotopic labeling experiments. 0 The high solid-state reactivity of LiBH4-TEA allowed the detailed study of the mechanism of H2 loss and covalent bond formation in this dihydrogen-bonded complex, both at macroscopic and molecular level. Using in situ solid-state llB NMR and optical microscopy, it was found that, like most solid-state reactions, this decomposition is heterogeneous, with separation of the product phase from the parent crystal. Kinetic analysis and H/D exchange experiments established that proton transfer between the OH groups of the TEA and the BHa' anions, at the reactant/product interface, is the rate- limiting step, with activation parameters AHat and AS"e of 20.1 i 2.4 kcal/mol and —l6.8 i 6.2 eu. These values are comparable with the analogous values found for the aqueous 120 hydrolysis of BHa' in neutral water, suggesting similar mechanisms for the solid and solution decompositions. o Preservation of crystallinity was demonstrated to be possible in the topochemical dihydrogen to covalent bonding transformation of N aBH4-THEC, which was designed to satisfy both the closed loop and large globular cation geometrical criteria. Remarkably, its solid-state decomposition is a crystal-to-crystal process, as indicated by polarizing microscopy and powder XRD. However, the observed 9% shrinkage of the unit cell led to significant deterioration of the single crystals’ quality, which in conjunction with the insolubility of the covalent product in organic solvents, precluded its detailed structural characterization. Nevertheless, this study is the first example to demonstrate that judicious engineering of dihydrogen-bonded crystals permits transfer of crystallinity to the covalent products resulting from their solid-state decomposition. 0 Preliminary results suggest that NaBH4 and various hydroxyethylcyclen ligands consistently self-assemble in dihydrogen-bonded dimers, as demonstrated by the crystal structure analysis of NaBI-14~TI~{EC, NaBH4-THPC, and N aBHa-[9-(l’,4’,7’-tris- hydroxyethyl-cyclen)methylanthracene]~Et2O. Consequently, incorporation of these dimeric building blocks in more elaborated systems may allow the rational construction of extended dihydrogen-bonded networks, and ultimately covalent solids, with controlled structures and functions. 121 3.3. Experimental Section Materials. Common chemicals were purchased from commercial sources and used without further purification. N—[2-(6-Aminopyridyl)]-acetamidine hydrochloride was prepared by a literature procedure. '20 TEA-d3 was obtained by repeated dissolution of TEA in D2O (99.9% D) followed by removal of the water at 85 °C under vacuum, until the 1H N MR (DMSO-d6) signal for the OH protons completely disappeared. LiBD4 was synthesized by metathesis,121 from NaBD4 (98%D). THEC was prepared by a literature procedure. 122 1 ,4,7, 10-Tetrakis((S)-2-hydroxypropyl)- 1 ,4,7, 10-tetraazacyclododecane) (THPC) was synthesized from cyclen and S(-) propylene oxide.123 9-(1’,4’,7’,10’- Tetraazacyclododecyl)methylanthracene was obtained by a literature procedure.“9 Physical and Chemical Characterizations. Melting points were taken on a Thomas Hoover instrument and are uncorrected. FT-IR spectra were measured in KBr pellets on a Perkin Elmer Spectrum 2000 instrument. 1H N MR (300.1 MHz) spectra were recorded on a Gemini-300 instrument. l3C NMR (75.4 MHz) and “B (96.23 MHz) NMR spectra were recorded on a Varian VXR-300 instrument. “8 (128.33 MHz), 23Na (105.81 MHz), and 7L1 (155.44 MHz) solid-state MAS NMR spectra were measured on a Varian VXR-400 instrument. llB solution NMR chemical shifts were referenced to B(OCH3)3 in CDC13. llB solid state MAS NMR chemical shifts were referenced to solid boric acid; 23 N a and 7Li chemical shifts were referenced to solid N 8C] and LiCl, respectively. A negative sign indicates chemical shifts upfield from references. X-ray powder diffraction measurements were conducted on a Rigaku-Denki RW400F2 diffractometer with monochromatic Cu Kat radiation, operated at 45 kV and 100 mA. Optical micrographs 122 were obtained with a Nikon AFX-DX microscope equipped with a Mettler FP82HT hot stage. TGA was recorded on a CAI-IN TGSystem 121 instrument at a heating rate of 5 °/min. Hydridic Content Analyses. A 30-120 mg sample of the hydride-containing material was placed in a 50 mL side-armed round bottom flask, connected by a rubber tube to a 50 mL burette, half-filled with distilled water. About 10 mL of diluted HCl (1- 2%) was rapidly added from a 25 mL addition funnel. The amount of the evolved H2 was obtained by measuring the volume of the displaced water in the burette, after adjusting the gas pressure in the flask (by rising it to the corresponding height) to the atmospheric pressure. Typical uncertainty: :1: 0.3 H' equivalents. In Situ llB MAS Solid State NMR. All spectra were recorded at a MAS frequency of 3.8 kHz, using 2.0 [.18 N2 pulses, with a recycle delay of 5 s, to allow full relaxation of all boron species.124 For each experiment, 48 scans were acquired. The observed spectra were deconvoluted into Lorentzian lines, and the ratios of the resulting integrals, corresponding to product and starting material, respectively, were used to estimate the extent of decomposition. Theoretical Calculations. All calculations were carried out with the GAUSSIAN 94 and 98 packages.'25 All structures were fully optimized and confirmed (except for the model of the decomposition product of 49) by vibrational analysis to be minima on the potential energy surface. Single Crystal X-Ray Structure Determination. X-Ray crystallographic measurements were carried out on a Siemens SMART CCD diffractometer with graphite- monochromated Mo Kat radiation (A = 0.71073 A), operated at 50 kV and 40 mA. The 123 structures were solved by direct methods and refined on F2 using the SHELXTL software '26 Absorption corrections were applied using SADABS, part of the SHELXTL package. software package. All non-hydrogen atoms were refined anisotropically. Hydrogen atoms were located from difference Fourier maps and refined isotropically for all the structures. For the borohydride ions in 49, however, only the hydridic hydrogens involved in dihydrogen bonding could be found, and the remaining H atoms from 8114' were calculated and placed in idealized positions. Similarly, for B(2)H4' in 51, the two noninteracting hydridic hydrogens were calculated, and, in this case, the 4 H were subsequently constrained in an idealized tetrahedral geometry, with B-H distances of 1.21 A. 124 Table 3.4. Crystal data and structure refinement for NAPA H3BCN (44). Empirical formula Formula weight Temperature Wavelength Crystal system Space group Unit cell dimensions Volume Z Density (calculated) Absorption coefficient F(000) Crystal size 0 range for data collection Index ranges Reflections collected / unique Refinement method Data / restraints / parameters Goodness-of-fit on F2 Final R indices [I>20(I)] R indices (all data) Extinction coefficient Largest diff. peak and hole 125 C8H14BN5 191.05 143(2) K 0.71073 A Triclinic P] a = 7.104(2) A b = 9.390(4) A c = 17.736(10) A a = 99.38(4)° 13 = 95.99(4)° y = 111.53(3)° 1068.3(9) A3 4 1.188 g/cm3 0.077 mm" 408 0.6 x 0.3 x 0.15 mm 2.37 to 28.47° -8$h$8,-11$k$12,-10$1$23 3118 / 3095 [R(int) = 0.0469] Full-matrix least-squares on F 3095 IO / 366 1.095 R, = 0.0455, sz = 0.1130 R1 = 0.0577, sz = 0.1238 0.022(3) 0.262 and -0.163 eA'3 Table 3.5. Crystal data and structure refinement for NaBHa-TEA (45). Empirical formula Formula weight Temperature Wavelength Crystal system Space group Unit cell dimensions Volume Z Density (calculated) Absorption coefficient F(000) Crystal size 0 range for data collection Index ranges Reflections collected / unique Refinement method Data / restraints / parameters Goodness-of-fit on F2 Final R indices [I>20'(I)] R indices (all data) Extinction coefficient Largest diff. peak and hole 126 C6H|9BNNaO3 187.02 173(2) K 0.71073 A Monoclinic P21/Il a = 9.23510(10) A b = 7.35120(10)A = 15.8083(2) A or = 90° 8 = 103.6880(10)° Y = 90° 1042.73(2) A3 4 1.191 g/cm3 0.123 mm" 408 0.52 x 0.38 x 0.31 mm 2.34 to 28.15° -11$h$12,-9Sks9,-20.<_1S.20 9787 / 2482 [R(int) = 0.0211] Full-matrix least-squares on F2 2482 / 0 / 186 1.060 R1 = 0.0250, sz = 0.0726 R. = 0.0292, sz = 0.0743 0.0102(31) 0.357 and -O.165 eA'3 Table 3.6. Crystal data and structure refinement for NaNCBH3-TEA (46). Empirical formula Formula weight Temperature Wavelength Crystal system Space group Unit cell dimensions Volume Z Density (calculated) Absorption coefficient F(000) Crystal size 0 range for data collection Index ranges Reflections collected / unique Refinement method Data / restraints / parameters Goodness-of-fit on F2 Final R indices [I>20'(I)] R indices (all data) Largest diff. peak and hole 127 C7H13BN2NaO3 212.03 173(2) K 0.71073 A Monoclinic sz/C a = 7.55160(10)A b = 15.1232(2) A c = 10.57070(10) A or = 90° [3 = 105.5380(10)° Y = 90° 1163.10(2) A3 4 1.211 g/cm3 0.121 mm'1 456 0.62 x 0.18 x 0.10 mm 2.41 to 28.17 deg. -9.<.h510,-195k$20,-13515 14 9974 / 2753 [R(int) = 0.0549] Full-matrix least-squares on F2 2753 / 0/ 199 1.053 R. = 0.0352, sz = 0.0891 R. = 0.0442, sz = 0.0945 0.232 and -0.219 eA'3 Table 3.7. Crystal data and structure refinement for LiBHa-TEA (47). Empirical formula Formula weight Temperature Wavelength Crystal system Space group Unit cell dimensions Volume Z Density (calculated) Absorption coefficient F(000) Crystal size 0 range for data collection Index ranges Reflections collected / unique Refinement method Data / restraints / parameters Goodness-of-fit on F2 Final R indices [I>2o(I)] R indices (all data) Absolute structure parameter Extinction coefficient Largest diff. peak and hole 128 C6HloBLiN03 170.97 173(2) K 0.71073 A Monoclinic P21 a = 7.7156(5) A b = 17.0039(11) A c = 7.8656(5) A 01 = 90° 13 = 98.1330(10)° 7: 90° 1021.55(11) A3 4 1.112 g/cm3 0.081 mm" 376 0.44 x 0.23 x 0.21 mm 2.40 to 25.00° -9Sh$6,-l7$kS20,-8SIS9 5330/ 3229 [R(int) = 0.0339] Full-matrix least—squares on F2 3229/ l / 370 0.979 R. = 0.0417, wR2 = 0.0768 R1 = 0.0619, wR2 = 0.0839 1.7(10) 0.015(3) 0.145 and -0.123 eA’3 Table 3.8. Crystal data and structure refinement for NaBHatTHEC (49). Empirical formula Formula weight Temperature Wavelength Crystal system Space group Unit cell dimensions Volume Z Density (calculated) Absorption coefficient F(000) Crystal size 0 range for data collection Index ranges Reflections collected / unique Refinement method Data / restraints / parameters Goodness—of-fit on F2 Final R indices [I>20(I)] R indices (all data) Absolute structure parameter Extinction coefficient Largest diff. peak and hole 129 CtoH4oBN4NaO4 386.32 173(2) K 0.71073 A Orthorombic 1222 a = 8.8240(3) A b = 14.1344(6) A c = 17.35330) A 0: = 90° B = 90° Y = 90° 2164.34(15) A3 4 1.186 g/cm3 0.100 mm'1 848 0.23 x 0.18 x 0.13 mm 1.8610 28.20° -11ShSll,-185k518,-225.1.<_22 12854 / 2608 [R(int) = 0.0979] Full-matrix least-squares on F 2608/ 11 / 200 1.029 R1 = 0.0487, wR2 = 0.1221 R. = 0.0599, wR2 = 0.1399 0.7(5) 0.0027(9) 0.488 and 0250 eA'3 Table 3.9. Crystal data and structure refinement for the decomposition product of NaBHa-THEC in DMSO. Empirical formula Formula weight Temperature Wavelength Crystal system Space group Unit cell dimensions Volume Z Density (calculated) Absorption coefficient F(000) Crystal size 0 range for data collection Index ranges Reflections collected / unique Absorption correction Refinement method Data / restraints / parameters Goodness-of-fit on F2 Final R indices [I>20’(I)] R indices (all data) Extinction coefficient Largest diff. peak and hole 130 C16H34BN4N305 396.27 173(2) K 0.71073 A Monoclinic P21/n a = 8.7696000) A b = 130571000) A c = 17.7724(2) A 01 = 90° 13 = 96.36° Y = 90° 2022.500) A3 4 1.301 g/cm3 0.112 mm'1 856 0.47 x 0.39 x 0.29 mm 1.94 to 28.26° -11 shs 11,-16Sks 17,-23s1s22 20248 /4840 [R(int) = 0.0261] None Full-matrix least-squares on F2 4840 / 0 / 462 1.043 R. = 0.0352, wR2 = 0.0921 R1 = 0.0439, wR2 = 0.0966 0.0000(9) 0.261 and 0250 eA'3 Table 3.10. Crystal data and structure refinement for N aNCBH3-THEC (50). Empirical formula Formula weight Temperature Wavelength Crystal system Space group Unit cell dimensions Volume Z Density (calculated) Absorption coefficient F(000) Crystal size 0 range for data collection Index ranges Reflections collected / unique Refinement method Data / restraints / parameters Goodness-of-fit on 1:2 Final R indices [I>20'(I)] R indices (all data) Absolute structure parameter Extinction coefficient Largest diff. peak and hole 131 C 17H3gBN5NaO4 411.33 173(2) K 0.71073 A Orthorombic P112121 a = 25.2624(3) A b = 9.4053(2) A c = 9.4430(2) A 01 = 90° B = 90° 7 = 90° 2243.660) A 4 1.218 g/cm3 0.102 mm"1 896 0.65 x 0.29 x 0.26 mm 1.61 to 28.20° -32$h$32,-12$k$12,-12$1S 12 22515 / 5336 [R(int) = 0.0187] Full-matrix least-squares on F2 5336/ 1 /410 1.022 R1 = 0.0251, sz = 0.0637 R1 = 0.0276, WRz = 0.0646 007(17) 0.0058(9) 0.205 and -0125 eA'3 Table 3.11. Crystal data and structure refinement for N aBHa-THPC (51). Empirical formula Formula weight Temperature Wavelength Crystal system Space group Unit cell dimensions Volume Z Density (calculated) Absorption coefficient F(000) Crystal size 0 range for data collection Index ranges Reflections collected / unique Refinement method Data / restraints / parameters Goodness-of-fit on F2 Final R indices [I>20’(I)] R indices (all data) Absolute structure parameter Largest diff. peak and hole C20H483N4N304 442.42 173(2) K 0.71073 A Monoclinic P21 a = 9.7915(4) A b = 18.14490) A c = 14.6111(6) A 01 = 90° 13 = 96.867(2)° Y = 90° 25772708) A3 4 1.140 g/cm3 0.092 mrn'I 976 0.49 x 0.41 x 0.18 mm 1.80 to 28.24° -125hS12,-24S.k_<.24,-19Sl_<. 16 23846 / 11681 [R(int) = 0.0198] Full-matrix least-squares on F2 11681 I20 / 923 0.789 R1 = 0.0358, WRZ = 0.0968 R. = 0.0428, sz = 0.1023 0.09( 18) 0.517 and -0203 eA'3 132 Table 3.12. Crystal data and structure refinement for 9-(1’,4’,7’-tris-hydroxyethyl- cyclen)methylanthracene (52). Empirical formula Formula weight Temperature Wavelength Crystal system Space group Unit cell dimensions Volume Z Density (calculated) Absorption coefficient F(000) Crystal size 6 range for data collection Index ranges Reflections collected / unique Refinement method Data / restraints / parameters Goodness-of-fit on F2 Final R indices [I>20'(I)] R indices (all data) Largest diff. peak and hole 133 C29H42N403 494.67 173(2) K 0.71073 A Triclinic P1 a = 8.7352(3) A b = 1229710) A c = 13.7789(3) A at = 69.152(2)° 13 = 79.456(2)° y: 78.736(2)° 1346.11(7) A3 2 1.220 g/cm3 0.080 mm'l 536 0.39 x 0.18 x 0.08 mm 1.59 to 28.27° -11 shs 11,-16Sks 15,-18s15 17 12550 / 6195 [R(int) = 0.0383] Full-matrix least-squares on F2 6195 / 0 I489 1.268 R. = 0.0825, wR2 = 0.2187 R. = 0.1467, wR2 = 0.2369 1.571 and -0.708 eA'3 Table 3.13. Crystal data and structure refinement for NaBHa-[9-( l ’,4’,7’-tris- hydroxyethyl-cyclen)methylanthracene]rEt2O (53). Empirical formula C33H56BN4Na04 Formula weight 590.62 Temperature 173(2) K Wavelength 0.71073 A Crystal system Monoclinic Space group P21/n Unit cell dimensions a = 12.7329(6) A b = 19.0589(2) A c = 13.7260(6) A 01 = 90° 13 = 95.459(2)° “ y = 90° Volume 3315.8(2) A3 Z 4 Density (calculated) 1.183 g/cm3 Absorption coefficient 0.086 mm'l F(000) 1288 Crystal size 0.49 x 0.18 x 0.13 mm 0 range for data collection 1.83 to 28.25° Index ranges -16ShS16,-2SSkS25,-1851518 Reflections collected / unique 30419 / 7907 [R(int) = 0.0676] Refinement method Full-matrix least-squares on F Data / restraints / parameters 7907 / 0 / 541 Goodness-of-fit on F"2 1.027 Final R indices [I>20'(I)] R. = 0.0672, wR2 = 0.1646 R indices (all data) R. = 0.1547, wR2 = 0.1969 Largest diff. peak and hole 0.767 and -0.478 eA'3 134 Syntheses: N-[2-(6-Aminopyridyl)]-acetamidine cyanoborohydride (44). 2.61 g (0.014 mol) of N—[2-(6-Aminopyridyl)]-acetamidine hydrochloride and 0.88 g (0.014 mol) of NaNCBH3 were dissolved in 65 mL methanol and the resulting clear solution was left standing at rt for 12 h. After the removal of methanol in vacuo, 20 mL of THF were added and the solution was filtered. The filtrate was concentrated in vacuo and 100 mL of CH2C12 were added. The resultant precipitate was collected to yield 2 g (75%) of product as a white solid. X-Ray quality crystals were grown by diffusion of cyclohexane vapors into an isopropanol solution of 44. Mp = 119-120 °C; IR (KBr): v3” = 2338, 2219 cm", ) VCN = 2168 cm"; ‘H NMR (DMSO-d6): 5 7.47 (t, J = 8 Hz, 1H, py), 6.52 (s, 2H, py), 6.26 (t, J = 8 Hz, 2H. py), 2.31 (s, 3H, CH3), 0.21 (q, 1.... = 88.5 Hz, 3H, NCBH3); ”B NMR (DMSO-d6): 5 - 55.95 (q, JBH = 88.8 Hz, NCBH3); l3C NMR (DMSO-d6): 8 164.0, 158.1, 150.6, 140.1, 104.4, 100.1, 20.0. Anal. Calcd for C3H14N5B: C, 50.26; H, 7.33; N, 36.65. Found: C, 49.64; H, 7.57; N, 36.11. NaBH4-TEA (45). A mixture of 0.38 g (0.01 mol) NaBH4 and 2.8 mL (0.02 mol) triethanolamine in 25 mL THF was stirred at rt for 2 h. Filtration of the precipitate obtained and washing with THF yielded 0.95 g (51%) of 45. Crystals suitable for X-Ray crystallography were obtained by slow cooling of a solution of 45 in isobutanol from rt to -25 °C. Mp = 107-108 °C (dec.); IR (KBr): VBH = 2288 cm"; 1H NMR (pyridine-d5): 5 6.68 (br, 3H, OH), 3.87 (t, J = 5.1 Hz, 6H, OCH2), 2.67 (t, J = 5.1 Hz, 6H, NCH2), 1.30 (q, 1511 = 81.3 Hz, 4H, BH4); llB NMR (DMSO-d6): 5 —49.28 (quintet, 1311 = 81.4 Hz, BH4); ”B MAS NMR (6000 Hz): 5 - 47.9 (Aw/2 = 1826 Hz); ”C NMR (DMSO-d6): 5 135 58.9 (OCH2), 56.9 (NCH2); 23Na MAS NMR (6029 Hz): 5 -9.9 (40.,2 = 1649 Hz). Anal. Calcd for C6H19NO3BNa: C, 38.50; H, 10.16; N, 7.49. Found: C, 38.09; H, 10.47; N, 7.39; H' content: 4.11. NaNCBH3-TEA (46). A mixture of 0.63 g (0.01 mol) of NaCNBH; and 1.4 mL (0.0] mol) of triethanolamine in 30 mL isopropanol was stirred at rt for 2 h. The undissolved material was filtered out and the isopropanol was slowly evaporated from the filtrate using a gentle stream of nitrogen. When it was almost dry, 50 mL of CH2Cl2 were added and the white solid was collected. Yield 1.1 g (52%). X-Ray quality crystals were grown by slow evaporation of a solution of 46 in isopropanol. Mp = 85-86 °C; IR (KBr): w... = 2341 cm“, vCN = 2180 cm"; 'H NMR (CH3CN-d3): 5 3.55 (t, J = 5.4 Hz, 6H, OCH2), 3.46 (s, 3H, OH), 2.52 (t, J = 5.4 Hz, 6H, NCH2), 0.26 (q, 13.. = 88.6, 3H, NCBH3); “B NMR (DMSO-d6): 5 -559 (q, 1.... = 88.6, NCBH3); 13C NMR (DMSO-d6): 5 59.0 (OCH2), 56.9 (NCH2). Anal. Calcd for C7H18N2O3BNa: C, 39.62; H, 8.49; N, 13.21. Found: C, 38.73; H, 9.23; N, 13.21. LiBH4-TEA (47). 0.22 g (10 mmol) LiBH4 and 1.4 mL (10 mmol) TEA were dissolved in 20 mL 2-propanol. Slow evaporation of most of the solvent under Ar followed by addition of 20 mL CH2C12 afforded 0.95 g (56 %) complex. X-Ray quality crystals were grown by diffusion of a diethyl ether layer into a solution of 47 in 2- propanol. 1H NMR (CH3CN-d3): 5 = 4.21 (s, 3H; OH), 3.60 (t, J = 5.1 Hz, 6H; OCH2), 2.58 (t, J = 5.1 Hz, 6H; NCH2), -O.23 (q, JBH = 81.0 Hz, 4H; BH4); 13C NMR (CH3CN- d3): 5 53.5 (NCH2), 58.4 (OCH2); ”B NMR (CH3CN-d3): 5 -55.98 (quintet, 1.... = 80.9 Hz; B114); "8 MAS NMR (5997 Hz): 5 -48.7 (Aw/2 = 1586 Hz); 7L1 MAS NMR (6000 136 Hz): 5 2.8 (411.,2 = 1735 Hz); IR (KBr): w... = 2290 cm"; Anal. Calcd for C6H1903NBLi: C, 42.15; H, 11.11; N, 8.19. Found: C, 42.39; H, 11.56; N, 8.17. LiBH4-TEAod3. A mixture of 0.22 g (10 mmol) of LiBHa and 1.4 mL ( 10.3 mmol),of TEA-d3 in 30 mL THF was stirred under Ar at rt for 1 h. The resulting precipitate was filtered and washed with THF. Yield quantitative. 1H NMR (CH3CN-d3): no OH signal; IR (KBr): V01) = 2516 cm"; Anal. hydridic content: 4.06. LiBDa-T EA (47a). A mixture of 0.065 g (2.5 mmol) of LiBDa and 0.4 g (2.7 mmol) TEA in 8 mL THF was stirred under Ar at rt for 2h. The resulting precipitate was filtered and washed with THF. IR (KBr): V31) = 1704, 1652 cm"; Anal. hydridic content: 3.90, 90% D based on 'H NMR (DMSO-d6). LiBI-Ia (50% D)-TEA (47b). A mixture of 0.044 g (2 mmol) of LiBHa and 0.052 g (2 mmol) of LiBDa was dissolved in 12 mL THF. An amount of 0.64 g (4.3 mmol) of TEA was subsequently added and the mixture was stirred under Ar at rt for 2h. The resulting precipitate was collected and washed with THF. IR (KBr): v31) = 1752, 1703, 1649 cm", w... = 2384, 2292, 2226 cm"; Anal. hydridic content: 3.80, 54% D based on ‘H NMR (DMSO-d6). NaBHa-THEC (49). 0.038 g (1 mmol) of NaBHa were dissolved in 10 mL 2- propanol and an amount of 0.366 g (1 mmol) of THEC-H2O was subsequently added. After approximately 1 min a white precipitate was formed. The mixture was stirred for 1 h at rt under an Ar atmosphere. Filtration of the mixture yielded 0.3 g (78%) of 49. Single crystals suitable for X-ray crystallography were grown by diffusion of Et2O in a solution of 49 in CH3CN. Mp = 180-181 °C (dec.); IR (KBr): var. = 2298, 2225 cm"; 1H NMR (pyridine-d5): 5 6.03 (br s, 4H, OH), 3.85 (t, J = 4.5 Hz, 8H, OCH2), 2.49-2.41 (br, 24H, 137 NCH2), 1.51 (q, 13.. = 81.3 Hz, 4H, B114); 13C NMR (CH3CN-d3): 5 51.17 (NCH2), 56.45 (NCH2), 58.96 (OCH2); ”B NMR (CH3CN-d3): 5 -54.21 (quintet, 1.3.. = 81.4 Hz, B114); “13 MAS NMR (4013 Hz): 5 - 45.4 (Aw/2: 2198 Hz); 23Na MAS NMR (6028 Hz): 5 -12.3 (Av1/2 = 1440 Hz). Anal. Calcd for C16H40N404BNa: C, 49.74; H, 10.36; N, 14.51. Found: C, 49.37; H, 10.87; N, 14.80. NaNCBH3-THEC (50). A solution of 0.032 g (0.5 mmol) of NaNCBH3 in 3 mL THF was added to 0.183 g (0.5 mmol) THEC-H2O in THF. A white precipitate was formed immediately. The mixture was stired 1 h at rt under an Ar atmosphere. The solid was subsequently filtered and washed with THF to yield 0.17 g (83%) of 50. Single crystals suitable for X-ray crystallography were grown by diffusion of Et2O in a solution of 50in CH3CN. Mp = 230-234 °C (dec.); IR (KBr): v3" = 2348 cm'l, VCN = 2162 cm"; lH NMR (CH3CN-d3): 5 3.61 (t, J = 4.8 Hz, 8H, OCH2), 3.25 (br s, 4H, OH), 2.47 (br, 24H, NCH2), 0.25 (q, 3H, NCBH3); ”C NMR (CH3CN-d3): 5 51.17 (NCH2), 56.50 (NCH2), 59.06 (OCH2); llB NMR (CH3CN-d3): 5 -56.60 (q, 13.. = 88.3 Hz, NCBH3). NaBHa-THPC (51). 0.285 g (0.7 mmol) of THPC were added to a solution of 0.027 g (0.7 mmol) in 8 ml 2-propanol. In approximately 2 min a white precipitate was formed. The mixture was left standing for l h at rt, then the solid was filtered. Yield 0.085 g (27%). No melting occurs when the solid is heated up to 200 °C, at which temperature it turns brown, and eventually starts to melt at approximately 210 °C. Single crystals suitable for X-ray crystallography were grown by slow evaporation from 2- propanol. IR (KBr): VB" = 2290 cm'l; 1H NMR (H20-d2): 5 3.89 (br m, 4H, OCH), 2.90- 2.87 (br (1, 8H, NCH2), 2.42-2.34 (br t, 4H, NCH2), 2.09 (br s, 12 H, NCH2) -0.27 (q, 1311 = 80.72 Hz, 4H, BH4); 13C NMR (DMSO-d5): 5 21. 79 (CH3), 49.17 (br, NCH2), 51.03 138 (br, NCH2), 61.58 (NCH2), 62.31 (OCH2); “8 NMR (DMSO-d6): 5 -52.55 (quintet, 1.3.. = 82.8 Hz, BH4); Anal. Calcd for C20H43N4OaBNa: C, 54.30; H, 10.86; N, 12.67. Found: C, 54.66; H, 11.39; N, 12.41. 9-(1’,4’,7’-tris-hydroxyethyl-cyclen)methylanthracene (52). Warning: Ethylene oxide is a volatile (Bp = 11 °C), extremely toxic, carcinogenic substance. 0.2 g (0.57 mmol) of 9-(1’,4’,7’,10’~tetraazacyclododecyl)methylanthracene were dissolved in 3 mL absolute ethanol. 2 mL of water were added, and the solution was cooled to 0 °C. To this solution, 0.158 g (3.6 mmol) of ethylene oxide (cooled to 0 °C) were added, and the mixture was stirred at 0 °C for 2 h, and then another hour at rt. Solvent evaporation under vacuum at 35 °C yielded a yellow oil. A few mL of water were subsequently added, and the resulting aqueous solution was extracted with chloroform. Evaporation of CHC13 yielded a yellow oil, which was dissolved in minimal amount of 2-propanol, and a large excess of ether was added. Upon standing in the freezer over night, a yellow crystalline solid precipitated, which was filtered and washed with ether, to yield 0.12 g (42%) of 52. Single crystals suitable for X-ray crystallography were grown by diffusion of Et2O in a solution of 52 in 2-propanol. Mp = 131-133 °C; 1H NMR (CH3Cl—d3): 5 8.36 (d, 2H, ArH), 8.29 (s, 1H, ArH), 7.87 (d, 2H, ArH), 7.38 (m, 4H, ArH), 4.8 (br, OH), 4.39 (s, 2H, CH2-Ar), 3.47 (t, J = 4.5 Hz, 2H, OCH2), 3.40 (t, J = 4.5 Hz, 4H, OCH2), 2.73 (t, J = 4.5 Hz, 6H, NCH2), 2.43-2.32 (br, 16H, NCH2); 13C NMR (CH3C1-d3): 5 130.83, 128.51, 127.11, 125.32, 124.50, 58.35, 58.01, 57.73, 55.65, 52.37, 52.11, 51.90, 50.88, 50.56; Anal. Calcd for C29H42N403: C, 70.45; H, 8.50; N, 11.34. Found: C, 70.23; H, 8.83; N, 11.04. 139 NaBHa-[9-(1’,4’,7’-tris-hydroxyethyl-cyc1en)methylanthracene]~iPrOH (53). 0.1 g (0.2 mmol) of 9-(1’,4’,7’-tris-hydroxyethyl-cyc1en)methylanthracene were added to a solution of 0.008 g (0.21 mmol) of N aBH4 in 4 mL 2-propanol. In approximately 4 min a white precipitate was formed. The mixture was stirred 2 h at rt under an Ar atmosphere, and subsequently filtered to yield 0.09 g (53%) of 53. The solid turns red when heated, and partly melts at 190 °C. Single crystals suitable for X-ray crystallography were grown by diffusion of Et2O in a solution of 53 in CH3CN. IR (KBr): V311 = 2384, 2293, 2228 cm"; 'H NMR (CH3CN-d3): 5 8.54 (s,1H, ArH), 8.43 (d, 2H, ArH), 8.06 (d, 2H, ArH), 7.60 (m, 2H, ArH), 7.49 (m, 2H, ArH), 4.61 (s, 2H, CH2-Ar), 3.72 (septet, 1H, iPrOH), 3.6-1.9 (br), 1.08 (d, 6H, iPrOH); 13C NMR (CH3CN-d3): 5 132.47, 130.23, 128.98, 127.77, 126.22, 125.46, 58.54, 57.98, 56.54, 56.15, 54.51, 52.65, 51.59, 51.26, 50.74; HB NMR (CH3CN-d3): 5 -53.75 (quintet, JBH = 80.4 Hz, BH4); Anal. Calcd for C32H54N4O4NaB: C, 64.87; H, 9.12; N, 9.46. Found: C, 64.20; H, 9.06; N, 9.33. Decomposition Studies: Solid-State Decomposition of NaBH4-TEA. 05-07 g of N aB114~TEA were heated at 82 °C under Ar or open atmosphere in a 7 x 27 mm glass tube immersed in refluxing acetonitrile, for 30-58 days. The resulting white solid is insoluble in common organic solvents and it does not melt up to 300 °C. The powder XRD shows a single phase with a broad peak at 20 = 94° but no NaBHa or initial complex. IR (KBr): VBH = 2387, 2293, 2226 cm"; ”8 MAS NMR (6058 Hz): 5 -7.1 (Aw/2 = 1390 Hz); 23Na MAS 140 NMR (6004 Hz): 5 -14.6 (Aw/2 = 2769 Hz). Anal. NIB/Na = 1/ 1/0.96; H' content: 1.05- 1.38. Decomposition of NaBHa-TEA in DMSO. A solution of 0.5 g of NaBH4-TEA in 10 mL DMSO was heated at 110 °C under Ar for 55 h. The resulting precipitate was collected and washed with CH2Cl2 to yield 0.17 g of white solid insoluble in common organic solvents. Mp > 300 °C; IR (KBr): no B-H stretching vibration; llB MAS NMR (6000 Hz): 5 -5.0 (Av.,.= 1559 Hz); ”Na MAS NMR (6011 Hz): 5 .135 (Av... = 2665 Hz). Anal. H' content: 0.22. Decomposition of NaBH4-TEA in the Melt. 0.5 g of NaBH4-TEA were placed in a 21 x 50 mm vial and heated under Ar at 130 °C in an oil bath for two hours. In about 1 min the solid melted with effervescence. After approximately another 10 min the melt solidified. The powder XRD confirmed the presence of NaBH4. llB MAS NMR (5998 Hz): 5 -2.3 (Aw/2: 1394 Hz, 1 B), 5 -50.4 (Aw/2: 1017 Hz 1.2 B). Anal. H' content: 2.03. Solid-State Decomposition of LiBIL-TEA. Aliquots of 0.1-0.5 g of LiBHa-TEA were placed in a 21 x 50 mm vial and were preheated under Ar at 80-100 °C for 10-15 min. The temperature was subsequently raised to 120-130 °C, and heating was continued for about 2 h. The resulting white solid is insoluble in common organic solvents and it does not melt up to 300 °C. IR (KBr): v8.. = 2301, 2247, 2171 cm"; “13 MAS NMR (6000 Hz): 5 -4.6 (Av.,. = 1693 Hz); 7L1 MAS NMR (6000 Hz): 5 3.1 (Av.,. = 1611 Hz). Anal. H’ content: 0.97-1.08. 141 Annealing of the LiBH4-TEA Solid-State Decomposition Product. A suspension of 0.07 g of the solid-state decomposition product 48 (H' = 1.05) in 5 ml. of freshly distilled DMSO was heated at 120 °C under an Ar atmosphere for 4 days, then filtered and the resulting solid was washed with CH2Cl2. IR (KBr): v3" = 2368, 2340 cm"; ”13 MAS NMR (6000 Hz): 5 -33 (Av.,. = 1677 Hz); 71.1 MAS NMR (3540 Hz): 5 1.0 (Aw/2 = 2496 Hz). Anal. H' content: 1.05. Decomposition of LiBHa-TEA in DMSO. A solution of 0.5 g of LiBH4-TEA in 10 mL DMSO was heated at 120 °C under Ar for 4 days. The resulting precipitate was collected and washed with CH2Cl2. Yield 0.2 g. Mp > 300 °C; IR (KBr): no EH; "8 MAS NMR (6000 Hz): 5 -5.9 (Aw/2 = 1726 Hz). Anal. H' content: 0.26. HID Isotope Exchange Experiments. Aliquots of 0.1-0.3 g of LiBHa-TEA-d3 were heated under Ar at 85-110 °C, monitoring the decompositions by H' analysis. The partly or fully decomposed samples were analyzed by IR (KBr) and 1H NMR (CH3CN- d3). No signals for the V131) in the IR or for OH in the 1H N MR spectra were observed throughout the reaction. Reaction of the LiBIL-TEA Solid-State Decomposition Product (48) with Pd(AcO)2. A sample of 0.1 g (z 0.61 miliequivalents of H’) of the solid-state decomposition product 48 was suspended in 10 mL freshly distilled THF containing 0.136 g (0.61 mmol) Pd(AcO)2. The mixture was stirred at rt under an Ar atmosphere for 24 h. After filtering and washing with THF, a green hygroscopic powder was obtained, which turned into a black paste if left in open atmosphere. Powder XRD analysis Showed that this compound was completely amorphous, and no Pd black, or unconverted Pd(AcO)2 was present. IR (KBr): vAco = 1580 cm"; no B-H; ‘H NMR (DMSO-d6): very 142 complex, and indicates the presence of at least three different AcO groups: 5 1.72, 1.69, 1.63. '3C NMR (DMSO-d6): 5 22.94, 24.02, 25.17, 59.20, 61.64, 64.43, 66.26, 67.08; “B NMR (DMSO-d6): 5 -14.61. Solid-State Decomposition of NaBHa-THEC. A sample of 0.07 g of NaBI-I4-THEC was heated at 160 °C under an Ar atmosphere for 20 h. The resulting white solid is insoluble in common organic solvents and it does not melt up to 300 °C. IR (KBr): no B-H; “13 MAS NMR (4038 Hz): 5 80 (Av.,. = 2462 Hz); ”Na MAS NMR (6001 Hz): 5 -15.6 (Aw/2: 3421 Hz). Decomposition of NaBHa-THEC in the Melt. A sample of 0.04 g of NaBHa-THEC was rapidly heated to 200 °C under Ar. Melting occurred under these conditions, followed by immediate solidification. Heating was continued for 10 min. The IR spectrum of the resulting solid is virtually identical with the corresponding spectrum of the solid-state decomposition product. Powder XRD indicates an amorphous phase, and annealing at 160 °C for 24 h did not improve its crystallinity. llB MAS NMR (6000 Hz): 5 -6.9 (Aw/2 = 1745 Hz); ”Na MAS NMR (6000 Hz): 5 -8.9 (Av.,. = 2704 Hz). Decomposition of NaBHa-THEC in DMSO. A solution of 0.15 g of N aBHa-THEC in 5 mL DMSO was heated at 120 °C under Ar for 48 h. A crystalline solid precipitated, which was filtered and washed with CH2Cl2. Yield 0.13 g. “B MAs NMR (6060 Hz): 5 -7.5 (Aw/2 = 1880 Hz); 23Na MAS NMR (6058 Hz): 5 -9.0 (Aw/2 = 2476 Hz). 143 BIBLIOGRAPHY 144 BIBLYOGRAPHY (1) Jeffrey, G. A. An Introduction to Hydrogen Bonding, Oxford University Press, Oxford, 1997. (2) Jeffrey, G. A.; Saenger, W. Hydrogen Bonding in Biological Structures, Springer- Verlag, Berlin, 1991. (3) (a) Crabtree, R. H.; Siegbahn, P. E. M.; Eisenstein, O.; Rheingold, A. L.; Koetzle, T. F. Acc. Chem. Res. 1996, 29, 348. (b) Crabtree, R. H. J. Organomet. Chem. 1998, 577, 111. (c) Shubina, E. S.; Belkova, N. V.; Epstein, L. M. J. Organomet. Chem. 1997, 536- 537, 17. (d) Alkorta, 1.; Rozas, I.; Elguero, J. Chem. Soc. Rev. 1998, 27, 163. (e) Crabtree, R. H. Science 1998, 282, 2000. (f) Crabtree, R. H.; Eisenstein, O.; Sini, G.; Peris, E. J. Organomet. Chem. 1998, 567, 7. (g) Kelly, R; Loza, M. Chemistry in Britain 1999, 35, 26. (h) Calhorda, M. J. Chem. Commun. 2000, 801 (4) Zachariasen, W. H.; Mooney, R. C. L. J. Chem. Phys. 1934, 2, 34. (5) We have recently redetermined the crystal structure of ammonium hypophosphite at higher resolution, which allowed us to locate the H atoms and thus probe the existence of dihydrogen bonding. However, we found no close H-r-H contacts in this salt. Stein, R.; Custelcean, R.; Jackson, J. E. Manuscript in preparation. (6) Burg, A. B. Inorg. Chem. 1964, 3, 1325. (7) Titov, L. V.; Makarova, M. D.; Rosolovskii, V. Y. Doklady Akademii Nauk 1968, 180, 381. 145 (8) (a) Brown, M. P.; Heseltine, R. W. Chem. Commun. 1968, 1551. (b) Brown, M. P.; Heseltine, R. W.; Smith, P. A.; Walker, P. J. J. Chem. Soc. A 1970, 410. (9) Brown, M. P.; Walker, P. J. Spectrochim. Acta 1974, 30A, 1125. (10) (a) Shubina E. S.; Bachmutova, E. V.; Saitkulova, L. N.; Epstein, L. M. Mendeleev Commun. 1997, 83. (b) Epstein, L. M.; Shubina, E. S.; Bakhmutova, E. V.; Saitkulova, L. N.; Bakhmutov, V. 1.; Chistyakov, A. L.; Stankevich, I. V. Inorg. Chem. 1998, 37, 3013. (c) Shubina, E. S.; Belkova, N. V.; Bakhmutova, E. V.; Saitkulova, L. N.; Ionidis, A. V.; Epstein, L. M. Russ. Chem. Bull. 1998, 4 7, 817. (11) Jackson, J. E.; Huang, R.; Eckert, J.; Sharma, M.; Argyriou, D.; Sheldon, R. Unpublished results. (12) Peris, E.; Lee, J. C.; Rambo, J. R.; Eisenstein, O.; Crabtree, R. H. J. Am. Chem. Soc. 1995, 117, 3485. (13) Cramer, C. J.; Gladfelter, W. L. Inorg. Chem. 1997, 36, 5358. (14) Klooster, W. T.; Koetzle, T. F.; Siegbahn, Per E. M.; Richardson, T. B.; Crabtree, R. H. J. Am. Chem. Soc. 1999, 121, 6337. (15) Popelier, P. L. A. J. Phys. Chem. A 1998, 102, 1873. (16) N5th, H.; Thomas, S. Eur. J. Inorg. Chem. 1999, 1373. (17) (a) Zottola, M. A.; Pedersen, L. G.; Singh, P.; Shaw, B. R. Modeling the Hydrogen Bond; ACS Symposium Series, Smith, D. A., Ed.; American Chemical Society, Washington, DC, 1994, Vol. 569, pp. 277-285 (b) Singh, P.; Zottola, M.; Huang, 8.; Shaw, B. R.; Pedersen L. G. Acta Cryst. 1996, C52, 693. (18) Padilla-Martinez, I. R.; Rosalez-Hoz, M. de J.; Tlahuext, H.; Camacho-Camacho, C.; Ariza-Castolo, A.; Contreras, R. Chem. Ber. 1996, 129, 441. 146 (19) Flores-Pana, A.; Sénchez—Ruiz, S. A.; Guadarrarna, C.; N5th, H.; Contreras, R. Eur. J. Inorg. Chem. 1999, 2069. (20) Atwood, J. L.; Koutsantonis, G. A.; Lee, F.-C.; Raston, C. L. Chem. Commun. 1994, 91. (21) Campbell, J. P.; Hwang, J. W.; Young, V. G.; Von Dreele, R. B.; Cramer, C. J.; Gladfelter, W. L. J. Am. Chem. Soc. 1998, 120, 521. (22) (a) Liu, Q.; Hoffmann, R. J. Am. Chem. Soc. 1995, 117, 10108. (b) Alkorta, 1.; Elguero, J.; Foces-Foces, C. Chem. Commun. 1996, 1633. (c) Remko, M. Molecular Physics 1998, 94, 839. (d) Kulkami, S. A. J. Phys. Chem. A 1998, 102, 7704. (e) Kulkami, S. A.; Srivastava, A. K. J. Phys. Chem. A 1999, 103, 2836. (f) Petterson, M.; Lundell, J. Phys. Chem. Chem. Phys. 1999, I, 1601. (23) Grabowski, S. J. Chem. Phys. Lett. 1999, 312 (5-6), 542. (24) Milstein, D.; Calabrese, J. C.; Williams, I. D. J. Am. Chem. Soc. 1986, 108, 6387. (25) Stevens, R. C.; Bau, R.; Milstein, D.; Blum, O.; Koetzle, T. F. J. Chem. Soc. Dalton 77ans.1990,l429. (26) Feracin, S.; Biirgi, T.; Bakhmutov, V. I.; Eremenko, 1.; Vorontsov, E. V.; Vimenits, A. B.; Berke, H. Organometallics 1994, I3, 4194. (27) Lee, J. C.; Rheingold, A. L.; Muller, B.; Pregosin, P. S.; Crabtree, R. H. Chem. Cbnunun.l994,1021. (28) Peris, E.; Lee, J. C.; Rambo, J. R., Eisenstein, O.; Crabtree, R. H. J. Am. Chem. Soc. 1995,117,3485. (29) Lough, A. J .; Park, S.; Ramachandran, R.; Morris, R. H. J. Am. Chem. Soc. 1994, 116, 8356. 147 (30) Park, S.; Ramachandran, R.; Lough, A. J.; Morris, R. H. Chem. Commun. 1994, 2201. (31) Park, S.; Lough, A. J.; Morris, R. H. Inorg. Chem. 1996, 35, 3001. (32) (a) Wessel, J.; Lee, J. C.; Peris, E.; Yap, G. P. A.; Fortin, J. B.; Ricci, J. S.; Sini, G.; Albinati, A.; Koetzle, T. F.; Eisenstein, O.; Rheingold, A. L.; Crabtree, R. H. Angew. Chem. Int. Ed. 1995, 34, 2507. (b) Patel, B. P.; Wessel, J .; Yao, W.; Lee Jr., J. C.; Peris, E.; Koetzle, T. F.; Yap, G. P. A.; Fortin, J. B.; Ricci, J. S.; Sini, G.; Albinati, A.; Eisenstein, O.; Rheingold, A. L.; Crabtree, R. H. New J. Chem. 1997, 21 , 413. (c) Sini, G.; Eisenstein, O.; Yao, W.; Crabtree, R. H. Inorg. Chim. Acta 1998, 280, 26. (33) Patel, B. P.; Yao, W.; Yap, G. P.A.; Rheingold, A. L.; Crabtree, R. H. Chem. Commun. 1996, 991. (34) Peris, E.; Wessel, J.; Patel, B. P.; Crabtree, R. H. Chem. Commun. 1995, 2175. (35) Desmurs, P.; Kavallieratos, K.; Yao, W.; Crabtree, R. H. New J. Chem. 1999, 23, 1111. (36) Shubina, E. S.; Belkova, N. V.; Krylov A. N .; Vorontsov, E. V.; Epstein, L. M.; Gusev, D.G.; Niederrnann, M.; Berke, H. J. Am. Chem. Soc. 1996, 118, 1105. (37) Belkova, N. V.; Shubina, E. S.; Ionidis, A. V.; Epstein, L. M.; Jacobsen, H.; Messmer, A.; Berke, H. Inorg. Chem. 1997, 36, 1522. (38) Messmer, A.; Jacobsen, H.; Berke, H. Chem. Eur. J. 1999, 5, 3341. (39) Scheiner, S.; Orlova, G. J. Phys. Chem. A 1998, 102, 260. (40) (a) Richardson, T. B.; Koetzle, T. F.; Crabtree, R. H. Inorg. Chim. Acta 1996, 250, 69. (b) Xu, W.; Lough, A. J.; Morris, R. H. Can. J. Chem. 1997, 75, 475. (c) Huang, L.- 148 Y.; Aulwurrn, U. R.; Heinemann, F. W.; Knoch, F.; Kisch, H. Chem. Eur. J. 1998, 4, 1641. (d) Junk, P. C.; Steed, J. W. J. Organomet. Chem. 1999, 587, 191. (41) Braga, D.; De Leonardis, P.; Grepioni, F.; Tedesco, E.; Calhorda, M. J. Inorg. Chem. 1998, 37, 3337. (42) Abramov, Y. A.; Brammer, L.; Klooster, W. T.; Bullock, R. M. Inorg. Chem. 1998, 37, 6317. (43) Gusev, D. G.; Lough, A. J .; Mom's, R. H. J. Am. Chem. Soc. 1998, 120, 13138. (44) Abdur-Rashid, K.; Gusev, D. G.; Landau, S. E.; Lough, A. J.; Morris, R. H. J. Am. Chem. Soc. 1998, 120, 11826. (45) Abdur-Rashid, K.; Gusev, D. G.; Lough, A. J .; Morris, R. H. Organometallics 2000, 19, 834. (46) Jackson, J. E.; Bass, S.; Huang, R. Unpublished results. (47) Lee, J. C.; Peris, E.; Rheingold, A. L.; Crabtree, R. H. J. Am. Chem. Soc. 1994, 116, 11014. (48) Ayllon, J. A.; Gervaux, C.; Sabo-Etienne, S.; Chaudret, B. Organometallics 1997, 16, 2000. (49) Orlova, G.; Scheiner, S.; Kar, T. J. Phys. Chem. A 1999, 103, 514. (50) Shubina, E. S.; Belkova, N. V.; Bakhmutova, E. V.; Vorontsov, E. V.; Bakhmutov, V. 1.; Ionidis, A. V.; Bianchini, C.; Marvelli, L.; Peruzzini, M.; Epstein, L. M. Inorg. Chim. Acta 1998, 280, 302. (51) Epstein, L. M.; Shubina, E. S. Ber. Bunsenges. Phys. Chem. 1998, 102, 359. (52) Orlova, G.; Scheiner, S. J. Phys. Chem. A 1998, 102, 4813. (53) Chu, H. S.; Xu, 2.; Ng, S. M.; Lau, C. P.; Lin, Z. Eur. J. Inorg. Chem. 2000, 993. 149 (54) Ayllon, J. A.; Sabo-Etienne, S.; Chaudret, B.; Ulrich, S.; Limbach, H.-H. Inorg. Chim. Acta 1997, 259, 1. (55) Grfindemann, S.; Ulrich, S.; Limbach, H.-H.; Golubev, N. S.; Denisov, G. S.; Epstein, L. M.; Sabo-Etienne, S.; Chaudret, B. Inorg. Chem. 1999, 38, 2550. (56) Chu, H. S.; Lau, C. P.; Wong, K. Y.; Wong, W. T. Organometallics 1998, 17, 2768. (57) Musashi, Y.; Sakaki, S. J. Am. Chem. Soc. 2000, 122, 3867. (58) Ayllon, J. A.; Sayers, S. F.; Sabo-Etienne, S.; Donnadieu, B.; Chaudret, B. Organometallics 1999, 18, 3981. (59) Caballero, A.; Jalén, F. A.; Manzano, B. R. Chem. Commun. 1998, 1879. (60) Patel, B. P.; Kavallieratos, K.; Crabtree, R. H. J. Organomet. Chem. 1997, 528, 205. (61) Bosque, R.; Maseras, F.; Eisenstein, O.; Patel, B. P.; Yao, W.; Crabtree, R. H. Inorg. Chem. 1997, 36, 5505. (62) Yao, W.; Crabtree, R. H. Inorg. Chem. 1996, 35, 3007. (63) Aime, S.; Gobetto, R.; Valls, E. Organometallics 1997, 16, 5140. (64) Aime, S.; Fém'z, M.; Gobetto, R.; Valls, E. Organometallics 1999, 18, 2030. (65) Aime, S.; Férriz, M.; Gobetto, R.; Valls, E. Organometallics 2000, 19, 707. (66) Haack, K.-J.; Hashiguchi, S.; Fujii, A.; Ikariya, T.; Noyori, R. Angew. Chem. Int. Ed. 1997, 36, 285. (67) Gatling, S. C.; Jackson, J. E. J. Am. Chem. Soc. 1999, 121, 8655. (68) Hwang, J .-W.; Campbell, J. P.; Kozubowski, J.; Hanson, S. A.; Evans, J. F.; Gladfelter, W. L. Chem. Mater. 1995, 7, 517. (69) W5hler, F. Pogg. Ann. 1828, 12, 253. 150 (70) (a) v Liebig, J .; W5hler, F. Ann. Phys. Leipzig, Ser. 2 1830, 20, 369; (b) Dunitz, J. D.; Harris, K. D. M.; Johnston, R. L.; Kariuki, B. M.; MacLean, E. J .; Psallidas, K.; Schweizer, W. B.; Tykwinski, R. R J. Am. Chem. Soc. 1998, 120, 13274. (71) (a) Ohashi, Y. Reactivity in Molecular Crystals, Kodansha, Tokyo, 1993. (b) Singh, N. B.; Singh, R. J .; Singh, N. P. Tetrahedron 1994, 50, 6441. (72) Schmidt, G. M. J. Pure Appl. Chem. 1971, 27, 647. (73) Cohen, M. D. Angew. Chem. Int. Ed. 1975, 14, 386. (74) Gavezzotti, A. J. Am. Chem. Soc. 1983, 105, 5220. (75) McBride, J. M.; Segmuller, B. E.; Hollingsworth, M. D.; Mills, D. E.; Weber, B. A. Science 1986, 234, 830. (76) Luty, T.; Eckhartd, C. J. J. Am. Chem. Soc. 1995, 117, 2441. (77) Curtin, D. Y.; Paul, C. I.; Duesler, E. N.; Lewis, T. W.; Mann, B. J .; Shiau, W. 1. Mol. Cryst. Liq. Cryst. 1979, 50, 25. (78) Gougoutas, J. Z. Pure Appl. Chem. 1971, 27, 305. (79) Wegner, G. Pure Appl. Chem. 1977, 49, 443. (80) Adler, G. Mol. Cryst. Liq. Cryst. 1988, 156, l. (81) Hirshfeld, F. L.; Schmidt, G. M. J. J. Polymer Sci. A 1964 , 2, 2181. (82) Hasegawa, M. Chem. Rev. 1983, 83, 508. (83) Addadi, L.; Lahav, M. J. Am. Chem. Soc. 1979, 101, 2152. (84) Epple, M.; Kirschnick, H. Chem. Ber. 1996, 129, 1123. (85) Epple, M.; Kirschnick, H. Chem. Ber. 1997, I30, 291. (86) Leiserowitz, L.; Schmidt, G. M. J. J. Chem. Soc. A, 1969, 2372. (87) Feldman, K. S.; Campbell, R. F. J. Org. Chem. 1995, 60, 1924. 151 (88) Feldman, K. S.; Campbell, R. F.; Saunders, J. C.; Ahn, C.; Masters, K. M. J. Org. Chem. 1997, 62, 8814. (89) Matsumoto, A.; Odani, T.; Chikada, M.; Sada, K.; Miyata, M. J. Am. Chem. Soc. 1999. 121, 11122. (90) Kane, J. J .; Liao, R.-F.; Lauher, J. W.; Fowler, F. W. J. Am. Chem. Soc. 1995, 117, 12003. (91) Xiao, J .; Yang, M.; Lauher, J. W.; Fowler, F. W. Angew. Chem. Int. Ed. 2000, 39, 2132. (92) Coates, G. W.; Dunn, A. R.; Henling, L. M.; Dougherty, D. A.; Grubbs, R. H. Angew. Chem. Int. Ed. 1997, 36, 248. (93) Coates, G. W.; Dunn, A. R.; Henling, L. M.; Ziller, J. W.; Lobkovsky, E. B.; Grubbs, R. H. J. Am. Chem. Soc. 1998, 120, 3641. (94) For excellent reviews about kinetics in the solid-state see: (a) Hannay, N. B. Treatise 0n Solid State Chemistry 1976, Reactivity of Solids, Vol 4, Plenum Press, New York. (b) Brown, M. E.; Dollimore, D.; Galwey, A. K. Bamford, C. H. (Editor) Comprehensive Chemical Kinetics, Vol. 22: Reactions in the Solid State 1980, Elsevier Scientific Publishing Company, Amsterdam-Oxford-New York. (c) Garner, W. E. (Editor) Chemistry of the Solid State 1955, Academic Press Inc., New York. (95) (a) Sekiguchi, K.; Shirotani, K—I.; Sakata, 0.; Suzuki, E. Chem. Pharm. Bull. 1984, 32, 1558. (b) Brill, T. B.; Gongwer, P. E.; Williams, G. K. J. Phys. Chem. 1994, 98, 12242. (96) Baughman, R. H. J. Chem. Phys. 1978, 68, 3110. (97) Galwey, A. K.; Brown, M. E. Proc. R. Soc. London A 1995, 450, 501. 152 (98) (a) Vyazovkin, S.; Wight, C. A. Annu. Rev. Phys. Chem. 1997, 48, 125. (b) Vyazovkin, S.; Wight, C. A. J. Phys. Chem. A 1997, 101, 8279. (99) (a) Davis, R. E. J. Am. Chem. Soc. 1962, 84, 892. (b) Pepperberg, I. M.; Halgren, T. A.; Lipscomb, W. N. J. Am. Chem. Soc. 1976, 98, 3442. (100) Mesmer, R. E.; Jolly, W. L. Inorg. Chem. 1962, 1, 608. (101) (a) Schreiner, P. R.; Schaefer, H. F.; Schleyer, P. v. R. J. Chem. Phys. 1994, 101, 7625. (b) Watts, J. D.; Bartlett, R. J. J. Am. Chem. Soc. 1995, 117, 825. (c) Custelcean, R. ow‘a J. Mol. Struct. Theochem 2000, 505, 95. (102) Rozas, I.; Alkorta, 1.; Elguero, J. Chem. Phys. Lett. 1997, 275, 423. (103) Custelcean, R.; Jackson, J. E. J. Am. Chem. Soc. 1998, 120, 12935. (104) Naiini, A. A.; Pinkas, J .; Plass, W.; Young, V. G.; Verkade, J. G. Inorg. Chem. 1994, 33, 2137. (105) a) Ashby, E. C.; Dobbs, F. R.; Hopkins, H. P.; J. Am. Chem. Soc. 1975, 97, 3158. b) Kadlec, V.; Hanzlik, J. Collect. Czech. Chem. Commun. 1974, 39, 3200. (106) Davis, R. E.; Gottbrath, J. A.; J. Am. Chem Soc. 1962, 84, 895. (107) (8) Davis, R. E.; Bromels, E.; Kibby, C. L. J. Am. Chem. Soc. 1962, 84, 885; (b) Kreevoy, M. M.; Hutchins, J. E. C. J. Am. Chem. Soc., 1969, 91, 4329. (108) Custelcean, R.; Jackson, J. E. Angew. Chem. Int. Ed. 1999, 38, 1661. (109) Brown, H. C.; Mead, E. J .; Scoaf, C. J. J. Am. Chem. Soc. 1956, 78, 3616. (110) (a) Onak, T. P.; Landesman, H.; Williams, R. E.; Shapiro, I. J. Phys. Chem. 1959, 63, 1533. (b) Golden, J. H.; Schreier, C.; Singaram, B.; Williamson, 8. M. Inorg. Chem. 1992, 31, 1533. 153 (111) (a) Price, W. C. J. Chem. Phys. 1949, 17, 1044. (b) Volkov, V. V.; Sobolev, E. V.; Grankina, Z. A.; Kalinina, I. S. Russ. J. Inorg. Chem. 1968, 13, 343. (c) Marks, T. J.; Kennelly, W. J .; Kolb, J. R.; Shimp, L. A. Inorg. Chem. 1972, 11, 2540. (112) Custelcean, R.; Jackson, J. E. J. Am. Chem. Soc. 2000, 122, 5251. (1 13) (a) Galwey, A. K. Pure Appl. Chem. 1995, 67, 1809. (b) Tanaka, H.; Koga, N. J. Phys. Chem. 1988, 92, 7023. (c) Koga, N.; Tanaka, H. J. Phys. Chem. 1994, 98, 10521. (114) For examples of other solid-state reactions where the kinetics are directly coupled to chemical transformations see: (a) Galwey, A. K.; Mohamed, M. A. J. Chem. Soc. Faraday Trans. 1985, 81, 2503. (b) Son, S. F.; Asay, B. W.; Henson, B. F.; Sander, R. K.; Ali, A. N .; Zielinski, P. M.; Phillips, D. S.; Schwarz, R. B.; Skidmore, C. B. J. Phys. Chem. B 1999, 103, 5434. (115) Davis, R. E. J. Am. Chem. Soc. 1962, 84, 892. (116) Custelcean, R. Vlassa, M.; Jackson, J. E. Angew. Chem. in press. (117) (a) Buoen, S.; Dale, J.; Krane, J. Chem. Commun. 1982, 1172. b) Whitbread, S. L.; Politis, S.; Stephens, A. K. W.; Lucas, J. B.; Dhilon, R. S.; Lincoln, S. F.; Wainwright, K. P. J. Chem. Soc. Dalton Trans. 1996, 1379. (118) Becke, A. D. J. Chem. Phys. 1993, 98, 5648. (119) Akkaya, E. U.; Huston, M. E.; Czamik, A. W. J. Am. Chem. Soc. 1990, 112, 3590. (120) Bernstein, J .; Stearns, B.; Shaw, E.; Lott, W. A. J. Am. Chem. Soc. 1947, 69, 1151. (121) Schlesinger, H. I.; Brown, H. C.; Hyde, E. K. J. Am. Chem. Soc. 1953, 75, 209. (122) Buoen, S.; Dale, J .; Krane, J. Acta. Chem. Scand. 1984, B 38, 773. 154 (123) (a) Hancock, R. D.; Shaikjee, M. S.; Dobson, S. M.; Boeyens, J. C. A. Inorg. Chim. Acta 1988, 154, 229. (b) Dhillon, R. S.; Madbak, S. E.; Ciccone, F. G.; Buntine, M. A.; Lincoln, S. F.; Wainwright, K. P. J. Am. Chem. Soc. 1997, 119, 6126. (124) Eckert, H. Prog. Nucl. Mag. Res. Sp. 1992, 24, 159. (125) Gaussian 98: Frisch, M. J .; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Zakrzewski, V. G.; Montgomery, J. A.; Stratmann, R. E.; Burant, J. C.; Dapprich, S.; Millam, A. D.; Daniels, A. D.; Kudin, K. N .; Strain, M. C.; Farkas, O.; Thomasi, J .; Barone, V.; Cossi, M.; Cammi, R.; Mennucci, B.; Pomelli, C.; Adamo, C.; Clifford, S.; Ochterski, J .; Petersson, G. A.; Ayala, P. Y.; Cui, Q.; Morokuma, K.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Cioslowski, J .; Ortiz, J. V.; Stefanov, B. B.; Lin, G.; Liashenko, A.; Piskorz, P.; Komarorni, 1.; Gomperts, R.; Martin, R. L.; Fox, D. J .; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Gonzalez, C.; Challacombe, M.; Gill, P. M. W.; Johnson, B. G.; Chen, W.; Wong, M. W.; Andres, J. L.; Head-Gordon, M.; Replogle, E. S.; Pople, J. A.; Gaussian, Inc., Pittsburgh PA, 1998. (126) SHELXTL: Structure Analysis Program 5.1; Bruker AXS, Inc., Madison, WI, 1997. 155 MICHIGAN STATE UNIVERSITY LIBRARI- lwmmww