MSU LIBRARIES an. RETURNING MATERIALS: Place in book drop to remove this checkout from your record. FINES wiII be charged if book is returned after the date stamped below. GALLERY AND SURFACE PROPERTIES OF LAYERED DOUBLE HYDROXIDES BY Kevin Julius Martin A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1986 ABSTRACT GALLERY AND SURFACE PROPERTIES OF LAYERED DOUBLE HYDROXIDES BY Kevin Julius Martin Layered double hydroxides (LDHs) are a class of compounds of general formula [M(II)1_XM(IIIIXKH02] [x/zY' °nH20], where x ranges from 0.16-0.33. The gallery anions are exchangeable, making LDHs negative analogs to clay minerals. The structural properties of LDHs are well docu- mented. Very little has been done to characterize the properties of the gallery anions or the surface properties. Most LDHs that are synthesized are dehydroxylated to form oxide catalysts. No reactions have been reported for LDHs in their pristine form. This dissertation will describe the investigations into the properties of the gallery anions of LDHS. LDHs containing a paramagnetic anion asaispin probe were synthesized. Because the ~SO3 groups of the spin probe K2(SO3)2NO bind to surface hydroxyl groups,ru>spin probe was intercalated even when it*was added to the syn- thetic mixture. Therefore, no information about the mobil- ity of gallery anions was obtained from EPR studies. The availability of gallery anions of [ZnZCr(OH)6] [Y‘nHZO], where Y’=I, C1, F, for nucleophilic displacement reactions was investigated. The reaction rate of iodide substitution of alkyl bromides under solution-solid conditions was independent of the alkyl chain length. The Kevin Jul ius Martin reaction rate depended on the surface properties of the LDH. The rates also depended on the nature of the solvent in which the reaction occurred. From adsorption studies it was deduced that the more tightly a solvent was bound to the solid, the slower the reaction. Chloride substitution reactions did not occur under solution-solid conditions, but did occur under gas-solid conditions. Fluoride substitution did not occur under either conditions. A detailed mechanism is proposed in which the bromo group of the reactant molecule must insert into the gallery of the reactant LDH. The mechanism is consistent with all observations regarding the kinetics behavior and adsorptive properties of the Zn-Cr LDH. TO JOHN CLIFFORD MARTIN 1944 - 1974 H ACKNOWLEDGMENTS I would like to thank Dr.1LJ. Pinnavaia for patience and guidance down the long road of graduate school. I look forward to future interactions. I also wish to thank Dr. C.H. Brubaker for careful editing of this manuscript. I hope we got all the typos and grammar mistakes. I would like to express my gratitude to all past and present group members with whom I have come in contact. Your support and questions have helped a lot. I would like to thank my mother, siblings and nephews for emotional and financial support. You helped me keep my sense of humor through it all. Finally, thanks to my mother-in-law Norma Donovan for use of the word processor, and to Susan and Elizabeth, words alone cannot express what I owe to you in the com- pletion of this work. Maybe now we can live like normal people. 111 TABLE OF CONTENTS a e LIST OF TABLES 0.0.0.0000...O...OOOOOOOOOOOOOOOOOOOOOVI LIST OF FIGURES OIOOOOOOOOOOOOO00.0.0000...OOOOOOOOOOViii CHAPTER 1 0 INTRODUCTION 0 I O O O O O O O O O O O O O O O O O O O O I C O I O O O 1 A. Synthesis and properties of layered double hYdIOXides 00.0.0.0...0.00.00.00.000000000001 B. Supported reagents .........................23 C. Objectives of current research .............34 CHAPTER 2. EXPERIMENTAL SECTION .....................35 A. Synthesis of materials .....................35 1. [ZnZCr(OH)6][C1'nHZO] ....................35 2. [Ca2A1(OH)6][C1°nHZO] ....................36 3. [LiA12(0H)6][0.5504-nH201 ................36 3a. Spin-labelled [LiA12(OH)6][0.5804'nH20] .37 4. Ion exchange reactions ...................38 B. Displacement reactions over [ZnZCr(OH)6] [Y.nHZO] 0..0....0...0...0.0.00.000000000000038 1. Reactions under triphase catalytic conditions ...............................38 2. Reactions under stoichiometric conditions OOOOOOOOOOOOOOOOOOOOOO0.0.0....39 3. Gas-solid reactions ......................40 C. Physical methods ............................41 1. X-ray diffraction ........................4l 2. EPR studies ..............................42 3. IR StUdies OOOOOIOOOOOOOOOOOOOOOO0.0.0....42 1v CHAPTER 3. page Gas chromatography .......................42 surface areas 0 O O O O O O O O O O O 0 O O O O O O O O O O O O O O O 43 Adsorption studies 0 O O O O O O O O O O O O O O O O O O O O O .43 RESULTS AND DISCUSSION 0.0.0.000000000000046 A. Properties Of LDHS 0.00.00.00.00.0.00.00.00.046 1. 2. Exchange and swelling properties .........46 EPR StUdies 0......COCOOOOOOOOOOOOOOOOO00.55 B. Nucleophilic displacement reactions .........66 1. Reactions under triphase catalytic conditions OOOOOOOOOOOOOOOOOCOOO0.0.00.0..66 Stoichiometric biphase reactions in t01uene 0.00....0.00.0.0....0.0.0.0000000070 Adsorption properties of Zn-Cr LDHs ......86 Reaction and sdsorption studies in other SCI-vents .0.0.00.0.0...0.0.0.0000000000000102 Thermodynamics of reaction ...............199 Proposed detailed mechanism ..............112 C. conCIUSions .0.0...0...0...0.0.0.0000000000000124 LIST OF REFERENCES OOOOOOOOOOOOO00.00.00.000000000000127 LIST OF TABLES Table Page 1 Some Natural LDH Minerals...............3 2 Phases Formed Upon Aging Ni-Al LDHs.....12 3 Variation of Basal Spacing In [Ca2A1(OH)6 ][Y' 2820] and [an Cr(OH)5 ] [Y nHZO] For Various Y Species.u.6u.u.l3 4 Dependence of Lattice Parameters on Octahedral Layer Composition for Some LDH SpeCieSOOOOOOOOOOOOOOOOIOOCOOOOOOOOOl4 S Dehydration Temperaturesa for Some Mg-Al and Ni-Al LDHSoooooooooooooooooooo18 6 Dehydroxylation Temperaturesa'b For some LDH carbonates.C...................19 7 Swelling of [Zn Cr(OH)6][C12H250803] in Alcohols an Amines..................2l 8 Reactions Utilizing Polymer-Bound ucleophiles of the Form r-CHNMeX- .00...0.0.00.0000000000026 2 3 9 EPR Correlation Times for PADS-doped [Ii-A1 LDH MateriaISOOOOOOOOOOOOOOOOOOOOO61 10 Conversion of Alkyl Halides by the Reaction RX + Y'-9RY'+ X' Under Triphase Consitions.....................67 11 Results of Blank Reactions for Halide EXChangeoooooooooooooooooo0000000000000069 12 Conversions of Nucleophilic Displacement Reactions of C4H98r Under Modified conditiOUSOO0.0.00.0...0.0.0.0000000000071 l3 Calculated Pseudo First Order Rate Constants, for Iodide Substi- tution of AlEyI Bromides over A. D. [ZnZCr(OH)6][I nHZO]....................72 v1 Table Page 14 Pseudo First Order Rate Constants, kobs' for Iodide Substitution of Alkyl Bromides Over O.D. [ZnZCr(OH)6] [I' OHZO]...............................74 15 Activation Parameters for Some Alk 1 Halide Iodide Substitutions at 90 C....76 16 Conversions of Gas-solid Iodide Substi- tutions of l-Bromobutane Over . [znzcr(OH)6][I]oooooooooooooooooooooooo.89 l7 Conversions of Gas-Solid Chloride Substitutions of l-Bromobutane Over [zn2CI(OH)6] [C1]........................8g 18 Maximum Adsorption Capacities and Desorption Experiment Results for Various Adsorbates on O.D. [ZnZCr(OH)6][I].........................106 l9 Pseudo First Order Rate Constants for ' Iodide Substitution of l-Bromobutane with 0.0. Zn/Cr/I LDH at 90°C in Various Solvents........................107 20 Activation Parameters for Iodide Substitution of l-Bromobutane at 90 °C by O.D. [ZnZCr(OH)6][I] in various Solvents................................108 21 Values of Der - DC and 24.2 dEGI for eac Various Alky Hal1 e Exchange tions RX + LDH(Y)#RY + LDH‘X)000000000000000111 v11 LIST OF FIGURES Figure Page 1 Structure of layered double hydroxides (water molecules omitted for clarity)u.2 2 Diagram of the coordination geometry around the calcium centers in [Ca Al(OH)6]HN12H 0] showing the dis ortion6 caused y coordination to gallery water. ..........................6 3 The structure of montmorillonite, a clay mineral of the smectite group......7 4 a.) The effect of aging on the product distribution of coprecipitated Ni-Al LDH for a solution value of x=0.25. b.)The effect of the initial product distri- bution on the shape of x-ray powder diffraction peaks. Adapted from reference S8............................l7 5 Structure of an alkylammonium-exchanged smectite swollen with a linear alcohol. From reference 2S.......................22 6 A conceptual diagram of triphase cataIYSiSOOOOOOOOOOO0.0.0.0000...0.0.00028 7 A diagram of a smectite clay with gallery ions as a result of a.) homoionic intercalation of cations, and b.) intersalation of two equiva- lents of cation and one of anion. Structure b acts as an anion exchangeru32 8 Powder x-ray diffractograms of (a). nCr(OH)6 ][C1' 2H O];~(b). [LiAl (OH)6 SO4°nH2601; and2 (c). [Ca2A1(OI-l%6] [0. SCO43'ZHZO]...........................47 9 Powder x-ray diffractograms of [Ca Al(OH) ][G.SCO3'2H O] (a). pressed pow er; b) oriente film.u.u.n.u.n49 V111 Figure Page 10 Powder x-ray diffractograms of the dodecylsulfate-exchanged forms of (a). Zn-Cr LDH; (b).Ca-Al LDH; and (c). Li'Al LDHOOOOOOOOOOOOOOOOOOOOOOOOOOO0.0.S]- ll Powder x-ray diffractograms of n-octanol-swollen DDS-exchanged forms of (a). Ca-Al LDH; and (b). Li-Al LDH...52 12 Infrared spectra from 4000 cm'1 to woo curl of [ZnZCr(OH)6][Cl'nH20] (a). pristine material; and (b). exposed to 0.1 M NaDDS solution.................S4 l3 EPR spectra of K PADS (a). 10"5 M solution in DMS ; and (b). frozen in DZOOOOOOOOCOOOOOOOOOOOO0.00.00.00.00000056 14 EPR spectrum of [Zn Cr(OH)6][C1°nH201. Frequency = 9.14 Hz....................58 15 EPR spectra of PADS-doped [LiAlg(OH)6] ) [Cl'nH O] (a). damp material; ( after 2 hr in vacuum dessicator; and (c). after 12 hr in vacuum dessicator. Frequency = 9.14 GHz....................S9 16 EPR spectra of an Li-Al LDH synthesized in the presence of NaDDS and K PADS (a). damp material; (b). after 4 hr in vacuum dessicator; and (c). after 12 hr in vacuum dessicator. Frequency = 9.14 GHz.62 1? Powder x—ray diffractograms of unaged Li-Al LDH product (a). filter-washed; (b). dialyzed until free of electrolyte.............................64 l8 Powder x-ray diffractogram of product formed by reaction of LiCl solution and gibbSiteOOOOOOOOOOOOOOOOOOOOOOOOO...65 19 First order kinetic plots for iodide substitution of alkyl bromides over [ZnZCr(OH)6] [I’2.3H20]: ll-bromo- butane; Al-bromohexane; and Ol-bromo- 3-methy1butane..........................73 20 IR spectra from 4000 cm’1 to 1000 cm"1 of (a). [ZnZCr(OH)6][I°2.3HZO]; (b). [ZDZCI(OH)6] [1.6H20]00000000000000.000075 1x Figure Page 21 First order plots for iodide substi- tution of l-bromobutane over OJL Zn/Cr/I LDI-l: (a). 122 °C; (b). 1001 °C; (c). 90 °C; ((1). 80 °C; and (e). 5e °c..77 22 Arrhenius activation energy plot, 1n kobs vs. l/T, for iodide substi- tution of l-bromopentane over OJL Zn/Cr/I LDH.............................78 23 First order kinetic plot for the extended iodide substitution of 1-bromo- butane over (LD. Zn/Cr/I LDHuu.n.u.u82 24 First order kinetic plot for iodide substitution of 1-bromobutane by OJL Zn/Cr/I LDH at RBr/LDH ratios of I4:1; 03:1; 02:1; A1:1, plotted as functions of (a). -1n fI-; (b) -1n fRBr'84 25 Isotherms for adsorption of toluene onto: (a). A.D. Zn/Cr/I LDH at .0 °C; 022 °C; (b). 0.0. Zn/Cr/I LDH at A0 °C; 322 °C88 26 Isotherms for adsorption of 1-iodo- butane'onto: (a). A.D. Zn/Cr/I LDH at on °C; 022 °C; (b). o.o. Zn/Cr/I LDH at no °C; I 22 °C......................89 27 Isotherms for adsorption of 1-bromo- butane onto: (a). A.D. Zn/Cr/Br LDH at .0 °C; 022 0C; (b). O.D. Zn/Cr/BrrLDH at A0 °C; I 22 °c......................90 28 An adsorption isotherm showing the location of Point B, the attainment of monolayer coverage......................92 29 Heat curves, qS vs. V/V , for adsorp- tion of toluene onto [ZnZCr(OH)6] [1.nH20] ‘AOD0.0.D...OOOOOOOOOOOOO00.093 30 Heat curves, q t vs. V/Vm for adsorp- tion of l-iodogutane onto [ZnZCr(OH)6] [I.nHZO] ‘AoDo .O.D.-00000000000000.0094 31 Heat curves, qst vs. V/Vm, for adsorp- tion of l-bromobutane onto [ZnZCr(OH)6] [1.nH201 ‘AOD0.00DOO0.0.00.0000000000095 Figure Page 32 Plot of the enthalpy of liquid adsorp- tion,AH ads' vs. V/V for toluene adsorbed onto [ZnZCrTOH)6][I'nH20] A.A.D. .O.D....O0.00000000000000000000000000.96 33 Plots of AHlads vs. V/Vm for adsorption of l-iodobutane onto [ZnZCr(OH)6] [I'nHZO] QA.D.OO.D.; and 1-bromo- butane onto [ZnZCr(OH)6][I'nH20] I A.D. ‘O.D....OOOOOOOOOOOOOOOOIOOOO00.0.000097 34 Adsorption of toluene onto O.D. [ZnZCr(OH) ][I‘nH O] as a function of pressure a? A0 0 I22 °C..............99 35 Plots of AHla 3 vs. V/Vm for O octane and Acarbon getrachloride adsorption onto O.D. [znzcr (OH)6][I'DH2010000000000163 36 Plots of AHlads vs. V/Vm for adsorption of.tchlorobenzene and . toluene onto O.D. [ancr(OH)6] [I'DHZO]ooooooooooooooolg4 37 Possible bonding interactions of hydroxyl group orbitals with an aromatic system. (a).e2 +6‘+ - 9 OH no net overlap; (b). e2 + H2p - p-orbital too high in eaergy; (c). 0.5 e2 +5+OH - probable interaCtio 0.000000000IOOOOOIOO0.0.0.000114 x1 Chapter 1. Introduction A. Synthesis and properties of layered double hydroxides. There are a large number of compounds that can be classified as layered double hydroxides1 (LDH). This class of materials, also known as hydrotalcite-like compounds (HT)2 and Feitknecht compounds3, has both synthetic and naturally occurring members. Most of these species, of general formula [M(II)1_XM(III)X(OH)2][x/zYz"nH20]1, con— sist of brucite-like octahedral sheets with trivalent metal cations periodically substituted for the divalent metal cations. This substitution imparts a net positive charge on the sheets which is compensated by anions intercalated between the stacked hydroxide layers together with n water molecules (Figure 1)l. These compounds will be the subject of this dissertation. Related compounds exist with only 3 anala- divalent cations in hydroxide-carbonate structures gous to those of malachite, Cu2(OH)2CO3; hydrozincates, Zn5(OH)6(CO3)2; and azuriteé, Cu3(OH)2(CO3)2. These mixed- metal hydroxides with no trivalent cations will not be considered in this work. Also, natural minerals with more than one octahedral layer per anionic layer, such as coa- lingite7 and carrboyditea, will not be considered. Only a small percentage of the known LDH species occur in nature, the most common of which is hydrotalcite, m. w. m. H\I.H H \I..H owom. wMo M A M Structure of layered double hydroxides (water molecules omitted for clarity). Figure l. 3 Table 1. Some Natural LDH Minerals.a Name Idealized Formula hydrotalcite, [Mg6A12(OH)16][CO3'4HZO] mannasseite pyroaurite, [M96F92(0H)16][C03'4H201 sjogrenite stichtite, [Mg6Cr2(OH)16] [C03'4H20] barbertonite reevesite [Ni6Fe2 (OH) 6][CO3' 4H20] eardleyite [N16 Al (OH) [CO3 4H meixnerite [M96 AI «n16][(OH)2%H20] takovite [Ni6 Al OH)1[(OH)2'4H20] hydrocalumite [Ca62Af(OH)6]6 [OH 2H 25] aFrom reference 9. [Mg6A12(OH)16] [CO3°4H2019. The other formulae in Table l with two entries in the name column exist in two polymorphic formsg. The most extensively studied natural LDHs from a structural standpoint are the pyroaurite/sjog- renite polymorphs, [Mg6Fe2(OH)16] [CO3'4H20]. Early powder diffraction work10 showed that sjogrenite is hexagonal with a=3.13 A and c=15.65 A. Pyroaurite is rhombohedral with a=3.l3 A and c=23.47 A. Single crystal studies of sjogren- ite and’pyroaurite confirmed these data11‘13. The hydroxyl groups of neighboring octahedral layers stack on top of each other. The rhombohedral pyroaurite has three octahe- dral layers per unit cell stacked in the sequence BC-CA- A812'13. The AB-BA-AB stacking sequence of sjogrenite leads to a hexagonal unit cell with two octahedral layers per unit ce11.11'12 In the samples examined, there was no the Mg(II) and order to the cation site distribution, i.e., Fe(III) centers were statistically distributed throughout the octahedral siteslz. Much of the early attention focused upon synthetic LDHs resulted because a calcium-aluminum compound formed as an intermediate in the setting of Portland cement was in fact a Ca-Al LDH14. A report from 192915 showed that two metastable compounds of "calcium sulphoaluminate" were formed by interaction of calcium aluminate and calcium sulfate. The formula of one of them, given as 3CaO°A12 3' CaSO4'12H20, can be rewritten as [Ca2A1(OH)6][1/2304'3HZO]. No structural data were given,lnn:the formula falls into the category of a LDH. Later studies16 of the Ca-Al system confirmed that it was an LDH. In the following years, many reports appeared on the synthesis of these basic double salts, as they were called before the structure was known17. From x-ray powder dif- fraction data came the knowledge that the 001 parameter of the materials varied as the size of the anion used in the synthesis or exchanged into the structurela.’ A double layer structure was proposed19 in which the divalent metal formed a hydroxide layer, like brucite, with the trivalent metal and anions intercalated between these sheets. One group20 has used infrared spectroscopy results to support this structural argument. Specifically, the C032' vibra- tions were consistent with Al—OCOz-Al bridges in the gal- leries between brucite-like sheets. This structure is today not generally accepted today because of the single crystal work reported more recently. Synthetic analogues of the natural mineral hydrocalumite, [Ca2A1(OH)6] HNP2H20]9, have been investigated by single crystal x-ray diffraction21‘24. The results of these investigations are in essential agreement with the reported structures of pyroaurite and sjogrenite in that the metal cations are contained in octahedral hydroxide layers with anions inter- 21422. Because of the large size of calated between them Ca(II) compared to Al(III), the metal cations were ster- ically limited to an‘ordered arrangement23. This result was different from the random distribution of Mg(II) and Fe(III)iJIpyroaurite and sjogrenite».12 The0.33 led to the formation of Al(OH)3 phases along with the LDH product. Similar results have been reported for aged Ni-Al products.57 For the Mg-Fe system,55 pure product was formed in the range of 0.27ng0.15. For the Mg-Al and Ni-Al compOunds the values of x indicate an upper limit on the M(II)/M(III) ratio of 5 and a lower limit of 2. The limits for the Mg-Fe system appear to be 5>M(II)/M(III)>3. All these data are for materials aged either for 4-5 days at room temperatures“, or 5 days at 60 0C.48"55'57 When freshly-precipitated Ni-Al materials were examined results different from the aged LDHs were reported.57 For freshly-prepared materials, single phases were found for values 0.5Zx20.15, as shown in Table 2. When the compounds were aged hydrothermally for 2 days at 150 0C, the samples with x>0.33 or x<0.25 contained LDHs along with an additional phase, as indicated in the table. This was interpreted as evidence that the compounds formed initially when x>0.33 or x<0.25 were metastable, and that they changed into the more stable LDHs and either bayerite 12 Table 2. Phases Formed Upon Aging Ni=Al LDHs.‘a Solution xb Before Aging After Aging 0.85 LDH LDH, Ni (0102 0.80 LDH LDH, Ni (0H)2 0.75 LDH LDH 0.66 LDH LDH 0.60 LDH LDH, boehmite 0.50 LDH LDH, boehmite 0.40 LDH, —- bayerite :From reference 57. X=[M(III)]/([M(II)]+[M(III)])o or Ni(OH)2 upon hydrothermal treatment.57 This conclusion was:h1accord with general observations concerning fresh precipitates of low crystallinity,34 that they are often metastable under their conditions of preparation and that their composition is usually flexible. This points to the importance of aging coprecipitated LDHs, a step mentioned in every synthesis of this type. The importance of impro- ving the crystallinity of the product can be seen in the Ca-Al example.15'21"24 The material prepared by coprecipi- tation and unaged was metastable in its reaction mixture.15 The same compound formed under hydrothermal conditions,21"24 with aging built into its synthesis, was very stable. So improving crystallinity is not only impor- tant for structural studies, but also for increased stabi- lity of the precipitates. An extensive literature exists on pOwder x-ray dif- fraction of synthetic LDHs. Almost all the patterns recorded can be indexed on the basis of either a hexagonal 13 or rhombohedral unit cell. As discussed earlier, this in- dexing corresponds to two layers and three layers per unit cell, respectively.12 In the following discussion, the rhombohedral species will be discussed on the basis of hexagonal unit cell parameters _a_ and _c_ For simplicity, 0. 9f, the actual layer repeat distance, will be used instead of c since 2' is independent of symmetry group. In —o' Table 3 the values of g' for two LDH series of constant octahedral layer composition are given. The 3' parameter varies as a direct function of the radius of the gallery anion.l If the octahedral layer thickness (4.8 A= two layers of hydroxidesg) is subtracted from gf, the result is close to the diameter of the anion. Based on this assumption, the values of g‘ for the C032" and N03 forms of the Zn-Cr LDH should be approximately the same. For the C0327 form, 5y-4.9 A = 2.7 A, which is the approximate diameter of an oxide group.59 This data agrees well with Table 3. Variation of Basal Spacing In [Ca Al(OH)6 ] [Y ZHZO] and [Zn2Cr(OH)6][Y nHZO] or Various Y Species. Metals Y' 2' (A) References Ca, Al OH 7.4 23 Ca, A1 C1 7.9 22 Ca, Al 0. 5CO3 7.6 8 Ca, Al 0. 5504 8.9 21 Zn, Cr F 7.51 31,32 Zn, Cr Cl 7.73 ,3l,32 Zn, Cr Br 7.85 31,32 Zn, Cr 0. 5CO3 7.60 31,32 Zn, Cr No3 8.88 31,32 14 the single crystal results12 that show carbonates in an orientation parallel to the hydroxide sheets. Apparently since there are twice as many nitrates per unit cell as carbonates, there is not room for them to orient in the same manner as the carbonates.32 Therefore, they must stand on end to all fit into the interlayer. This effect has been observed for other IJHL54'57 Table 4 shows the dependence of g-and E} on the constitution of the octahedral layers for several carbonate intercalated LDHs. There are several observations which can be made about the data presented in the table. For a given pair of metal cations, as the value of x increases, the value ofgy decreases. This decrease was said postulated as the result of increased electrostatic attraction between the octahedral layers and anionic interlayers as the den- sity of charge centers increases.5g This effect was not only observed in the Mg--Al"’9'50 and Ni-A150'57, but also in the Mg-Fe55 LDH series. Table 4. Dependence of Lattice Parameters on Octahedral Layer Composition for Some LDH Species. M(II) M(III) xa 3(A) g'(A) Reference Mg A1 0.16 3.07 7.9 49 Mg A1 0.25 3.06 7.90 54 Mg A1 0.30 3.05 7.73 50 Mg Al 0.34 3.04 7.57 54 Ni Al 0.17 3.05 7.86 50 Ni A1 0.22 3.04 7.76 50 Ni A1 0.28 3.02 7.58 50 ax=lMinter- actions of the clay surface with the cation and its. strongly-held solvation sphere. Gallery anions of LDHs lack this type of solvation sphere. Therefore, LDHs would not be expected to swell.in the same manner as smectites. In Ca-Al LDHs, however, because of its large radius, the Ca(II) is exposed to the interlayer region. In the crystal structure of [Ca2A1(OH)6] [OH'ZHéO], a water molecule is coordinated strongly enough to the Ca(II) center to distort its geometry-from octahedral.]‘2"13 It was suggested that this cation was exposed enough to hold an extra layer of water in the same way the cations in smectites attract extra layers. That this phenomenon was observed only when OH‘ was the gallery anion suggests hydrogen bonding inter- actions as important factors in the swelling. 21 The other example of swelling in LDHs was reported for Zn-Cr LDHs.31 When an alkyl sulfate exchanged form of the material was exposed to either a linear primary alcohol or amine, the interlayer spacing was found to increase» The magnitude of the swelling was dependent upon the hydro- carbon chain length of the swelling agent, as shown in Table 7. This swelling phenomenon was said to be the result of solvation of the alkyl chains of the anion to give an approximate double layer structure.31 This be- havior was compared to that of alkyl ammonium substituted layer silicates.25 A model of this interaction is shown in Figure 5. Among the patented uses of LDHs are applications as ion 41 64 in- exchangers, carriers for anionic pharmaceuticals, hibitors of thermal and ultraviolet degradation of thermo- 65 plastic resins and scavengers of Cl’ from Ziegler-Natta Table 7. Swelling of [ZnZCr(OH)6][C12H250803] in Alcohols and Amines.a Swelling Agent Interlayer Spacing (A) noneb 26.2 C4H90H 29.2 C6H130H 30.8 C10H210H 38.2 C12HZSOH 41.1 C12H25NH2 41.7 :From reference 31. Exchanged in water, air-dried. 22 \ .\\ NH3 on on m3'( NH3 1\\\\\\\\\\\\\\ Figure 5. Structure of an alkylammonium-exchanged smectite swollen with a linear alcohol. From reference 25. 23 66 catalyst residue in plastics. By far the greatest utili- zation of LDHs has come in the field of catalyst preparation.34.56-58,67-71 LDHs formed by coprecipitation have been calcined to form oxides to catalyze aldol condensations.7g'71 Ni-Al LDHs have been calcined, then \reduced, to give supported nickel catalysts for steam 67 Cu-Zn—Al LDHs have been treated in a similar reforming. manner to produce catalysts for methanol synthesis.69 Part of the extensive literature on the preparation of these catalysts is concerned with the characterization of the catalyst precursors (LDHs), and much of the data quoted in this work comes out these reports.55'58'67'68 To date, however, no reactions involving uncalcined LDHs have been reported. 2. Supported reagents. One of the major advances in organic chemistry in the recent past has been the use of supported reagents in synthesis72. In this technique, reactants for organic reactions are bonded to a polymeric organic backbone73, or somehow attched to an inorganic support72'74. This can be accomplished by covalent bonding, adsorption or intercalation. These supported reagents offer several 72-75. (1) The effective surface area for advantages reaction is increased. This accelerates the reaction rate by increasing the number of reactant pairs brought together. (2) Solids always have microscopic pores on their surfaces. These pores may serve to constrain the 24 substrate and reactant, and therefore lower the entropy of activation of the reaction. (3) Product workup after reac- tion is usually simplified. This is especially true where, without a solid support, a mixture of solvents must be used to bring the reactants together; This situation can lead to complex distillation schemes or azeotrope problems. With the supported reagents the solvent for the solution- phase reactant can be chosen so that separation problems are kept to a minimum, or in some cases no solvent is needed at all. (4) In many cases the supported reagent can be easily regenerated. Among the solid supports which have been utilized in the formulation of supported reagents are:72'73 Celite,‘ silica, alumina, graphite, activated carbon, montmoril- lonite, molecular seives, polymers and resins. Of the inorganic supports, Celite-supported silver carbonate was among the first to be utilized72. Acid and base forms of ion exchange resins were among the first organic solid 76. Since these early reagents to be used in synthesis uses, many other reagents supported on solids have been devised. The range of reactions studied encompasses all classes of organic synthesis. The focus here will be on nucleo- philic substitution reactions by anionic species. In general, these reactions are carried out by preparing an anion exchange resin in the form of the desired nucleo- phile, then mixing the substrate with the resin at the 25 desired temperature73. One of the first reactions to be reported using this technique was the cyanation of benzyl halides to their corresponding cyano compounds77 . Among the problems associated with this reaction in the solution phase was the formation of side products by competing elimination reactions. The polar solvent used in these reactions, usually acetonitrile, promoted these side reactions. With the supported reagent, only the substrate benzyl halide need be dissolved so that a nonpolar solvent could be utilized. The problem of elimination reactions was solved in most of the systems studied.76 A partial listing of other nucleophilic displacement reactions which have been carried out using polymer-supported reagents is given in Table 8. As can be seen, a number of reactions have been reported. Many of the reactions shown were difficult to carry out in the solutions phase. The polymer support seemed to increase the nucleophilicity of such 79 86 reactants as F“ and sulfonates, allowing reactions such as this to be performed simply. One application of supported reagents is the technique of triphase catalysi387'88, 89,90. which is an extension of phase transfer catalysis Phase transfer catalysis provides a way to react two mutually—insoluble reagents without the use of a cosolvent. All phase transfer reactions involve 9“ (1) transfer of one reagent from its at least two steps: "normal" phase into the second phase; and (2) reaction of the transferred reagent with the nontransferred reagent. 26 Table 8. R actions Utilizin Pol mereBound Nucleo hiles of t e Form ~CH g+Me §a p 2 3 X“ Reaction System References F conversion of sulfonyl chlorides and ' 78,79 alkyl halides to fluorides Cl,Br,I halogen exchange with alkyl halides 79,80 CN alkyl halides to nitriles 77,81 SCN alkyl halides to thiocyanates or isoe 81,82 thiocyanates OCN alkyl halides to ureas and urethanes 82 N02 . alkyl halides to nitroalkanes and 81,83 nitrite esters OCOR alkyl halides to esters 84 OH condensation reactions 85 OAr alkyl halides to ethers 83 OZSPh synthesis of alkyl sulfones 86 27 Some of the species which have been found to function as 90'91 quaternary ammonium and phase transfer catalysts are phosphonium salts, crown ethers, cryptands and linear poly- ethers. Details of the mechanism of reaction vary with these different catalyst systems, but the simplest case involved SNZ-type nucleophilic displacement reactions as shown in Scheme 1.89 The catalyst provides a way of fer- rying the nucleophile to the organic reagent, then regene- rating in thelaqueous phase. More than 1000 publications and patents on phase transfer catalysis attest to its util- ity in these and many other reactions.88'90 or + RX .\.-=-_-‘: RY + ox (organic) QY + MX : MY + QX (aqueous) Scheme 1 More recently, a synthetic approach has been developed that simplifies catalyst recovery and avoids the problem of emulsion formation sometimes encountered in phase transfer catalysi387'88. This approach, termed triphase catalysis, is an extention of phase transfer catalysis. A conceptual diagram of triphase catalysis is given in Figure 6. Most triphase catalyst systems have involved the attachment of one of the active phase transfer species to an insoluble support. The supports utilized include polystyrene- divinylbenzene polymers87 and silica gel92'93. Smectite 28 AQUEOUS ORGANIC Figure 6. A conceptual diagram of triphase catalysis. 29 clays modified so that they became anion exchangers94 have also been utilized In order for these materials to function as triphase catalysts, they must provide(): OJ anion ex- change capacity; (2) an environment in which these aqueous anion exchanges can take place (idh, a degree of hydro- philicity); and (3) an environment in which the organic reagent has access to the nucleophile (idh, a degree of organophilicity). The way in which these criteria were met in the aforementioned systems will now be described. By'far the greatest number of reports have concerned themselves with polymer-based catalysts.88 In these sys- tems the active sites consist of quaternary ammonium or phosphonium groups bound to the polymer matrix through a hydrocarbon chain. Crown ethers and cryptands have also been effectively utilized as active species.95 The polymer matrix employed in these studies is usually polystyrene crosslinked with 1-4% divinylbenzene. The organic matrix and hydrocarbon groups surrounding the anion exchange site make the solid organophilic enough for the organic reagent to interact with the nucleophile. At the same time the anion exchange site is exposed enough to be accessible to the aqueous solution of nucleophile that regenerates the active sites. One of the problems associated with the use of these supports is matching solvent with polymer.90 These polymers swell to differing degrees in different solvents as a function of their percentage of crosslinking. In the cyanation of l-bromooctane, the activity of the 30 catalyst depends on the swelling properties of the resin 9“ Specifically, the resins that swell more in the support. organic phase than in water showed greater activity. This effect can be explained on the basis of diffusion. If the resin were swelled with organic phase, there would be a large amount of reactant dispersed in the resin. Under these conditions diffusion of the organic reactant to the active site would be more favorable than if the aqueous phase predominated in the resin. Since the reaction rates were higher when the organic phase was present, the inter- pretation was that the rate is controlled by diffusion of the substrate to the active site on the catalyst. Another way in which the ammonium and phosphonium salts have been supported is by attachment to silica gel.92’93 Attachment was accomplished by way of a functionalized organosilane coupling agent. These catalysts were effective in many of the same reactions as the more prevalent resin-bound materials. These solids provided access to both organic and aqueous phases in much the same way that the polystyrene-supported materials did. There was 96'97 as there no problem with swelling in these catalysts, was with the resins. Therefore, in order to simplify product workup, the reactions could be run without an organic solvent. There was a problem, however, with water adsorbed on the surface of the silica that created a hydrophilic environment around the active sites of the catalyst under certain conditions. These conditions arose 3] if the hydrocarbon chain on the silane group was too short, or long enough to allow the active site to be bent to the silica surface.92 Also, with very long chains, phenyl- substituted substrates were solvated by the long chains. These solvated molecules were unavailable for reaction and 93 consequently the rate of reaction was slower. The use of smectite clays as triphase catalyst supports has been (L94'98 Montmorillonite intersalated with recently reporte benzyltri-n-butylammonium bromide catalyzed the reaction of phenol with l-bromobutane in the presense of base.98 The environment of the solid seems to have enhanced the nucleo- philicity of the phenoxy group, as the conversion with the supported system was greater than that using an analagous homogeneous phase transfer catalyst system. In other work that utilized clays, hectorite was intersalated with Ni(phen)3SO4 (phen=l,l0-phenanthroline) and with cetylpyridinium bromide.94'99 A comparison of intercalation and intersalation is given in Figure 7. The net effect of intersalation was to induce anion exchange 94 These materials were utilized capacity in the smectites. as triphase catalysts in halide exchange reactions on alkyl halides. ApparentLy, the use of organometallic complexes and organic cations in the galleries of the clays provided the organophilicity that, along with the aforementioned anion exchange capacity, are the two requirements for pro- ducing a solid catalyst for triphase reactions. In the halide exchange reactions, the [Ni(phen)3]2+ intersalates 32 H «I» 2+ 0 Mu): M(L)3 Homoionic, l8A° [ ____ .____ ____ ____. NHL)? NHL)? NHL)? NHL)? . NHL)? NHL)? NHL)? NHL)? _ 1._J Biloyer lntersoldtion,~30A° Figure 7. A diagram of a smectite clay with gallery ions as a result of a.) homoionic intercalation of cations, and b.) intersalation of two equiva- lents of cation and one of anion. Structure b acts as an anion exchanger. 33 were more effective than tflua cetylpyridinium 99 intersalates. Also, in the caserof the halide exchange reactions using the nickel-based material, the catalyst was 94 The rate of con- somewhat size and shape selective. version of l-chloropentane to its corresponding bromide was reported to be much greater than any of the larger, more bulky alkyl chlorides. The reactions proceeded at almost identical rates under homogeneous phase transfer conditions that used tricaprylmethylammonium chloride as the catalyst.94 Inherent differences in reactivity of these primary alkyl chlorides were apparently not responsible for the differences in conversion using the solid catalyst. There were two possible explanations put forth for the differences in reactivity.99 The first one stated that the differences arose from the layered structure'of the clay. There was only a limited amount of space in the interlayer region of the clay, allowing the smaller, unbranched mole- cules to diffuse more rapidly into the interlayer region where the nucleophile was immobilized. This explanation was made with the assumption that the reaction took place within the clay layers. The other explanation started out with the hypothesis that the reactions took place in an interface between the two liquid phases on the surface of 99 Since the nucleophiles were contained within the clay. the aqueous phase, the smaller, less hydrophobic alkyl would react more quickly by virtue of their greater ability to diffuse into this aqueous phase. This still did not 34 explain why the branched hydrocarbons reacted so much more slowly than their straight-chain analogs. Perhaps the true mechanism was some combination of these two. Objectives of Current Research. As outlined above, the physical properties of layered double hydroxides have been relatively well explored. The chemical properties of gallery anions and surface proper- ties of the materials have not been adequately addressed in the past. In this work, the goal of the research was an understanding of these properties. The properties of the gallery anions and surface were to be explored by use of LDHs as sources of nucleophile under triphase catalytic conditions and under stoichiometric conditions. The inter- actions of LDHs with aqueous and organic phases under these conditions would provide data about the availability of gallery anions for reaction and the relationship of surface properties to reactivity. Chapter 2. Experimental Section. A. Synthesis of Materials. 1. [ZnZCr(OH)6][Cl'nH20]. This compound was prepared by a modified version32 of the original reported procedure. A slurry of 45 g (0.55 mol) ZnO was prepared by adding water to the powder with stirring. To this slurry was added 150 mL of a 1.0 M CrCl3 solution (0.15 mol Cr(III)) in 50 mL water. This addition brought the pH of the solution to 3.4. The mixture was stirred at 60 0C until the solution became colorless, indicating that all of the Cr(III) had been consumed. The clear solution was decanted from the solid, and a further 150 mL of l M CrCl3 and 50 mL water added to the solid, bringing the total Zn/Cr ratio of the mixture to slightly less than 2, the small excess of Cr(III) assuring complete reaction of all ZnO. The mixture was stirred at 60°C for 5 h. The solid product was washed with distilled water by centrifugation until the wash water was free of C1“ (AgN03 test). The lavender-gray product was resuspended in water and poured onto a glass sheet to dry as a 5-10% suspension and allowed to air dry. The product was identified by x- ray powder diffraction (5001:7'735‘)‘ Completeness of reaction was verified by the absense of any peaks attributable to 260.100 35 36 2. [Ca2A1(OH)6][Cl'nH20]. In accord with one of the published methods for prepa- ration22 5.97 g (54.0 mmol) anhydrous CaC12 was dissolved in approx. 75 mL distilled water. To this solution was added 6.96 g (89.0 mmol) Al(OH)3 and 10.0 g (135 mmol) Ca(OH)2, resulting in a Ca/Al ratio of 2.10, a slight excess of Ca(II). The suspension was placed in a Parr stainless steel pressure reactor equipped with a Pyrex liner and a stirrer. The mixture was heated to 150 oC. The temperature was monitored and controlled through use of a temperature controller and a built-in thermocouple. The reaction was allowed to proceed for 100IL. The resulting white product was washed with distilled water by cen- trifugation until the wash water was free of Cl' (AgNO3 test). The material was resuspended and poured onto a glass plate to dry. Identity of the product was confirmed by x-ray powder diffraction through comparison with pub— lished data(d661= 7.9 A)22 and by the absence of any peaks associated with the starting materials.106 3. [LiA12(OH)6][SO4‘nHZO]. This compound was prepared by following the procedure al.101 A saturated solution of LiOH was of Serna et. prepared with freshly boiled deionized water to remove traces of carbonate. This solution was placed in a three- neck flask that had been purged with NZ to remove traces of C02 that could be absorbed by the strongly basic (pH=l3.0) solution. These precautions were necessary because of the 37 reported preference of LDHs for carbonate ion.l As N2 was bubbled through the Solution, a 0.5 M solution of A12(SO4)3 was added dropwise. The pH of the solution was monitored during the addition. When a value of pH=l0 was reached, the addition was stopped. The resulting powder product and its mother liquor were placed in a Parr pressure reactor and hydrothermally treated for 3 days at 130 °C. The resulting product was washed on a Buchner funnel under N2 until the wash water was free of Li+ (flame test). The identity of the product was confirmed by x-ray powder diffraction by comparison of the basal spacings with liter- ature values (d061=l0.45 A)102and the absence of peaks from any other Al or Li containing phases. 3a. Synthesis of spin-labelled [LiA12(OH)6][Cl'nH20]. Nitroxide-doped samples of [LiA12(OH)6][Cl‘nHZO] were prepared for EPR analysis with K2N0(SO3)2, peroxylamine disulfonate (PADS), present in the reaction mixture. Since the PADSZ' species was stable in basic solution and unstable in acid, 0.064 g (0.24 mmol) KZPADS was placed in 5 mL water with 1.0 g (24 mmol) LiOH'HZO. A 0.5 M solution of A1C13 was added as before. The solid was allowed to age at room temperature for 4 h and then washed under N2. The hydrothermal treatment used on the sulfate sample earlier would have caused the decomposition of the spin probe. The solid was stored under Ar in a dessicator to prevent the .decomposition of PADSZ‘ by oxygen. 38 4. Ion exchange reactions. In a typical ion exchange reaction, 1 g of solid was immersed in enough 1 M solution of the desired exchange anion to give a 30-fold molar excess of anion. The mixture was stirred at 60 °C for 2-3 h. The solution was decanted and the process repeated. Usually, one repetition was enough to make theloriginal LDH undetectable in the x-ray diffraction pattern. lecases where solubility problems precluded use of a 1 M solution, as in the case of sodium dodecylsulfate exchange, or thermal instability prevented reaction at 60 °C, as in attempted exchanges with PADS, Llonger exchange times and repeated applications of fresh solution were used to effect complete exchange. B. Displacement Reactions Over [ZnZCr(OH)6][Y'nHZO]. 1. Reactions Under Triphase Catalytic Conditions. The conditions utilized were similar to those used by earlier workers in triphase catalysis in order to facili- tate direct comparison of results. Iodide substitution reactions of alkyl halides were investigated. In these reactions the aqueous phase consisted of 0.899 g (6 mmol) NaI in 3.0 mL water. The organic phase was composed of 1 mmol alkyl halide in 2.0 mL toluene. In all cases a 20:1 molar ratio of alkyl halide:LDH was utilized. So for the reactions, 0.023 g (0.05 mmol) [ZnZCr(OH)6][I'2.3 H20], prepared by anion exchange with the original chloride 39 material, was added to catalyze the reaction. A dodecyl- sulfate-exchanged form (DDS, C12H250803') was also used in some reactions to test whether the presence of an organic species in the gallery would affect reaction yields. In these cases, 0.029 g of the DDS form of the solid was used in the reaction mixtures. All of these reactions were carried out in Pyrex culture tubes sealed with a Teflon- lined cap. The tubes were placed in a constant temperature (Omega Engineering model 4000 temperature controller) oil bath maintained at 90 °C. A magnetic stir plate and spin bars in the individual tubes provided agitation of the reaction mixtures. The reactions were allowed to proceed for 24 h. The products were analyzed by GC. Blank reactions were also run which included all the ingredients listed except the solid LDH. These were carried out and analyzed in the same manner as the triphase reactions. 2. Reactions Under Stoichiometric Conditions. The same reactions carried out under triphase catalytic conditions were also attempted in the absence of an aqueous phase. The iodide-exchanged form of the Zn-Cr LDH was again utilized. The air-dried (AJL) form with n=2.332 and a form dried at 150 0C for 2 h under Ar, the oven-dried (OJL) form with n=0, were tested in the reactions. In both cases 1 mmol equivalent of intercalated I" “L453 9 A.D. LDH or 0.412 g O.D. LDH) was added to 3.0 mL organic solvent in a pyrex culture tube with Teflon-lined cap. To this mixture were added 1 mmol alkyl halide and a magnetic 40 spin bar. The culture tubes were placed in the temperature controlled oil bath at the desired reaction temperature, typically 90 °C. A magnetic stir plate provided agitation of the reaction mixtures. The tubes were withdrawn from the oil bath every 10 min, cooled in an ice bath to quench the reaction and ca. 10 pl of liquid reaction mixture withdrawn. The culture tubes were quickly resealed and returned to the oil bath after each extraction of material. The samples were analyzed by GC to calculate percentage conversions. In some cases the RBr:I' ratio was varied to study the effect on the reactions. Ratios of 2, 3, and 4 were used in addition to the typical 1:1 ratio. 3. Gas-Solid Reactions. A vertically-mounted, fixed bed continuous flow micro- reactor previously used to test for catalytic activity of clay minerals162 was utilized to study the reactivity of LDHs toward vapor phase alkyl halides in nucleophilic dis- placement reactions. The reactor consisted of 7 mm I.D. quartz tubing encased in a tube furnace heated and control- led by a three stage temperature controller (Eurotherm model 919A). The alkyl halide was introduced with a syringe pump (Sage Instruments, model 341A) into a pre- heater zone at the top of the reactor. The vector gas, He, was purified with BASF R3-ll catalyst to remove oxygen and 4A molecular seives to remove water. The He flow rate was controlled through a flowmeter. 41 In a typical reaction, 1 mmol Zn-Cr LDH was loaded into the reactor tube. Several 2 mm glass beads were placed on top of the solid to ensure complete vaporization of the alkyl halide. The solid was pretreated at 150 °C for 2 h under a flow of He to remove water from the structure. The syringe pump delivered 1 mmol alkyl halide in 4.8 min into a He flow of 3.01nL/min. This combination of parameters gave a contact time on the solid of 0.20 seconds. The product mixture was collected at liquid N2 temperatures in a quartz trap at the bottom of the reactor for 30 min to ensure collection of all reaction products. The products were diluted with toluene and analyzed by CC. C. Physical Methods. 1. X-ray Diffraction. Powder x-ray diffractograms were recorded using either a Siemens Crystalloflex-4 or Philips x-ray diffractometer. Both instruments utilized Ni-filtered Cu K“ radiation. Some samples were prepared by placing approx. 1 mL of a l- 5% suspension of the solid to be examined on a microscope slide and allowing it to dry. The slide was then placed in the goniometer for examination. For samples such as DDS- exchanged species that could not be resuspended or would not form a film which adhered to the slide, a different technique was used. Double sided tape was placed on the microscope slide. The powder was applied to the tape and then pressed on‘with another slide. 'The second slide was 42 removed, leaving a pressed powder sample. The first tech- nique, which gave films with the LDH platelets preferen- tially oriented parallel to the slide, was the preferred method of preparation since the °001 reflections were accentuated in these diffractograms. Typically, the sam- ples were scanned from a 2 value of 2° to 30°. The inten- sity of the signal was recorded on a strip chart recorder. The 2 values were converted to d-spacings using a standard chart based on Bragg's equation for Cu K3 radiation (A =1.5405 A). 2. EPR Studies. EPR spectra were recorded using a Varian E-4 X-band spectrometer. The synthesized materials containing the - PADS spin probe were placed in a quartz tube while damp and forced to the bottom of the tube by a Teflon rod. EPR spectra were recorded while the sample was damp and after drying for various periods of time. Standard pitch served as a reference material (g=2.0023). 3. Infrared Studies. Infrared spectra were recorded by use of a Perkin Elmer model 457 or model 533 spectrophotometers. Both were double beam instruments with diffraction gratings for wave- length selection. The spectra were recorded from 4000 cm"1 to 250 cm'1 on samples prepared as KBr pellets. 4. Gas Chromatography. Reaction product mixtures were analyzed by using gas liquid chromatography (GC). Some samples were analyzed by 43 use of a Varian Associates model 920 CC with thermal con- ductivity detection. A 10' x 0.25" 5% SE-30 on Chromosorb- W column separated the components of the sample. The chromatograms were recorded by use of a Linear Instruments model 252A recorder-integrator which was calibrated with standand mixtures to insure accuracy of integration. All reaction mixtures involved in kinetic studies were analyzed by use of a Hewlett-Packard model 5890A capillary GC using a 60'm x 0.25 mm dimethyl polysiloxane (Supelco SPB-l) column and flame ionization detection. The chromatograms were recorded and analyzed by a Hewlett-Packard model 3392A programmable plotter/integrator which was capable of custo- mized data handling. All peaks were identified by comparison of retention times with authentic samples of reaction mixture components. 5. Surface Areas. Surface areas were determined using a Perkin-Elmer- Shell model 212B sorptometer with N2 as the adsorbing gas. These surface areas are referred to as B.E.T. surface areas, in honor of the developers of the technique, Brunauer, Emmett, and Teller. The samples were either outgassed for l h at room temperature under dry Ar or at 150 °c for 2 h under Ar. 6. Adsorption Studies. The adsortion capacities of Zn-Cr LDHs were measured by 103 using a McBain balance arrangemen of quartz springs in a vacuum system. The vacuum line was constructed of quartz. 44 Teflon stopcocks and valves were used to keep the system grease-free. Vacuum was achieved with rotary pumps and a CVC model PVMS-3l diffusion pump. Pressure was measured with a Datametrics model 1400 electronic manometer. In a typical experiment, 75 mg of an LDH was placed in a quartz bucket and suspended from a pre-calibrated quartz spring. The position of the bucket was measured using a telescope on a height-calibrated mount. The portion of the vacuum system containing the solid was then evacuated. Samples were either evacuated to le0’3 torr at room tem- perature before adsorption, or heated to 100 °C for 90 min under vacuum to remove water from the structure. Nhile the samples were being treated, the solvent to be used as adsorbate was degassed in a vacuum isolated from the solid samples. Heat treated samples were allowed to equilibrate for l h before adsorption studies began. The position was noted so that the final weight of the solid material could be calculated and used in subsequent data treatment. The adsorptions were carried out at 2 temperatures, 22°C and 0‘WL An ice bath was employed to cool the sam- ples to 0‘NL These samples were allowed to equilibrate for l h before beginning adsorption. The adsorbate was also maintained at 0 °C to minimize condensation on th solid at high values of p/po. After all temperatures were equilibrated the system was isolated from the vacuum pumps. A small quantity of adsorbate was leaked into the system from a smaller manifold. The- system was allowed to 45 equilibrate approx. 5 min, by which time the sample had stopped adsorbing. The position of the bucket was recorded to a precision of 0.05 mm by using the telescope. The pressure was also recorded. The procedure was continued up to a value of p=po for the adsorbate at 0 °C. Chapter 3. Results and Discussion. A. Properties of LDHs. 1. Exchange and Swelling Properties. The compounds [ZnZCr(OH)6][Cl'2H20], [Ca2A1(OH)6] [l/2CO3'2H20] and [LiAlZ(OH)6][0.5SO4'2HZO] were synthesized. The carbonate form of the Ca-Al LDH was formed in the initial synthesis due to improper precautions against C02 absorption. Proper precautions were taken in subsequent syntheses. The powder x-ray diffractograms of these materials are shown in Figure 8. As can be seen, the patterns are nearly identical, as the structural components of these species were very similar. When anion exchange reactions were attempted with these species, the Zn-Cr LDH exchanged more readily than the other two species. This difference may have been due to some adventitious carbonate in the structure of the Ca-Al and Li-Al LDHs, even when precautions against C02 adsorption were taken. It has been reported that carbonate forms of LDHs are difficult, if not impossible to exchange. This difficulty is due to the good fit of carbonate in the gallery and the -2 charge. The acidic conditions under which the Zn-Cr LDH was formed preclude the uptake of C02 by the reaction mixture. As can be seen from the diffractograms in Figure 8 the 46 47 1 i a 1 Leax 1 HL4SA 'Lssi ' flpflWNP 5.271 352A 1_ 1 j ( . ‘ V I C 0 378A ./ \ k'\de\Aw-v*Ffi- fl v 25 20 15 10 . 5 degrees 29 Figure 8. Powder x-ray diffractograms of (a). [ZnZCr(0H)6] [C1'2H20]; (b). [LIA12(OH)6][0.5304'01‘120]; and (C). [C32A1(0H)6][0.5C03'2H20]. 48 peaks from Zn-Cr LDH were broader than those of the other species and was partially due to the method of sample preparation for x-ray examination. The Ca-Al and Li-Al species formed oriented films on microscope slides. This preparation tends to minimize layer disorder that can broaden diffraction peaks. An x-ray diffractogram of a Ca- Al sample prepared by the pressed powder method is shown in Figure 9 along with that of the oriented film sample. Some broadening of the °00l peak is observed in the pressed powder sample. The Zn-Cr LDH did not form oriented films and therefore was examined in the diffractometer as a pres? sed pdwder sample. 'This random orientation of platelets probably helped broaden its peaks. Another factor which determines diffraction peak width is particle size. The synthesis of the Ca-Al and Li-Al LDHs featured a hydrothermal aging step. Aging increased particle size and in turn lead to sharper diffraction lines. The synthesis of the Zn-Cr LDH included no such aging step. Indeed, other workers31'32 have found that the particles of this compound cannot be made larger through aging. The small particle size, along with the absence of carbonate, probably contributes to making Zn-Cr LDHs the most facile anion exchangers. In a previous section the swelling behavior of alkyl sulfate intercalates of Zn-Cr LDH was discussed. The swel- ling behavior is a general property of the LDHs tested. Anion exchange reactions were carried out by using a 0.5 M 49 7.56 A (I 3.78 A i l i M i '3 Ii ’1 ! 1 -¢~4rw 25 20 15 10 5 degrees 20 Figure 9. Powder x-ray diffractograms of [Ca2A1(OH)6] [0.5CO3°2HZO] (a). pressed powder; (b) oriented film. 50 solution of NaDDS with the three LDHs. X-ray diffrac- togranusof the three intercalates are shown in Figure 10. Again the diffractograms of different LDHs with the same gallery anion are practically identical. The small dif- ferences (dual=22.6—24.5 A; gallery heights=l7.6-19.5 A) were probably due to instrumental uncertainties at low angles and to so-called kink block structures.1°4 These result from differences in the carbon chain arrangement when long chain hydrocarbon species are intercalated in layer structures. When the three DDS-LDHs were immersed in n—octanol, all were swollen as had been previously reported for the Zn-Cr LDH. The gallery heights ranged from 30.4 to 33.9 A. The x-ray diffractograms for the swollen species are given in Figure 11. There was a dramatic change in surface properties of the three LDHs upon exposure to NaDDS solutions. The DDS’ seemed to have a strong affinity for the surfaces of the LDHs. The affinity was first indicated in washing the materials after exchange of DOS“ into the galleries. Removal of the excess electrolyte by centrifugation nor- mally required 5-10 washings of the material after exchange. After exposure to DDS', however, 15-20 washings were required before foaming of the wash solutions ceased, indicating that excess NaDDS had been removed from the system. Upon drying, the DDS-exchanged materials were found to be hydrophobic to the extent that they floated on water. This behavior was completely opposite that of 51 22.6 A 23.9 A b 8.2 A 12.1 A C 15 10 5 2 degrees 25 Figure 10. Powder x-ray diffractograms of the dodecyl- sulfate-exchanged forms of (a). Zn-Cr LDH; (b). Ca-Al LDH; and (c). Li-Al LDH. 52 3L9 A O 30.4 A 10.53 13.3A 10 degrees 29 Powder x-ray diffractograms of n-octanol—swollen DDS—exchanged forms of (a). Ca-Al LDH; and (b). Li-Al LDH. Figure 11. 53 inorganic-anion exchanged LDHs. The pristine materials were all found to disperse in water, and also to be wetted by toluene and alcohols. When placed in the presence of both toluene and water, the water was preferred. The DDS materials dispersed in organic solvents but not in water. This behavior persisted even upon repeated washing in etha- nol and acetone, sonification and repeated washing to remove the DDS from the surface. The tenacity with which the DDS anion was attached to the surface of the LDHs is not unprecedented. It has been reportedla4 that DDS“ was strongly bound to -alumina through interaction of the -SO3' tail of the anion with the hydroxide surface. The "fit" of the -SO3' group to the surface groups was found to be very good, encouraging a strong electrostatic interaction. LDHs have a similar ar- rangement of hydroxide groups on the surface, plus a net positive charge in the layers not completely compensated by the gallery anions. Therefore, an even stronger attraction for DDS' probably exists here than in the alumina. Additionally, [ZnZCr(OH)6][Cl'2H20] was exposed to a 0.1 M solution of NaDDS for 30 seconds, then removed and washed as if an exchange had taken place. X-ray diffraction indicated that no exchange had occurred, but the material exhibited the same hydrophobic properties as the exchanged material had, indicating that the source of the change in properties was due to surface interactions. The IR spectra of pristine and DDS-exposed materials shown in Figure 12. 54 4000 3500 3000 2500 2000 P 2000 1800 1600 1400 1200 1000 wavenumber (cm'l) Figure 12. Infrared spectra from 4000 cm‘1 to 1000 cm'1 of [ZnZCr(OH)6] [Cl'nHZO] (a). pristine material; and (b). exposed to 0.1 M NaDDS solution. 55a The presence of DDS is indicated by the C-H stretching bands near 3000 cm-1 and bands at 1100 cm'1 and 1175 om'l due to S-O stretches of bound DDS. These IR data compare with those given for DDS on alumina, with S-O bandsat 1200 cm'1 and 1240 cm’l.104 The shift to lower energy of the S- O stretch in the present study is probably an indication of stronger binding of the -S03 group to LDHs. The authors also stated that after the first layer of DDS' was attached to the surface, additional layers were adsorbed by cosol- vation of the alkyl groups of bound and unbound anions. This phenomenon could account for the large number of washings required to remove excess NaDDS from an exchange mixture. The implications of this type of interaction will play a role in the interpretation of EPR data to be presented. 2. EPR Studies. EPR spectroscopy has provided a probe into the mobility .of gallery cations and anions of layered compounds. For smectite clays, paramagnetic metal cations such as Cu(II) and MMU) have been used as probes,l°5'106 as well as the protonated form of the radical 4-amino-2,2,6,64tetramethyl- piperidine-N-oxide (HITEMPO).1°7 The PADS dianion has been used as a probe for intersalated smectites. The mobilities of these species in the galleries has led to work in which gallery cations in clays were used as catalysts,108 and the triphase catalysis using intersalated clays outlined earlier. 55b In the present work, attempts were made to use the PADS dianion as a spin probe in LDHs. If LDHs were to be effec- tive as solid reagents or solid phase transfer catalysts for nucleophilic displacement, freedom of movement of the gallery anions would be necessary. The nitroxide spin probes in general have been observed to tumble rapidly enough at 25 °C to completely average the anisotropies in g tensors and hyperfine values. As the tumbling of the anion in solution is restricted, the spectrum becomes increasing- ly anisotropic. The two limits of anion mobility give the two spectra shown in Figure 13. The spectra obtained in this study were analyzed fol- lowing an approach used by biophysicists studying mobility of spin probes in muscle tissue under conditions of varying 109 water content. The relationships for the spin corre- lation times are shown in Equations 1 and 2. In these ’13.: 0.65W0(R+-2) (2) R+ = [(hg/h+1)1/2] + [(hg/h-1)l/2] (3) equations, and are tumbling times in nanoseconds, W9 is the linewidth of the central peak in G and H0 is thelmag- netic field in G of the central peak. The symbols 110' h+1 and h_1 represent the heights of the middle, low and high field peaks, respectively; The correlation time» c' was the average of the two values from Equations 1 and 2. The equations used to obtain the correlation times are strictly valid only for isotropic rotation, but they are useful for (I o—N( :3 b 20 GAUSS F—H H —'- Figure 13. EPR spectra of KZPADS (a). 10"5 M solution in DMSO; and (b). frozen in D20. 57 comparative purposes for studying anisotropic rotation at surfaces. The first attempt to prepare samples for EPR analysis was made by trying an anion exchange reaction of a 0.1 M solution of KZPADS into [ZnZCr(OH)6][C1'2H20]. The EPR spectrum of the resulting material is shown inFigure 14. As can be seen, if any PADSZ‘ were present, its signal was swamped out by the huge peak due to Cr(III). This spectrum is identical to one of the pristine Zn-Cr LDH. It was obvious from this experiment that an LDH with only diamag- netic constituents in the octahedral layers was need for EPR experiments. The [LiA12(OH)6][Cl'nH20] synthesis with the PADS-doped LiOH solution was undertaken to provide just such a material. The EPR spectra of the material obtained are shown on Figure 15. Analysis of the top spectrum taken while the material was still damp gave a correlation time of 1.8 ns. This value was essentially the same as that observed in the doped hectorite intersalates maintained a 98% humidity. Drying the LDH material in a vacuum dessi- cator for 411and overnight produced more anisotropic EPR signals and correspondingly longer correlation times of 12.0 and 40.1 ns respectively. That the figures for damp material compared favorably with those of intersalated hectorite seemed to be a good sign that reaction of the gallery anions could take place in the LDHs as they had in the intersalates. 58 0' 2.00 Figure 14. EPR spectrum of [ZnZCr(OH)5] [Cl'nHZOI- Frequency = 9.14 GHz. J fl 4 —_—’ EPR spectra of PADS-doped [LiA12(OH)6] [Cl'nHZO] (a). damp material; (b). after 4 hr in vacuum dessicator; and (c). after 12 hr in vacuum dessicator. Frequency = 9.14 GHz. Figure 15. 60 The numbers were not reasonable, however, when the basal spacings of the species involved were considered. The °00l spacing of the intersalated hectorite was 31.5 A, which corresponded to a gallery height of 6 A when the thicknesses of the main clay layer and two layers of com- plex cation were subtracted. That a material with a basal spacing of 7.7 A (gallery height= 2.9 A) could exhibit the same degree of tumbling freedom seemed unreasonable. There was no doubt that the PADS species was present in the Li-Al LDH system since the blue color persisted through numerous washings and the material gave an EPR spectrum. The answer to the problem lies in the structure of PADS, shown in Figure 13. It contains -SO3' groups of the same orien- tation as those found in DDS“. The PADS dianion would bind more strongly than the DDS due to greater electostatic attraction. As the material was dried, surface water was removed, restricting the freedom of movement of the PADS on the surface. Even if some of the spin probe made it into the galleries, its signal was overpowered by that of the surface species. To try to alleviate this adsorption problem, another synthesis of the Li-Al material was attempted. In this case, the LiOH solution was not only doped with PADS but also was 0.1 M in NaDDS. It was hoped that the greater concentration of DDS would block adsorption of PADS on the surface by blocking edge sites. In this way any PADS present in the product would be in the galleries. The EPR 61 spectra of the resulting materials are given in Figure 16, and a comparison of the correlation times obtained for this sample and the previously—prepared sample is given in Table 9. As can be seen the correlation times for the two simi- larly treated products were practically identical. Also, x-ray diffraction of the samples showed that Cl’ from the AlCl3 solution had been intercalated in both cases. The DDS concentration was not high enough for the larger DDS anion to be favored over the smaller C1". The dianionic PADS should have been preferred over Cl", since dianions have been found to possess a higher affinity for LDHs than 1 Even with the surface of the LDH covered with monoanions. DDS the surface concentration of PADS was larger than that of the gallery, if any were exchanged at all. Later experiments cast further doubt on the validity of the interpretation of this work with Li-Al LDHs. When Table 9. EPR Correlation Times for PADS-doped LibAl LDH Materials. Materiala Drying Timeb c(nanoseconds) LieAthI damp 1.8 Li-Al-Cl 4 hr 12.0 LihAl-Cl 12 hrc 40.1 Li-AleDDS damp 1.8 LiFAlflDDS 4 hr ' 11.8 Li-Al-DDS 12 hrc 43.0 aPrepared by coprecipitation. Li-Ala-DDS material had LiOH gynthesis solution 0.1 M in NaDDS. In vacuum dessicator. cOvernight. 62 o- 2.00 a i \ b c zoo I! Figure 16. EPR spectra of an Li-Al LDH synthesized in the presence of NaDDS and KZPADS (a). damp material; (b). after 4 hr in vacuum dessicator; and (c). after 12 hr in vacuum dessicator. Frequency = 9.14 GHz. ' 63 freshly—prepared Li-Al LDH was placed in dialysis tubing to be washed free of excess electrolyte, only Al(OH)3 remained, as shown by the x-ray diffraction patterns in Figure 17. The diffractogram of fresh material shows the expected LDH pattern. When this material was dialysed, the second diffractogram results. Elemental analysis confirms that >98% of the Li+ had been removed. Material which had been hydrothermally treated after synthesis did not lose Li+ upon dialysis. Gibbsite intercalated with lithium salts have been prepared110 by heating LiX solutions with gibbsite. A powder x-ray diffractogram of a material prepared in this manner from gibbsite and LiCl is shown in Figure 18. It looks very much like the patterns obtained from LDHs. The Li+ in such species is removeable by washing with fresh water. From this data it was concluded that upon coprecipitation of Al(III) and Li(I) salts, an intercalated aluminum hydroxide is formed. When this material is treated hydrothermally, the Li+ ion is driven into the octahedral sheet. Reactions of this sort111 also occur in montmorillonite, a smectite clay mineral with an empty site in its octahedral layer (Figure 3). These later results would have raised doubts as to the validity of any correlation times obtained had the PADS been successfully intercalated. In summary, the reactions with PADS did not provide a mechanism to examine anion mobility in LDHs. A majority of the spin probe was apparently adsorbed on the surface of 7.63 A “— '11 1 9w“ Mfiwk . WNW” MflummwflNAWNTfifibflu AWMNHWNNWMAA 4.5.4.8 A "VA” 1 \NM ., W)“ m ”WWW 25 20 15 10 5 b 1’ «flwhm°v°1’°'¥‘o‘pfil~mvw\/ degrees 26 Powder xaray diffractograms of unaged Li-Al LDH Figure 17. filter—washed; (b). dialyzed until product (a). free of electrolyte. 65 7.63 A 3.80 A 4.3=4.5 A 25 20 15 10 5 degrees 20 Figure 18. Powder x-ray diffractogram of product formed by reaction of LiCl solution and gibbsite. 66 the LDH in a manner analogous. As a result of the attempts to synthesize samples of an Li-Al LDH with PADS present, the variable structure of unaged Li-Al LDHs was discovered. EPR analysis could be helpful in studying the gallery if a different probe could be used, one which did no interact appreciably with the surface, such as an anionic complex of a paramagnetic transition metal. B. Nucleophilic Displacement Reactions. 1. Reactions Under Triphase Catalytic Conditions. One of the main interests in developing the chemistry of anion exchangers was working toward an effective system for triphase catalysis. As outlined earlier, the require- ments for a successful system include anion exchange capa- city and sturdiness under reaction conditions. The LDH materials seemed to fit these two requirements. Questions remained as to the availability of the anions for nucleo- philic attack under reaction conditions and about the ability of the solid to be regenerated under these con- ditions. Also, since the nucleophile for the reaction would be in a restricted environment, the possibility of size and/or shape selectivity existed. It was postulated that access to the anions might be more restricted for larger, bulkier alkyl halides than for smaller ones. With these thoughts in mind the triphase reactions were carried out. 67 The choice of reactions to study was very large, as outlined earlier. Halide substitution reactions were cho- sen for study because of their simplicity and because of expertise developed in the group in previous work of this type.99 Iodide substitution reactions of alkyl halides were chosen as the most likely to succeed because aL.the nucleophilicity of the 1' species is greater than that of Cl’ and Br“; and b). I“ substituted Zn-Cr LDH provides a larger basal spacing than Br‘ or Cl' forms and conceivably more space for a reaction to take place. The results of the reactions are shown in Table 10. These reactions were carried out for 24 h at 90°C and were vigorously stirred. The data seemed to indicate that some of the predicted behaviors were occurring. The larger Table 10. Conversion of Alkyl Halides by the Reaction Rx + Y’—9»RY + X‘ Under Triphase Consitions.a RX Y“ Solidb % Conversion n-C4898r I 1 85.4 n~C 5H1 13: I 1 54.9 nI-CS I 1 40.7 H38”}3H3)CH2CH23I I 1 30.4 n-C4flgBr Cl 1 9.2 naCGRl Br c1 1 1.6 D“C4H9 I I 2 88.6 “*CSHIIBI I 2 56.3 3Reaction conditions: reaction time: 24 hr; temperature: 90 oC; Y‘/RX mole ratio: 6.0; RX/solid mole ration: 20.0; [Y ] 2. 0 M; 2. 0 mL toluene; 3. 0 mL H O. b30113 1: [2n2 Cr(OH) III 2. 3H 0] when Y = I; [Zn Cr(OH)6 1 [C1'2HZO] when Y' 81. Soli 2: [ZnZCr(OH)5][DD§ nHZO]. 68 alkyl halides were reacting more slowly than the smaller 1- bromobutane. Also, the branched alkyl halide, l-bromo-B- methylbutane reacted more slowly than the straight-chain counterparts. The conversion of RBr to RCl was slower than conversion of RBr to RI. The low conversion for the chlo- ride substitution reactions could be attributed to the energetic requirements for the replacement of intercalated C1' by Br“, a process requiring lattice expansion of the LDH. All of these observations were interpreted as evidence that anions for reaction were coming from the galleries of the LDH. However, the above data were open to another interpre- tation, one that was shown to be correct when blank experi- ments were carried out. The same reactions were performed with no solid in the mixture, but otherwise duplicating the conditions of the triphase reactions. Table~ll gives the results of the blanks. A comparison of the numbers in Tables 10 and ll shows that the conversions are virtually identical. The conclusion drawn from these data was that the reactions were taking place much more rapidly at the organic-aqueous interface than at the solid. The selec- tivity toward smaller alkyl groups can_be explained on the basis of decreased miscibility of longer alkyl groups with the aqueous phase. The observed selectivity toward iodide substitution reactions over chloride substitutions appar- ently arises from the greater miscibility with organic solvents of iodide over chloride. 69 Table 11. Results of Blank Reactions for Halide Exchange. Rx I“ % Conversion nhC4H98r I 86.2 naC H Br I 57.9 CH H}&H3)CH2CHZBr I 31.2 0" 489Br C]. 8.9 aReaction conditions: reaction time: 24 hr; temperature: 90 oC; Y'/RX molar ratio: 6.0; [Y-laq: 2.0 g; 2.0 mL toluene; 3.0 mL H20. The interpretation of these results was that the:sur- face of the LDH was too hydrophilic to allow interaction of the organic reagent with the gallery anion under the reaction conditions. Since it was known that exposure to NaDDS solution made the surface more hydrophobic, some of the Cl' form of Zn-Cr LDH that had been so treated was used in a series of triphase reactions. It was thought that the I“ would quickly displace the Cl' under reaction conditions and that iodide substitution would proceed more readily on the hydrophobic surface. The results of the reactions are given in Table 10. The conversions with the DDS-treated materials were slightly higher than those of the plain 1" form. However, x-ray diffraction showed that even after 24 h at 90 °C the Cl' had not been displaced from the gallery. Therefore none of the reaction was taking place with the gallery anions; indeed, none of the iodide even made it into the galleries. The increase in conversion when the DDS-treated LDH was used may have been due to increased interfacial surface area. The surface area of the DDS- 70 treated solid was found to be 15.0 m2/g versus 7.7 mz/g for the pristine material under air-dried conditions. The result of this experiment was to transform a material that was too hydrophilic for access by organic reagent into one that was too hydrophobic for interaction with the aqueous phase in the reaction mixture. Under triphase reaction conditions with.the untreated LDH the anions were unavailable for reaction with alkyl halides. The surface properties of the LDH seemed to be responsible. In order to test this hypothesis, another series of reactions was undertaken. In these reactions, the components of the triphase mixture were removed one or two at a time to study the effect on conversion. The conditions of reaction and conversions are listed in Table 12. These reactions were carried out under stoiciometric conditions. A dramatic improvement in conversion was noted when the water was removed from the system (11.5% vs.76.6%). This seemed to confirm that water on the sur- face was blocking access of the organic substrate to the anions. A series of experiments characterizing the avail- ability and reactivity of gallery anions in relation to the” surface properties of the LDHs proceeded from this point. 2. Stoichiometric Biphase Reactions in Toluene. The probe reactions used in this portion of the research were the same as those used in the triphase reactions. Toluene was chosen as the organic solvent to 71 Table 12. Conversions of Nucleophilic Displacement Reactions of C4H9Br Under Modified Conditions.a Run H20 toluene 1 mmol NaI 1 mmol LDHb % conversion 1 yes yes yes no 51.6 2 no yes yes no 0 3 yes yes no yes 11.5 4 no yes no yes 76.6 ‘31.0 mmol C H Br present initially in all runs. 3.0 mL H20, 3.0 mL toluene present where indicated. Reactions run 24 hr gt 90 °C. [ZnZCr(OH)6][I'2.3H20]. keep some continuity with previous work in the triphase area. The experimental procedure was designed to allow the conversion to product to be followed over the course of the reactions. Analysis of the conversion data would provide information about the kinetics of reaction and possibly a greater understanding of the mechanism of the reactions [taking place. Data treatment was simplified by the programmable plotter/integrator from which the percent conversion could be obtained directly. A flame ionization detector was used with this particular GC and meant that all conversions reported by the integrator were weight percent values. The weight percentages were converted to molar percentages by a routine on a programmable calculator before any subsequent data treatment. Air-dried (AJLJ [ZnZCr(OH)6][I’2.3 H20] was used in the initial studies of displacement reactions. The data collected from these reactions fit a first order kinetic expression with a high degree of correlation. From plots 72 of '1n(fRBr) vs. time, 1where fRBr=the fraction alkyl bromide remaining unconverted, pseudo-first order rate constants were determined. Several kinetics plots are shown in Figure 19 with the values of Rob and their 90% s confidence limits given in Table 13. The data in this table indicate that the the values of kob did not vary to s a significant extent with the length of the alkyl chain of RBr. The branched l-bromo-3-methy1butane was the slowest of the reactions, suggesting that there was some steric hindrance in the reaction mechanism. In the triphase systems, reactions on the solid were seemingly inhibited by the presence of water in the system. Since A.D. Zn-Cr LDHs had water on their surfaces,32 it seemed reasonable that removing this surface water would accelerate the bdphase reactions. Previous studies32 had indicated that all the water could be removed from both the surface and galleries of the I' Zn-Cr LDH by heating the material to 150 0C under argon. Treatment at this Table 13. Calculated Pseudo First Order Rate Constants, kobs' for Iodide Substitution of Alkyl Bromides Over A.D. [ZnZCr(OH)6][I'nHH20]. RBr 105 kobs(s“1)b labromobutane ’ 3.79 + 0.28 l-bromopentane 3.66 + 0.59 lhbromo-3-methylbutane 2.08 + 0.14 lbbromohexane 2.89 + 0.34 lbbromooctane 3.03 + 0.21 aReaction conditions: temperature: 90 0C; 0.453 g (1.00 mmol) [ZnZCr(OH)6][I'2.3HZO]; 1.00 mmol RBr; 3.0 mL toluene; gampled every 10 min for 1 hr. . + 90% confidence limit based on linearity of data fit. 73 0.3:; 0.2: I 0.1 ‘ O A O 00!] 1— ; : ; g t 0 600 1200 1800 2400 3000 3600 time (s) Figure 19. First order kinetic plots for iodide substi= tution of alkyl bromides over [ZnZCr(OH)6] [11.31120]: labromobutane; lhbromohexane; and l-brom0h3kmethylbutane. 74 temperature apparently did not cause any dehydroxylation of the octahedral layers. With these thoughts in mind, some of the I" Zn-Cr LDH was treated at 150 °C for 2 h under Ar. Part of the IR spectrum of this material, along with that of an A.D. sample, is given in Figure 20. A comparison of the region from 4000 cm"1 to 1000 cm’1 shows the virtual disappearance of the bending modes of water around 1600 cm‘ 1 after heat treating. The OH-stretching region of the spectra, due mainly to the hydroxide groups of the octahe- dral layers, remains almost identical upon heat treatment. When this OJL.material was used as the iodide source in the conversion of alky bromides, the kinetics results were similar to those obtained with the A.D. material. The data could be fit with a high degree of linearity to a first order rate expression. The values of for Rob ob- s tained for CLD. materials are given in Table 14 along with their 90% confidence limits. The values obtained using OJL LDH are approximately 3-4 times those of their AJL Table 14. Pseudo First Order Rate Constants, kob , for Iodide Substitution of Alkyl Bromides Sver O.D. [ZnZCr(OH)6][I' 01120].a RBr 10S kobs (s'l)b labromobutane 13.2 t 0.92 lebromopentane 12.7 i 0.43 lhbromoa3amethylbutane 9.40 1.0.26 lhbromohexane 12.1 i 0.23 1rbromooctane 9.39 i 0.11 aReaction conditions: 90 0C; 1.00 mmol RBr; [Zn Cr(OH)6] [I]; 3.0mL toluene; sampled every 10 ndJi for l r. b+ 90% confidence limits based on linearity of data fit. 75 L r 1 4000 3500 3000 2500 2000 { L J I 1 ' L F t r ' ' 2000 1800 1600 1400 1200 1000 A. wavenumber (cm'l) Figure 20. IR spectra from 4000 cm'1 to 1000 cm‘1 of (a). [thCI(OH)6][I°2.3HZO]; (b). [ZnZCr(OH)6] [I'NGH20]. 76 counterparts. The increase in surface area upon drying the material was only approximately 25% (7.7 mz/g A.D. vs. 9.9 m2/g O.D.), so increased surface area could not adequately account for the 3-4 fold increase in rates. This seemed to confirm the proposition that removing water from the sur- face did improve access to the gallery anion. One of the experiments carried out to characterize the reactions measured the rate constants at different tempera- tures in order to calculate an Arrhenius activation energy 112 and thermodynamic parameters of activation. The plots of kinetics data used to calculate activation parameters for l-bromobutane are shown in Figure 21. The iodide substitution of l-brompentane and 1-bromo-3-methylbutane were also carried out at different temperatures. The plot of In kobs v3.1JW'for 1-bromopentane is shown in Figure 22. The values of the activation parameters for the three alkyl halides are given in Table 15, along with some values Table 15. Activation Parameters for Some Alkyl Halide Iodide Substitutions at 90 °C.a RX 1“ source AH‘(kcal/mol) AS’(cal/mol°K) ref n~C4HgBr LDHb 13.1 1 1.0 ~40.6 1 3.0 this work nacsallar LDHb 14.5 1 1.3 -36.8 1 3.7 this work isoscsHllBr LDHb 15.1 1 0.6 -35.7 1 2.7 this work CH3Br solution 15.8 1 1.0 a7.6 1 3 113 CH CHZBr solution 16.3 i 2.0 516.4 i 6 113 n‘33H7C1 solution 20.3 515.4 114 3+ 90% confidence limits for values from this work; of limits not stated in referenced material. O.D. [ZnZCr(OH)6][I]. source Figure 21. 77 L l f r ' ' V 600 1200 1800 2400 3000 3600 time(s) First order plots for iodide substitution of lubromobutane over CLD. Zn/Cr/I LDH: (a). 129 °C; (b). 100 °C; (c). 99 °c; (d). 80 °C; and (e). 50 0C. 78 144» ‘14 e 1 V V '_1 0.0 1.0 2.0 3.0 103(1/T)(K‘1) Figure 22. Arrhenius activation energy plot, 1n kobs vs. l/T, for iodide substitution of labromopentane over O.D. Zn/Cr/I LDH. 79 calculated for similar reactions in solution.]'l3'114 The values.of AH’ foundxfor the reactions over solids and in solution were similar. This similarity was inter- preted as evidence that the reactions over the solid were following a similar mechanism to that in solution, 8N2 displacement. The negative values of AS‘ indicate a coming together of particles in the transition state, supporting the picture of a nucleophilic attack. The magnitude of entropy change was significantly larger for the reactions over solids as compared with their homogeneous cogeners. This difference was interpreted as evidence for the reaction taking place on the surface of the solid. The arranging of an adsorbed molecule into the special orien- tation required for reaction on the solid would conceivably be a more ordered state than the coming together of two particles in solution. The activation data argued against another mechanism that couldbe postulated. That was one in which an equili- brium concentration of halide ion existed in the toluene solution and was responsible for the nucleophilic displace- ment reactions occurring. If this mechanism were operating, it seems that the ASt data would be approxi- mately the same as that for the homogeneous reactions. Instead, the values are much more negative. The data from gas-solid reactions also argue against a soluble nucleophile mechanism. As can be seen from the conversions for these reactions listed in Table 16 the 80 Table 16. Conversions of Gasssolid Iodide Substitutions of 1~Bromobutane Over [ZnZCr(OH)6][I].a Temperature(°C) % Conversion 90 38 130 64 140 74 150 80 aReaction conditions: He folw rate: 3.0 mL/min; contact time: 0.20 3; total C4H9Br: 1.0 mmol; 1.0 mmol LDH treated 2 hr at 150 0C in He flow before reaction. reaction proceeds in the gas phase from 90 °C to 150 °C. That the reactions proceeded in the gas phase at all where there was no solvent present suggests that the soluble nucleophile mechanism was certainly not necessary'to ex- plain the conversions in the solution phase. Gas-solhfl chloride substitution reactions of lsbromo- butane were also performed. The conditions employed were identical with the iodide substitutions except that [ZnZCr(OH)6][Cl'2HZO] was the solid phase in the reactor. The conversions are given in Table 17. The values are very close to the iodide substitution results. Chloride substi- tution of l-bromobutane under solution-solid conditions Table 17. Conversions of GaseSolid Chlorinations of lvBromobutane Over [ZnZCr(OH)6][Cl].a Temperature(°C) : % Conversion 90 33 130 60 140 71 150 80 aReaction conditions: He flow rate; 3.0 mL/min; contact time; 0.20 8% total C H9Br: 1.0 mmol; 1.0 mmol LDH treated 2 hr at 150 C under e before reaction. 8] gave a conversion of approximately 1% after 1 h, compared with 30-35% iodide substitution conversion. When fluoride substitution of 1-bromobutane with [ZnZCr(OH)6][F] was attempted under solution-solid and gas-solid reaction con- ditions, no conversion was noted. The significance of these data to the proposed mechanism of reaction mechanism will be discussed in a later section. From this evidence, then, it appeared that the mechanism was one of adsorption of RBr onto the solid, followed by the rate determining step of nucleophilic dis- placement, then desorption of product. This chain of events is illustrated by Scheme 2. Kads kobs Kdes ____> __—_\. ——.L RB“:soln‘ RBrads< RIads ‘ RIsoln Scheme 2 The questions that remained at this point concerned the relative availability of all the anions in the solid and the importance of the adsorption properties of the solid to the course of the reaction. These problems will be addressed. In the reactions at 90 °C of alkyl bromides with toluene as the solvent, pseudo-first order kinetics behavior was observed for the first hour of reaction. When reactions were carried out for longer periods of time, deviations from first order behavior occurred. The first order plot for extended iodide substitution of l-bromo- butane is shown in Figure 23. The graph clearly shows the 82 . o o o 0161' . o O A 0.4+ 1.: CD a: “-4 o c 4, H .I 0.2«r 4» 0.0 _$ i ‘: +f : f + j' 0 1000 2000 3000 4000 5000 6000 7000 8000 time(s) Figure 23. First order kinetic plot for the extended iodide substitution of l-bromobutane over OJL Zn/Cr/I LDH. 83 deviation from first order behavior. The data from reactions carried out a higher temperatures also sheds some light on the situation. An examination of the data given in Figure 21 for reactions a 100 oC and 120 OC reveals deviation from first order behavior at values of '1n(fRBr) of 0.35-0.40, corresponding to conversions of 30-35%. This was also the region of the plots at 90 °C at which first order behavior was no longer observed. Two explanations could be advanced to explain this behavior. The first has to do with the availability of gallery anions. It seemed reasonable to assume that as more of the I’ was consumed and replaced by Br' in the galleries of the LDH, the remaining 1' would be relatively inaccessable compared with that near the outside edges of the particles. Perhaps 35% conversion, the point at which deviation from first-order behavior became pronounced, was the value at which access to I' became the rate controlling factor of the reaction. This hypothesis was easily tested. Reactions were run at 90 °C with excess alkyl bromide, but were otherwise identical to previous reactions. The data from the reactions run a 2-, 3-, and 4-fold excess of l-bromobutane plotted as first order data are shown in Figure 24. The data was not very revealing until it was plotted in a different manner. Since the stoichiometry of the reaction was known, as was the initial ratio of RBr to I', it was possible to calculate the fraction of the initial amount of ‘1n(fRBr)4 Figure 24. 2.4” 1. 84 64> I r I, ' O I L O . I O O 0 .2A a 1' . ‘ ‘ A o A 4» ' O o 1 0 s 4" fig ? ¢ 1 0 600 1200 1800 2400 3000 3600 time (5) First order kinetic plot for iodide substitution of 1-bromobutane by O.D. Zn/Cr/I LDH at RBr/LDH ratios of I4:l; 03:1; 02:1; A1:1, plotted as functions of (a). Eln £13; (b) -1n fRBr' 85 1' remaining (fI-). When the data were plotted as -ln(fI-) vs. time, the results shown in Figure 24a were obtained. These plots were linear out1x>a consumption of 85-90% of gallery iodide when a 4-fold excess of RBr was used. These data suggested that the rate of iodide substitution was not related to what was going on in the galleries of the LDH. The linearity of the plots calculated from fI- indi- cated that the proper statement of reaction kinetics was that the reaction was pseudo-first order in fI- for the first hour of reaction. Under the original reaction con- ditions, where fI- =fRBr' this distinction was not made. It now seemed that the reaction wduld follow pseudo-first order kinetics as long as the surface concentration of RBr was kept approximately constant. The data supported the other possible explanation for the deviation from first order behavior after 30-35% conversion. This explanation had to do with the adsorptive properties of the LDH. In the first stages of the reaction, the reactant molecules were competing for ad- sorption sites on the solid with the solvent. Later in the reaction, as iodide substitution product was accumulated, it would begin to compete for adsorption sites also. If the affinity of the product for the solid were comparable to that of the reactant, which seemed likely given the similarities in structure, fewer reactive sites would be available for the reactant, slowing the reaction. Knowledge of the adsorptive properties of the solids 86 involved would be useful in understanding the reactions taking place. Therefore, adsorption studies of the Zn-Cr LDHs were undertaken. 3. Adsorption Properties of Zn-Cr LDHs. In general the affinities of solids for vapors can be determined by measuring adsorption isotherms for adsorption at two different temperatures. The Clausius-Clapeyron equation is then applied to calculate an isosteric heat of adsortion, qst' as shown below: qst a -R 5(10 P) (6 (UT) 9 (4) In this equation, ais the surface coverage. From ad- sorption isotherms, pressure and temperature data are entered into the equation for points from the two isotherms of the same surface coverage. For the isotherms described here surface coverage was expressed in mmoles adsorbed per mg of solid. Physical adsorption is a spontaneous process and as such the free energy change is negative. Since the change in entropy is always positive when a gas is changed tmaa more ordered adsorbed state, the enthalpy of adsorp- tion must always be negative. The values calculated using by Equation 4 are positive for physical adsorption. The sign convention was adopted so that the numbers to be dealt with would be positive. To convert to an isosteric enthal— py of adsorption one must change the sign on qst' The adsorption of a vapor can be viewed as a two-step process, as shown ix: Scheme 3. 87 The value of AHgad is the inthalpy change that is calculated from the two adsorption isotherms by using the Clausius—Clapeyron equation. The value which has bearing A(9) ’ A(1) 'AHvap l A(1) —7 A(ads) AH ads A(9) 1 A(016$) AHgads (3 ‘qst) Scheme 3 on the reactions under study that are carried out under solution-solid conditions is Afllads' the enthalpy of ad- sorption of a liquid phase species onto the surface of a solid. The equations in Scheme 3 may be rearranged to calculate AHlad as shown in Scheme 4. 5 Am ——* Am AHvap A(9) -__) A(ads) AI"gads A(1) '—-—9 A(ads) AHlads=AHgads+AHvap Scheme 4 Adsorption studies of toluene and 1-iodobutane by [Zn2Cr(OH)6] [£‘nH20] were undertaken. These isotherms for both the AJL and O.D.Inaterials are shown in Figures 25 and 26. The isotherms for l-bromobutane shown in Figure 27 were measured for adsorption onto the Br' form since adsorption onto AJL 1' material resulted in reaction to form l-iodobutane. The solvent source was maintained at 18 °C for those isotherms measured at room temperature (22 °C) and at 0 °C for isotherms recorded at 0 °C. The po in the p/po expressions refers to the saturation pressure of the adsorbate at the temperature of the adsorbent. 88 o 6.01 o O o O .0 00 . o 4.0» ° ° 0 ‘ ° ".4 . O H o o I m .58. m I 5 2.0:: H (E) D Q. G H '0 $00 A 4 Q) .0 H O m E 6.0“ 0) E 3 H o A > 4.00 . I . I I 3 . ‘ ‘ I ' ‘ . 2.01 . . ‘ I . ‘ I .1 I... "A 0.0 ,1 * 1 ‘ i 0.0 0.2 0.4 0.6 0.8 1.0 p/po Figure 25. Isotherms for adsorption of toluene onto: (a). A.D. Zn/Cr/I LDH at .0 °C; 022 °C; (b). O.D. Zn/Cr/I LDH at 'A0 °C; I22 oC. 89 600" 4.0" 8 o 2:: O 8 , ° ° - 0‘ . ... o. . . 0 E 2.0 . .. 5 ° r-l .09 o 38 E '° Q‘ 0 a O .-4 0.0 4 ‘7 'O (D .Q :4 O 8 m 6.01 (D E :3 H A O > 4.01 2.00 . . ‘ I I .1 . I . :5. . . I .I I a.“ f A i 0.0 0.2 0.4 0.6 0.8 1.0 p/po Figure 26. Isotherms for adsorption of l-iodobutane onto: (a). A.D. Zn/Cr/I LDH at .0 0C; 022 °C; (b). O.D. Zn/Cr/I LDH at I0 °C; I22 °C. 90 6.0 O 4.0 A °. 'U o I "-0 O H . ° 9 a - .- ‘ O m 00 ' E 2.0" .0. \ 0". H ‘53 i‘ Q‘ & H V 00“ L i1 '5 CD .0 H 8 ,0 6.0» (U 0) E 5 0-1 0 > 4.0 200" ‘ I ‘ ' I I . a . ‘ . .- I ' ..I..‘A ‘ ‘ ‘ 0.“. t 4 0.0 0.2 0.4 0.6 0.8 1.0 p/po Figure 27. Isotherms for adsorption of l-bromobutane onto: (a). A.D. Zn/Cr/Br LDH at 90 °C; .22 °C; (b). O.D. Zn/Cr/Br LDH at 10 °C; I22 °C. 91 Isosteric heats of adsorption were calculated for these systems. For a given specific adsorption volume at 0 0C, a value of pressure was interpolated from the room temper- ature isotherm. This procedure produced a (p,V) point from each curve so that a value of qst could be calculated at that point. The procedure was repeated for each point on both curves to generate pairs of (p,V) data. These values were plotted as so-called heat curves as function of V/Vm, the fraction of monolayer coverage. The values of Vm' the volume of adsorbate required for monolayer coverage of the adsorbent, were calculated for each adsorbate by using the isotherms. .Adsorption isotherms can generally be broken down into three regions (Figure 28): (1) a non-linear region at low values of p/poj (2)1ailinear region at mid— range values of p/po; and (3) another non-linear region at high partial pressures. The transition point between the first two regions, the so-called Point B, has been postulated117 to be the point at which monolayer coverage is completed since the uptake of adsorbate is changing in character most rapidly at this point. Point B analysis has proved successful in calculating monolayer coverages, yielding good agreement in surface area measurements with N2 B.E.T. measurements.115' 117 The agreement was good for a variety of adsorbates on different adsorbents. The values for the present study were estimated from the iso- therms at the two temperatures studied and averaged. The heat curves are given in Figures 29-31. Plots of Afilads volume adsorbed Figure 28. 92 Point B An adsorption isotherm showing the location of Point B, the attainment of monolayer coverage. 93 101 A... (.111 7‘ A A A 50> ‘ A 001’ 1 o o 3 I e D 91 8 ‘ o as r 0 tn) . . I m . o O o -51} I o o O .01 ~10 ; : ’ t 1 : 1 fl 0.4 0.8 1.2 1.6 V/Vru Figure 29. Heat curves, qst vs. V/Vm, for adsorption of ‘AOD. .OID. toluene onto [ZnZCr(OH)6][I°nHZO] 94 . . I . . O 220 . I I I O ‘P 18‘? . I I '3' ' : d E 14 P \ H ‘0 O .2 V I u 0 “.‘. ‘ m A A A ‘ A 100 O A 1* A‘A A A A 64- ; ‘1— 1 *1 + 4 0.6 1.0 1.4 1.8 v/vm Figure 30. Heat curves, qst vs. V/Vm for adsorption of 1-iodobutane onto [ZnZCr(OH)6][I'nHZO] AA.D. OO.D. qst(kcal/mol) 96 ‘A 10-r A. A A A. ‘ A A 1‘ ‘ A I 5.. A A I I A 1 I I I I 01> . I I I . I I I '5 ‘fi VP ‘1 1 .L 1* 1 0.6 1.0 1.4 1.8 V/Vm Figure 31. Heat curves, qst vs. V/Vm, for adsorption of AA.D. 1-bromobutane onto [ZnZCr(OH)6][I°nHZO] O O.D. ‘ 1)- l60 .0 0- O. I I I I 8*. . I L ‘ ° I I I E ‘ ‘ ‘ .0 E . . I b I (U A . 3 “f I‘_ To ‘ . . A‘A A '3 I ‘ “ A‘IA P4 3: G -31 0.4 0.8 1.2 1.6 V/Vm Figure 32. Plot of the enthalpy of liquid adsorption, Afllads' vs. V/Vm for toluene adsorbed onto [ZnZCr(OH)6][I'nH20] IA.D. IO.D. 97 121+ ‘ A I I 8.» I I A I I I I 41 I ‘ I I I A . I H n 2, '0'. 1b :5 I '0 U I 1’ ' - ' g ' . I. .34 ‘IIIA~ ’. . . (U I .I H :1: Q . I -841 I I I ~12‘L I I 0 I I I I -16+r . 0.4 0.8 1.2 1.6 V/vm Figure 33. Plots of Afllads vs. V/Vm for adsorption of l-iodobutane onto [ZnZCr(OH)5][I'nH20] OA.D. O O.D.; and l-bromobutane onto [ZnZCr(OH)6] [I'm-120] IA.D. AO.D. vs. V/Vm are given in Figures 32 and 33. From the plots it can be seen that the values of qst are sometimes negative, giving a positive AHgad Even where‘Aflgad is sometimes is negative, the value of AHlad S 5 -positive once AHva is added in. For simple physical P adsorption, these values are not possible. When positive values for these parameters are calculated from adsorption isotherm data, it means that an activated chemisorption is taking place rather than or along with physical adsorption. Most often, chemisorption is thought of as formation of directed chemical bonds between a solid and an adsorbed species. The most common example of chemisorption are found in catalytic systems where simple gases such as H2, CO and C2H4 are involved. These gases can be chemisorbed by a variety of metals and oxide surfaces, either forming covalent bonds through interactions involving pi-electron systems or empty orbitals (CO, C2H4), or through dissoci- ation and bonding (H2),113:119 Chemisorption may also involve the formation of ionic bonds, or simply dipolar interactions stronger than the van der Waals interaction responsible for simple physical adsorption.118 Simple physical adsorption is an exoergonic process with essentially no activation energy. The process is thermodynamically controlled, so a cooler sample should always adsorb more vapor than a warmer one. A plot of the volume of toluene adsorbed onto O.D. Zn/Cr/I LDH as a function of pressure is given in Figure 34. The source of 99 A 6IG‘b 'o "-1 H 0 U) A 61 E b E 4.0" ‘ V‘ I a A H A v . . I '00, g A. . I .0 I ' 3 2.0» . ' . In I I '0 A (U I..A o l‘ 5 1 H o > 0.0 A f Ar 7L 4' v: 0.0 2.0 4.0 6.0 p(torr) Figure 34. Adsorption of toluene onto O.D. [ZnZCr(OH)6] [I'nHZO] as a function of pressure at I0 °C I22 °C. 100 the positive values ofAHgadS calculated for the adsorption of toluene onto OJL.Zn/Cr/I LDH can be seen in this plot. For the range of pressures illustrated here, the warmer LDH sample adsorbed more. Therefore, for a given value of V/Vm the pressure of toluene over the solid is greater for the cooler sample than for the warmer one. These values, when plugged in the Clausius-Clapeyron equation produce the positive values of AHgad The greater adsorption at 3' higher temperature is indicative of a chemical reaction taking place upon adsorption. In a given amount of time, a reaction taking place at a higher temperature will produce more product than one at a lower temperature. The product in this case is a chemisorbed molecule. A physical inter- pretation of the activation of the solid for adsorption will be suggested in a later section. Values of qst for physical adsorption are usually close to the value of AHva for a particular adsorbate. The P values of qst for l-iodobutane on O.D. Zn/Cr/I LDH are in the vicinity of 20 kcal/mol at V/Vm values near 1. Calculated values of qst of this magnitude are usually, though not always indicative of chemisorption.119 The difference between this type of chemisorption and that discussed earlier is the activation barrier. So-called activated chemisorption has au1 activation barrier to adsorption which is surmounted by significantly more molecules at 22 0C than at 0 OC. The activation barrier for the second type of chemisorption observed, like that of 101 l-bromobutane on LDH, is so low that there is no signifi- cant difference between the number of interactions which occur at 0 °C and 22 oC. and AH9 The values of AHI calculated by using ads ads isotherm data are not valid in an absolute sense when activated chemisorption occurs. A true value of the heat of adsorption and chemisorption could be obtained calori- metrically were such a device available. Some of the data from the adsorption experiments is useful as given. The curves for the adsorption of l-bromo- butane and l-iodobutane on AAD. material, shown in Figures 26 and 27, represent physical adsorption for much of the curve. The calculated values of AHlad for the two species 5 are very close and help confirm the speculation about the deviation of the iodide substitution reactions away from first order behavior. Since the affinity of the solid for the two compounds is approximately the same, when the concentration of RI reaches a significant fraction of the RBr concentration, RI is able compete for adsorption sites on the solid. The accumulation of product on the surface would inhibit reaction and give the observed deviation from first order behavior. The data for the alkyl halides on O.D. material is harder to interpret. 'Thelnagnitude ofAHladS for l-iodo- butane on the O.D. material puts it in the realm of chemi- sorption and indicates a great affinity of the solid for the alkyl iodide. The positive value of AHlad for l- S 102 bromobutane indicates that chemisorption is also occurring here. Some idea of the relative affinity of the solid for these molecules can be inferred from the adsorption iso- therms. At room temperature, the O.D. Zn-Cr LDH adsorbed 2.05x10‘4 mmol/mg of l-bromobutane at p/po = 0.90 while 2.27X10'4mmol/mg of l-iodobutane was adsorbed at a partial pressure of 0.92. These data indicate that the affinity of the solid for the two compunds was about the same. The differences in reactivity between the AmD. and CLD. materials are harder to reconcile by using the adsorption data. The AND. material adsorbed more of each of the adsorbates than did the O.D. material (Figures 25-27). This observation was surprising given the larger surface area of the O.D. LDH vs. the A.D. material. The so-called AWD. material which was used for adsorption studies was somewhat different from that used in the reactions. The reaction material was air-dried solid scraped off a glass sheet and used in the reaction mixture without further treatment. The material referred to as "AJL" that was used in adsorption studies was outgassed at room temper- ature under le0'3 torr before adsorption commenced. This treatment removed loosely-held surface water that would otherwise have volatilized during adsorption measurements, resulting in inaccuracies in pressure measurements. Also, water would have condensed in the reservoir of adsorbate, causing contamination of the liquid being used. The out- gassing procedure resulted in a 4.4%_weight loss, corre- 103 sponding to loss of 1.1 water molecules per mole of LDH out of the original ZJLJZ Of the original 2.3 waters, 0.3 were considered surface-bound, with 2.0 gallery waters. Therefore, not only surface water was removed, but also some gallery waters from near the edge of the particles. The surface of the AmD. LDH used in reactions was different from that of the "AJLN solid used for adsorption studies. The difference in surface composition may have lead to results that are hard to reconcile with observed reactivity trends. The importance of these observations will be dis- cussed later in relation to a proposed detailed reaction mechanism. 4. Reaction and Adsorption Studies in Other Solvents. Studies of other solvent were undertaken to elucidate the detailed mechanism of reaction and to ascertain the importance of surface properties. The CLD. solid material was used in these studies because the solid used in ad- sorption studies could be treated to give a material very similar to that used in the reaction mixtures. Zn/Cr/I LDH treated at 100 0C for 90 min under le0'3-1Xl0'4 torr experienced an average weight loss of 7.5%, corresponding 32 This loss to 1.9 water molecules of the original 2.3. compares to a loss of 8.0-8.5% upon treatment at 150 0C under argon. The slight difference in composition was accepted because more severe heating in a vacuum caused partial dehydroxylation of the solid as indicated by color 104 change (lavender-gray to green) and weight loss. Adsorption studies of chlorobenzene, n-octane and carbon tetrachloride were undertaken in the same manner as in the experiments just described. As before, fresh samples were used for each isotherm. The isotherms were obtained and from these plots of AHlad vs. V/Vm were s generated. These plots are shown in Figures 35 and 36. The heat curve of chlorobenzene is very similar to that of toluene, showing maxima and minima at the same values of partial monolayer coverage. Octane also showed activated chemisorption behavior, though judging from the magnitude of the positive value of AH1 , not to as great an extent ads as toluene and chlorobenzene. Carbon tetrachloride showed no signs of chemisorption, based on the heat curve that was generated. An additional experiment was performed with all the samples used for adsorption experiments. After the final point on the adsorption isotherm was recorded, the solvent source was closed off. The system containing the solid was evacuated to 1X10“3 torr for 5 min. The amount of adsor- bate remaining on the solid after this treatment was calcu- lated from the position of the quartz spring. The amount of adsorbate remaining on the solid after the treatment for O.D. materials at 22 0C is given in Table 18 along with their maximum measured capacities at the partial pressures given. This crude desorption experiment provides another measure of the strength of interaction between the solid 105 12+ AHlads(kcal/mol) Figure 35. Plots of AHlads vs. V/Vm for I octane and Icarbon tetrachloride adsorption onto O.D. [ZnZCr(OH)6][I°nH20]. «J.- 106' .10 121 O. I I I ‘ I I .II ? z . A ‘ . I A ‘g‘ :. . . g 4* I‘I \. I ‘l. 3 ’ I I U .3 3 ad. 10 0-1 I <3 {1.4-0 .581» 0.4 0.8 1.2 1.6 V/Vm Figure 36. Plots of ‘Hlads vs. V/Vm for adsorption of Ichlorobenzene and Itoluene onto O.D. [Zn2Cr (0H) 6] [I'flHzO] . 107 Table 18. Maximum Adsorption Capacities and Desorption Experiment Results for Various Adsorbates on OJL [ancr (0H) 6] [I] . Ma imum Capacitya Af er Desorptionb Adsorbate (1a mmol/mg solid) (10 mmol/mg solid) toluene 4.03 1.21 lliodobutane 2.27 6.58 lubromobutane 2.05 ' 6.63 chlorobenzene 4.93 ’ 1.33 nnoctane 2.11 9.13 carbon tetrachloride 2.41 0.G5 aAt 22°C. Value corresponds to the last point on the gdsorption isotherm (p/p = 9.88b0.92). Evacuated for 5 min at 2X10‘3 torr after last adsorption point. and adsorbate. The amount of adsorbate remaining after 5' min is.dependent on the rate of desorption, the more tightly held adsorbates being removed more slowly than other more loosely bound molecules. These data will be utilized in the interpretation of the reaction results to be discussed next. Iodide substitution reactions were carried out in these solvents under the same conditions used for the reactions in toluene. l-Bromobutane was the substrate in these reactions in which LEG mmol O.D. solid and 1.00 mmol 1- bromobutane were allowed to react in 3.0 ml solvent at 90 °C. The values of kobs for these reactions are listed in Table 19. The rate constants for iodide substitutions in toluene and chlorobenzene were nearly identical. This similarity was not surprising given the similarity in adsorption iso- therms and heat curves of the two adsorbates. The results 108 Table 19. Pseudo First Order Rate Constants for Iodide Substitution of lsBromobutane with OJL Zn/Cr/I LDH at 90°C in various Solvents.a Solvent 104kobs(8“1)b toluene 1.32 i 0.09 chlorobenzene 1.34 i 0.10 neoctane 10.4 i 1.0 carbon tetrachloride 2.89 i 0.17 a1.00 mmol lsbromobutane; 1.00 mmol [ZnZCr(OI-I)6][I] (O.D.); .0 mL solvent. ‘ i 90% confidence limits. of adsorption measurements of octane could also be corre- lated to reactivity. Evacuation after adsorption showed that a much smaller amount of octane remained on the sur- face than toluene (1.31X10'5mmol octane/mg solid vs. 1.21Xl0‘4mmol toluene/mg solid), indicating that adsorbed octane was more easily removed. Since the octane was more easily displaced, the rate of reaction should be faster since more 1-bromobutane could be adsorbed on the surface. The reaction which did not fit the adsorption data was that carried out in carbon tetrachloride. There was less carbon tetrachloride left on the surface after evacuation than any other adsorbate»(5.06Xl0’6mmol CC14/mg solid). This obser- vation was consistent with the observation that no acti- vated chemisorption was noted in the adsorption of carbon tetrachloride, indicating that there was no extra energy barrier to desorption either. The equilibrium concen- tration of carbon tetrachloride on the surface should have therefore been lower than that of any of the-other solvents, producing a faster reaction. Ixafact, the rate 109a constant for reaction in carbon tetrachloride was lower than octane and not much above toluene and chlorobenzene (Table 19). An activation energy study of these reactions was undertaken to help resolve the discrepency inadsorp- tion and reaction results. Iodide substitution of l-bromobutane was carried out at several temperatures in each of the solvents. Rate con- stants were obtained for the initial linear portion of the reactions and Arrhenius plots prepared. The calculated values of the activation parameters for reactions in the different solvents are given in Table 20. The activation parameters for reactions in toluene and chlorobenzene were nearly identical, as expected. The values of 48* for reactions in toluene, chlorobenzene and octane were essen- tially the same, perhaps indicating a similar structure in the transition state of the reaction. The activation para- meter values obtained for reactions in carbon tetrachloride were different from the others. The enthalpy of activation is lower than that of the other solvents, but the entropy Table 20. Activation Parameters for Iodide Substitution of l-Bromobutane at 90 °c by 0.0. [ZnZCr (om61 [I] in Various Solvents.a solvent 4H*(kca1/mol) AS’(cal/mol°K) chlorobenzene 13.1 1 1.1 -38.7 i 3.5 n-octane 11.7 1 0.9 -38.4 i 2.9 carbon tetrachloride 8.6 i 0.8 -49.6 i 3.2 aFrom Arrhenius plots. Reaction temperatures 50-120 °C. 1.00 mmol 1-bromobutane; 1.00 mmol LDH; 3.0 mL solvent. 109b of activation is more negative, indicating perhaps that more order was required in the transition state. A detailed mechanism in which these points are addressed will be suggested after a section dealing with the thermo- dynamics of halide exchange reactions over LDHs. 5. Thermodynamics of Reaction. The thermodynamics of the halide exchange reactions can be examined by looking at the individual steps involved and analyzing the known data. The individual steps for solution-solid reactions, along with expressions for the enthalpy change associated with them, are given first, followed by those for gas-solid reactions. (a) Rx(solv)—’ RX(ads) (5) AHa = Afllads(Rx) ' AHsolv(RX) ’ AHl adS[(solvent)LDH(Y)] (6) where AHl (RX) is the enthalpy of adsorption of liquid RX ads onto the LDH, AHsolv(RX) is the enthalpy of solvation of RX in the solvent utilized, and AH1 [(solvent)LDH(Y)] is the ads enthalpy of adsorption of liquid solvent onto Y-substituted LDH. This model assumes that the surface of the LDH is initially'saturated with solvent molecules. This satur- ation a likely occurrence given the excess solvent in the system and because the solvent was always added to the solid before the bromobutane. AHb a DCX " DCY + (Aflladswy) ' AHlads(RX) (8) 110 where DCX and DCY are the bond dissociation energies of carbon-X and carbon-Y bonds, respectively. The AHlads terms must be added to account for any differences in the enthalpies of adsorption of the product and reactant. The conversion of the substrate is accompanied by the conver- sion of the LDH. (c) LDH(Y) —-) LDH (X) (9) ABC = 24.2 kcal mol-1 A’1[Ad901{LDH(X-Y)] (10) where Adaal{LDH(X—Y)} = [d091{LDH(X)} — dggl{LDHun}] and 24.2 kcal mol'1 A'1 is a calculated value for the change in enthalpy for changing the gallery height of LDl-Is.1]'8 The difference in energy for different size gallery anions results from changing the distance between the centers of positive charge in the octahedral layer and the anions. The final step in the reaction sequence is the desorption of product and solvation in the reaction mixture. 65d = -AH1ads(RY) + AHsolv(RY) + Afllads[(solvent)LDH(X)] . (12) Totalling the enthalpy terms for reaction in the solution phase gives. Aern = -AHsolv(RX) + DCX - DCY + 24.2[AdMl{LDH(x-Y)}] + AHsoivaY) + {AHlads[(solvent)LDH(X)] - AHlads[(solvent)LDH(Y)]} (13) For gas-solid reactions, steps (b) and (c) are the same as those for solution phase» The initial steps and final steps are slightly different. 111 Ana. = Augadsmm = -AHvap(RX) + Anladsmm (15) and. = -AHgads(RY) = AH (RY) - AHI vap um (17) ads The sum of the enthalpies for vapor phase reaction is Ann“. = -AHvap(RX) + DCX - DCY + 24.2[Ad001{.LDH(X-Y)}] + AHvap(RY) (18) If the enthalpies of solvation of RX and RY are assumed to be equal, and the enthalpies of solvent adsorption onto LDH(X) and LDH(Y) are close to each other, the expression for the solution-solid reaction reduces to the middle terms concerning the bond strengths and energy change of the LDH. For the gas-solid reactions a similar value of the enthal- pies of vaporization can be assumed.62 Eliminating those two terms gives the same expression as for the solution- solid reaction, namely Aern = DCX - BC! + 24.2[Adggl{LDH(X-Y)}] (19) The values of (DCx - DCY) and 24.2[Ad601{LDH(X-Y)}] for some of the reactions studied are listed in Table 21. The importance of the thermodynamics in relation to the Table 21. Values of D - DCY and 24.2 600 for Various Alkyl Halide Exchange Reactionslkx + LDH(Y)—e»RY + LDH(X).a RX Y‘ DCX - DCYb 24.2Adaol(X—Y) Total RBr I 17.0 -11.0 6.0 aAll values in kcal/mol. bBased on bond dissociation energies from reference . 112 proposed detailed mechanism will be discussed next. 5. Proposed Detailed Mechanism. A.detailed mechanisnlof the reactions under study can be proposed which is consistent with the results obtained from the present study. The simple mechanism proposed in Scheme 2 will be presented in light of all the data col- lected and discussed in some detail. The first step of the reaction is adsorption of the reactant alkyl halide. This adsorption occurs after dis- placement of a solvent molecule from the surface in solution-solid reactions, but occurs unhindered in gas- solid reactions. The next step of the reaction is nucleo- philic attack by a gallery ion. After this step, the product alkyl halide is either retained on the solid or replaced by solvent or by reactant alkyl halide. One of the questions surrounding the first step of the reaction concerns the nature of the adsorption of solvent and alkyl halide. The interaction of various adsorbates with hydroxyl surfaces has been studied in det:ail.12]"122 IR data obtained from the study of substances adsorbed on porous Vycor glass indicate that even nonpolar molecules such as alkanes interact with hydroxyl groups, shifting the O-H stretching frequency to lower energy; The magnitude of the'O-H band shift is a direct indication of the strength of interaction between the adsorbate and the hydroxyl group. The results indicated that, for a series of similar compounds, such as a aromatic compounds or alkanes, the 113 adsorbate with the lowest ionization potential caused the greatest shift in the hydroxyl band, indicating the strongest interaction. Detailed data of this type cannot be obtained for LDHs due to the presense of broad, overlap- ping bands in the'hydroxyl stretch region. The data from the other studies can be applied to LDHs. The data in the published reports indicate that n-octane should bind more strongly than carbon tetrachloride,121 as is borne out in the present study since more octane than carbon tetrachloride remained on the surface of the solid after desorption at 1X10"3 torr. Toluene is bound more tightly than either alkanes or carbon tetrachloride as would be expected from its lower ionization potential [8.82 eV vs. 11447(CC14); 9.90 eV’(octane)].62 One of the pd orbitals is also of the proper symmetry to overlap the sigma bonding orbital on the hydroxyl H-aton1(Figure 37). Chlorobenzene should bind in the same manner, but the interaction will be slightly weaker due to its slightly larger ionization energy (9.07 eV).62 The adsorption of the two aromatic species was similar according to the adsorption data from the present study. There are few data available on the adsorption of alkyl halides on hydroxyl- groups, but the polar halide group should provide a site for binding to hydroxyls. The ionization potentials for the two alkyl halides of interest indicate that l-iodo- butane should bind more strongly than 1-bromobutane (ioni- zation potentials: 10.13 eV C4H98r; 9.21 eV C4H91). 114 Figure 37. Possible bonding interactions of hydroxyl group orbitals with an aromatic system. (a). e29 + CFO“ - no net overlap; (b). e2g + H29 - p-orbital too high in energy; (c). 0.5 e2g + C7+OH - probable interaction. 115 The entire surface of LDHs is covered with hydroxyl groups, including the exterior basal planes, the edges of the particles and the surface of the hydroxide layers within the galleries. In the LDH samples that were adsor- bents in the present study, surface water and some gallery water, probably from near the edges of the particles, had been removed prior to the adsorption studies. The removal of water allows the hydroxyl groups to interact with adsor- bates. It seems likely that adsorption takes place not only on the exterior surface of the LDHs but also within the galleries. Not only are the gallery hydroxyl groups available for interaction, but the net positive charge of the octahedral layers is concentrated in the galleries by the presence of 2 charged layers. Polar groups, such as the halide groups of the reactant and product molecules, would be attracted to the positive charge and would bind through ionic-dipole type interactions. Adsorption of the halide group of a reactant alkyl halide is probably neces- sary for reaction to occur. The results of the experiments will be discussed with that assumption in mind. (aL. Under solution-solid conditions, iodide substi- tution of alkyl bromides occurs at a much greater rate than chlorinations even though chloride substitution is thermo- dynamically favored over iodide substitution (Table 21). Under gas-solid reaction conditions in the temperature range of 90 °C to 150 °C, chloride substitution and iodide substitution of l-bromobutane by LDH gallery anions give 116 similar conversions (Tables 16 and 17). Fluoride substitution of vapor phase l-bromobutane does not occur at 150 °C, however. The pattern of reactivity is related to the gallery heightq the free space between the octahedral layers, of the LDHs involved. This pattern is consistent with the assertion that the reactant bromide group must insert into the gallery in order for reaction to take place. The O.D. iodide-exchanged LDH has a gallery height of 3.4 A (6001 - 4.8 A, the approximate thickness of a brucite sheet). The chloride exchanged form has a gallery height of‘2.7 A; the fluoride LDH, 2.5 A. The diameter of the bromo group of l-bromobutane lies somewhere between the non-polar diameter of 2.28 A and the ionic diameter of 3.90 3.59 Assuming an intermediate value of 3.0 A for the bromo group diameter, the gallery heights of the chloride and fluoride LDHs would have to expand to allow the bromo group to enter. The iodide gallery would admit the bromo group without further expansion. If the estimate of 3.0 A is a reasonable one for the bromo group diameter, the energy required to expand a chloride gallery to accomodate bromo group is approximately 7 kcal/mod (24.2 kcal mol"1 A‘1 X 0.3 A). In order to expand the fluoride lattice the required 0.5 A, approximately 12 kcal/mol is required. The value of AHgad for l-bromobutane is 6-7 kcal/mol more S exothermic at reaction temperatures than AHlad The dif— S. ference is AHva of l-bromobutane which is liberated when P the vapor condenses on the solid prior to adsorption. The 117 extra energy given off upon vapor phase adsorption is enough to activate the chloride-substituted LDH in the gas- solid reactions. Recall that chloride substitution did not proceed at an appreciable rate under solution-solid conditions. The extra energy of adsorption is not enough to activate the fluoride substitution reaction since it is not enough to expand the fluoride gallery enough to(allow insertion of the bromide group. There is a parallel between these observations and the behavior of LDHs in gallery anion replacement reactions in aqueous solution. When the reactions are carried out, the platelets of LDHs do not separate and disperse as do those 25 The anion ex- of smectite clays under going exchange. change reactions apparently involve the adsorption of the exchange anion to the edge of the platelet, incorporation into the gallery and expulsion of the starting anion from the gallery. The new anion is then mixed into the interior of the gallery, away from the edge of the platelet. When the exchange anion is larger that the anion present in the galleries, exchange is accomplished only by use of a large 31'32 These con- excess of anion and high temperatures. ditions are apparently necessary to overcome the activation energy of exchange, the expansion of the gallery to accomo- date the larger anions. Exchange of smaller anions and multiply-charged anions into LDHs is more easily accomplished, the reactions proceeding rapidly at room temperature with“, moderate concentrations of anion in 118 solution. (b). The values of the entropy of activation for reactions over the solid material are more negative than those for reactions in homogenous solution. The ordering of the system for reactions is greater than that in solution. Not only are two particles required to come together to make the reaction proceed, but the arrangement of molecules needed to for the reaction on the solid is more specific than that required in solution. The arrange- ment must allow the halide group of the alkyl halide to be partially extracted while the gallery anion moves to replace it. (c). The values of‘AH' calculated for reactions over the solid are of the same order of magnitude as those of homogenous reactions, but are consistently lower. This observation reflects the assisted bond-breaking of the C-X bond in the gallery region. Part of the AH' of a nucleo- philic substitution reaction goes into breaking the bond between the attacked carbon and the leaving group. When the bond is weakened by interaction with another species the barrier to breakage is lowered. When AH‘ and AS’ are combined to give AG4 values for solution-solid iodide substitutions and the homogenous reactions listed in Table 15, the values are lower for homogenous reactions than for solution-solid reactions. The advantages of a lower enthalpy of activation on an LDH are overshadowed by the entropic changes of the solid 119 system. .There is another element in the activation energy. The gallery nucleophile is located in a potential energy well from which it must move»to react. Reactions that used an excess of RBr showed that approximately 90% of the inter- calated iodide was available for reaction. That the reaction rate depended only on the fraction of iodide remaining and not its location in the solid (iéh, the distance from the end of the gallery) suggested that the barrier to gallery diffusion was surpassed at 90 oC. Aqueous anion exchange reactions that occur without particle delamination support the supposition that the barrier to movement is relatively low. This discussion is also related to the next point to be considered. (d). When l-bromobutane was adsorbed onto O.D. [ZnZCr(OH)6][I] at 22 °C, normal adsorption behavior was observed. When the same experiment was tried with A.D. material, a reaction occurred in which the weight of the material decreased. Analysis of,the adsorbate revealed that l-iodobutane had been formed in the vacuum line and had condensed in the reservoir of l—bromobutane that was open to the vacuum line. That reaction occurred on the A.D. solid and not on the O.D. solid is at first surprising since the O.D. material was more reactive in solution studies. The treatment of the samples must be kept in mind, however. Before adsorption, the so-called AJL material was equilibrated at a pressure of le0"3 torr. 120 This treatment caused a weight loss of approximately 4.4%, enough to remove the surface water and a small amount of gallery water. In the adsorption studies, the difference between the samples examined was mainly a difference in gallery water and not surface water. The solids used in reaction studies were different in surface and gallery composition. The differences in behavior of materials used in ad- sorption studies was related to the gallery properties. The d00l spacing of an.AJL.iodide-exchanged Zn-Cr LDH is 8.34 A vs. 8.21 A for an O.D. sample. Assuming an octa- hedral°layer thickness of 4.8 A, the gallery heights of the O.D. and A.D. samples are 3.4 and 3.5 A, respectively. The diameter of the iodide ion is 4.32 A, larger than the gal- lery height. The octahedral layers somehow distort to fit more tightly around the gallery anions. The electron cloud of the anions could also be distorted to decrease the dis- tance between charge centers in their respective layers. If the anions are located over a metal center in the octa- hedral layer, the three hydroxyl groups could conceivably distort to allow closer approach of anion. The gallery water of AND. samples apparently props the layers apart to some extent since the gallery height of an A.D. sample is greater than that of an O.D. sample. The propping action of the water molecules is not the same as a column holding up a roof of a building. The water functions more as a dielectric. The oxygen of water compensates for some of 121 the positive charge of the octahedral layers, decreasing the amount of distortion that the iodide undergoes. Because of the distortion of the anions, they are in physical "wells" in addition to the potential energy wells caused by electrostatic attraction. In order for reaction between a gallery anion and an alkyl halide to occur, the anion must come out of its physical well, i.e., either the anion must further distort or the gallery height must expand. Either process requires energy. 131the reaction that occurred in the vacuum line, the A.D. sample had a gallery height that was 0.13 A larger than the O.D. sample. The difference, which corresponds to a difference of 3.1 kcal/mol, *was enough to inhibit reaction on the (LD. material while it was permitted on the~AnD. material. A question that remains concerns the nature of the activation step in the species that exhibit activatied chemisorption. The activation barrier for chemisorption involves some rearrangement of the adsorbent to enable the interaction to occur. Examination of the AHlad curve for s toluene indicates that activated chemisorption occurred in the early stages of adsorption on A.D. LDH (Figure 32). The extent of chemisorption was much greater for CLD. LDH. Since the adsorption interaction presumably takes place through hydroxyl groups, the activation step at least par- tially involves preparing the hydroxyl groups for binding by adsorbate. The preparation may involve breaking hydro- gen bonds that exist between particles.32 Since hydrogen 122 bonds are weak interactions, a larger number of them could conceivably be broken at 22 0C than at 0 oC. The greater extent of activated chemisorption on O.D. samples indicates that, compared to the AJL material, a greater number of the hydroxyl groups are unavailable at 0 °C than at 22 °C. There could be more hydrogen bonding between particles in OJL.LDHs, or adsorption into the galleries could require an activation in O.D. solids not required for A.D. material. This difference in activation required for chemisorption, when examined in terms of gallery heights, is additional evidence for adsorption into galleries by adsorbates. The absence of chemisorptive behavior of carbon tetra— chloride also suggests that adsorption into the galleries takes place. Carbon tetrachloride is unique among the adsorbates studied because it is the only one too large to fit into the gallery an iodide-exchanged LDH. The strength of interaction between carbon tetrachloride and hydroxides is predicted to be the weakest of the adsorbates examined. Since the interaction of carbon tetrachloride is only slightly weaker than that of octane, according to reports, some chemisorption behavior might have been predicted for carbon tetrachloride. These data can be interpreted as evidence for chemisorption into the galleries since carbon tetrachloride alone doesn't chemisorb onto CLD. LDH. In the discussion of differences in reaction rates and activation parameters in different solvents, theigreater 123 rate in octane has been attributed to greater access of reactant alkyl bromide to the galleries. However, if a greater number of collisions alone were responsible for the increased reaction rate, the value of AS' for reactions in octane should have been less negative than that for reactions in toluene and chlorobenzene. The value of AS‘ is derived from the A parameter of the Arrhenius expression and is a reflection of the frequency'of reactions leading to reaction. Instead, the AS’ values for the 3 solvents are the same (Table 19). The reactions in octane have a smaller value of AH’ than those in toluene. 'The value of AH’ in carbon tetrachloride is even smaller than for toluene. These values point to the importance of a step discussed in relation to the first step of the proposed reaction mechanism. Recall that a solvent molecule must be desorbed in order for a reactant molecule to adsorb. Since the adsorption was exoergonic, the solid loses energy when the solvent molecule is desorbed. The amount of energy the solid loses is proportionate to the strength of interaction between the solid and the adsorbate. Therefore, the solid loses more energy in desorbing toluene and chlorobenzene than in desorbing octane. Desorbing carbon tetrachloride requires the least energy since it interacts more weakly than the other solvents.‘ The energy that is lost by the solid is energy that could have gone into the activation of the iodide substitution reaction. Therefore, since less additional energy is required to activate the reaction, the 124 solvent system which causes the smallest loss of energy upon desorption should have the smallest value of AH’. The above explanation accounts for the enthalpies of activation observed in various solvents. It