‘5'I47II‘II ”,5. III ,..II' 1 wfiw 77'7" 7".77 . 7 I II II III I'" “if: , '1: )7 "I‘ 777 ”"177. ":7'",-'7 “I: '777i‘77757t77777777‘77771777777777 "H7,“ 7'I77'I7I 7:77:77RP' .nII “fl EMA“ ’3 I§I.3I:I.:'II "772?" 'II777 7I7 7777M II‘J'I'“ 7": 77 ' 7I77I 7777771777537: 7"”I’I7III'II31' 7&7?77"'5:J '7'! 7‘71: ‘77. 37:19:. ' :7 I717 7‘-,"I‘I-7I.’ I, I I7'7j 777 7' 77 77 77 77:;77739‘ 77 ',;II"'!h{ ' 7 777: : l 2&17'7I'IIVLI17 I. }77"77' 777 'I77 77III' 7I77 'II' .I'II'I i'§::"7'7:7 77777777 {7:177:73‘777377 I I I 'l' I, 1",, I I . . .2; . I' II‘IzIEI‘IP ',1,|'7u737'7;':"7l:"I'I7‘:|I|:7':I77II7'II1"IH ‘II I "7; 77777:; li""'I‘$ 77|7"‘ ("7 I!" ’ 'I:'777'7\ ' i " I: I. ‘I'ImIL, ,. ,II.I|I:..‘I 77:717.,h77’ "HEW!" 77;E7L7EI7,;:IIE&iEI:.;77H 'Isz' i' in,,L,I .5!’ III, 1.7'.'." 'I'77"'"I'IT"7"'... ..I’I-'I:.HII’-'II'I r" “‘II‘II '7 II 7 ,1 A V x ' ~_ . 'z' gt.’ a . 2.- ..‘..l- -- '25.. _/ - -_-'--.-' ... .4: _.;. “I“: - ”..— _..‘ b.- 7777;77:717 “'0‘!" V", I: , 'I 7'I' ' ":7’ {'7'7 “"’ I"I"I '-' I:5 "III" II 77:37." ":1 ‘7 II'W' I‘I'L'77EEL'EW 3177777 7| 71!;775or 7"777777777 .hl,: '_"I ' I0 ' .. 77'II 'HIU7'7 7717:." '7577'1'7I “{77'I 7:" I7‘"' -.-.I; IIII'I‘I.III:I"ItTI-4II7 7'” . - 7'31: IIIU' 'tIIL-‘IIIIII'I'I‘IIi'II .H' 71"I7III7I‘ ,jI.. "LEI I'I7'II 7 7E. I ; v ‘ , . . ' hl , 1 .' 'Wm'""'77" '“'7I7"7'I77'II'-'I7‘7‘I’IIII.:II In: :7 IIII"7"77 Mi; "'I|II IIIII II‘II'II17 :79. fig}- :sz "77$ 1 II~LIII III :Jffi'i" 1,,7I7IISI II,.,I.IIIIII I;;;fI.I 71‘ 'If‘. “1,5? "7"II!""77‘7 13,}, ”I: ,4! .I I ‘ 'LIII’. :7 I ,7‘ 7“] 5 ‘ ‘7 l7 J E‘ '. .‘ I H . ' ' '“uJ :‘E’I: 321:7 '7 I77 |:;| {77‘1' ' 77‘7“ ' .7'7177 7'1" .| Ir‘:777'"1'j7~,' 7:777" "77 77:7;7' I77 9 I '7' 71'1" ‘ 1’37 F‘fi‘flflb 377;)“: 77. h, 7777777777'1'15777hl7’7 .,7,',, .’ , '7..lll::'l|;‘, :II.I‘.I ;-,:II:.II,:,I:;I IN , ._|1,,.~;|;II:;I .I, I:I+.,.I,;f4;v,|,I,;fl,. I. :~:-.I‘.,. ‘,I,$7’.:‘-,.L' l,v " 7 "3,, IL=’II‘~‘=I‘I:-I'EII "' I'I'I’ II'"? ‘ I'II .i‘I'I" (”II'I‘II ‘IIII III'I ‘I' 'III. IIII‘II I§;’I.5.I4I."¥II. .' I. 77 77“; :0 I7 . I777: “I"III. . ,5‘I . ‘ ", ::~.-,,?x'i'§ "-YIIII" [ {'1' .7:77 577735.; :1": 7,757“- '777 77'7“”; 'I7'~ '77777'7' 7777 777777777 7 7 ' ' '7' 7I'7I 17: '7' ~I'77"'7"'.II7 ’7' 7':""‘7' .'-II ENS; 75v, 7,77 777673747337#5775777?'71‘77 M '77 ‘ ‘-‘; ‘ ' I. ‘ ' II I . I III 1" u" o I '- I' I E IN 7E, 2.]. 775,. 77:7 7‘7 ,7 .I777u7l77777 (I7 7:7; 777747,“ 77 7‘: 777.7 "7) t . _.‘ I' '33: s 77.71; '7‘, . 77:7"( “-7 77:377f7:..77"‘3:-1 7 «adufu .-;'_'7I7H ',I.,}:. ”‘1 _ I ; 777E77777 ‘1‘. 77: H .II .. | I g I" ,0“ I In Q7\".t..:7;‘vli4.7‘l:b’ 72577'7'7I f." v ', .I_,. ,. IE7: I . I 33411;. ;, 77.,::2; I” . . ',-':'.7. 'I,| {77 l I 11,”. my": I II M, .. ;.,I ,, 1‘” |..r".,;§;v'l,:-;I .51! ., :hu ‘. I ,L, ‘ ‘7, 01"" t I :.I:;, l : | “1;. '1. fl "7:777 ‘ 9;; 9;, :.I; .| 7.: . I .‘ '.'I' ‘l7 " '7 | ..‘I . I' 37:. I 7 .7031 '-I '07 II 'l;. . H1: ..' :' ;'II.II,.II|’I it}. ;I, I u |. l ,Zz'r;...;!;.;. ;;,- 0-,:1.’ 1 1" ’1: ..’ ,I,, I I, I 3.") : e 2: ..': ,'|.; ,: .;_ I I ; ';I;,II,*,I, -.-";' ,I,-, g“,- 5, ,:II;;II ”:III ,,;.I,. .;- Vial-75' ,_'..I, «2-: MN.” ,,.I:; I ,,. "1'7: :I-“I' I. ..«JI‘V' II. "I ‘r'f'IIIII:"' .. . .‘t, ‘1: L7) \ .I 79.79" h . 77:24:; 2:377:7777:3'7:77 “3-350”. ,7!“ 1;»: I ”2:, . t't.|3 "h, ‘7'”.7: . , I .'~; “"‘II .I"‘.' .7. .. i-J‘ I'.' I15} ;_. ' ..‘ 7'7 v. I Lit-7V '7’t777‘ 'a;?,;;' ;-." . ',;|M I . MAIN“ ;a _I . ,. I. ,I l I O, l . w. I I I II c I, 5 'I .. ~ a II .1. .I - 7 . I I7_-7I‘ - - - I; I I I , I. I I- ':-7I-: I727 7 7 7 7 0 \ § 77 I 7' 7 - I! ’ - -I - :‘:I7"'-I~- I I" "7 '. ‘ - I . I ' ‘ -Ii" .II' - 7 7 77 777' 7I7777 777177.777“: 777:3- ’77 7 7 7 s7 a . 7 7 7.777 I 777 7 I. 7 - ., . I . '72,. . ‘.,7; ‘7 _ I ' | 7 7 ' . '2‘ ,. I . 'f'," I 7: - ' ‘ ‘I | ‘ I .- I - "" u ' ' I' ,- I7 7 ' I "31 ' I' ':" I 7 7 '7 L I» ;-2".I . I t'?- ' " " ' ' 7 7’- ,7' .7,:,'.. .7“I: 7;. I 7 , I I. I. . _, I' 77 . {‘34: 77 ’ ..‘ ’ 1.7.77 ..' ; \ 7" C . - . - , I o ..‘ . {II‘ 7,, I :I 7'7 ‘ ,, ; ‘5' '77« 7‘7" 77757 7 *I ' 7 '7 L 7 ’ 7 I 77.; 7 “'7 I 711- .‘I 7‘I 7.1.7 I II ' ' ' 7~ ,7 I '- .- 7 ." . “.. 0 . 'I 'I . I '- I’ I. :- --' - f; . II II ' . II . :a I I17. 777 ' 7 7 :07 7 7' 77 7 7 : ' '- ' I 77' 77" l' ,. ',"I7 .I;1I , 'I_ ,. .g 1 . ' - "If" 7. "' .‘ 7,“I 7. I7 7,.I-c , 7 . .‘ ' 7.7'77l .7 I 7‘8; I: n, ' I" o '73. ‘73:: Io;‘|. , I , :u I ' c , ., :il.7:7 r ..‘. . ‘I,:.:f| : 'I . II 0' I - 3 ‘ J" ' . I: ', g - ;, ,I,7 I I 717.7 . I I: II I II'» ,I' . I' | E ,|-NII t ' 1,. - . .,;v ".5, ..’ l I ' I I 7. , I ' ‘I I L I-I'I I: I. . . I I III ,,, , I_I,I‘ I'I:,I,, 7 1‘7 7 :7I ‘7Il' ‘.u'. I . ' " I ' 77 ; 777.'Z,I"71.7':;l77"73.' '1' , l7 I'II,;7 II| 7I:|7. /' |' , l I. I." 7 .‘ I 117',77l7..' ~77" .,;\ 7 E 5:073;,|Ii|,_,: 7,||,;.,I I ‘I: : , 7 ’ ' 7 ,‘7,| I: ,;' 7 I. :I\I.:'_'7 I77 fit, 7177‘77'i77'l7777:7w'f7712L‘77;'07:: . II' II I‘ - 'I I I'I'IIrI I‘I II‘-I'- II ._ III .I. ‘ II. . 77"; ‘I'II " -II- . 'I h 97 II,T;',I: 7.! ‘ ;I,, | I, II'7| “I. I‘l u, .I' I. I' 'I, ;.,";,'., II | . . .7'777;--.- I |§:I|I;|I|I|.' 'l,’,' I. III' IIII' 1,1It, ' I..I I,'.I' ,'I' '|~I;II‘I I}: ,.; 7 """d, I'M "’ 4477’ III I‘ “'1‘,-, g I a". ‘ I I. I ,' ".7 ,' ")1 ., I . ' ,. I,7.,I ”777177;. In."77 , ‘ I ”n .‘I I, I_I'I ‘I I,:|I{7'11I:l7:£t:7|h7'1‘ |I I I.IIII ,.:,.III I I7|§7n III II.I‘.II IIIII ”I ,,.I ..; ':'E: 'I I.IIIIII I : :I;|:7t7" '|7,.‘”' I‘7 .477I'7 ..|I ': 'H'H !"::,",'|||1 I'- h 77':|'7 |I||\II| I,”,IIIt|I7 MIN, |,‘P'I I:I'IIII| 77"" 7711!,9,II1,|II:1;I- 77I"7 7"I"I:77 'I7'I' |II’I-I 'I.|‘ ' _: II" l. :'I H: 7'!“ I,I " 7:7"; '5‘”: :IIII' ""‘I ”if. III: II'” ‘17":7':'HII'I".I‘;."":" I"I‘I 'FI |"I‘|" I,II||| ,HIIHIII I""I|7I"" if 777L '3'9: ,, 7II , . I III; III||:; 1"!“ III: | . 7|: {III N W, .2'55‘I ' I " ’ ':",‘II III] E "7 ' ""I" 43:1,?" EtizIfI '27s f'HE I: ff: ,7. § .II, ,7.” I,.,, .:.IIIIII:,I: ”5| :9in ,fl, ,‘7, "III 77'7"" 7., I ' ,7.'7I.II|I,,,I.7.II. :: III'E‘I | I I II I I I I '7'77777 ":I77'. II'II Eh! I ‘1 E; *3: I:I||LII’I 1“: I771 2.4% “1‘7 P71 I 773777! I If. Y ["71 I...II.I. II} V; - . i" '7IL,‘,1 I77g'III;I1I:. I ':7IuI.II ‘ M ‘ I A»: :':f777 7""“I ,, ”in“, .I .,, .,,II,. ,II,I;77 I‘IIII'..II IIII: IIIIIJ,II,,L,,, ‘III'.2'(.I§?:,I~III III fun“. :II III, F, $.11..— u .3: ca. 3253.”?— ” "..‘ QC... Iaxrm'g‘ . w - HI "' ., .43....— .-E m... ‘_ A; a; ..‘: “run-A '. .....7.."‘"‘ .- ?‘A—-oq ’11,,“ .121 :33“? ”hf—‘wr—VFVT.‘ «gm u: :s .. $19.25.??? *Mam‘w,uprwnmmo :' 77": 3:71:77 I . m AXE. «, .u—am ‘ . -.. . . .. “-30:1- ' W -‘ 7:4." SEE 1‘3 I I I 5 1:43. 23?“! IIFIIIE: 3' 7 37.77:: 74:514.? "I' .- I' I ’IiIIfiI" 1 ‘I :15. I ‘I I... llllllHllllllllIllUH“lllllllllllllllllllllllllllllllllllll 31293 00904 9416 This is to certify that the thesis entitled Preparation and Characterization of Vanadium Oxide/Aluminum Antimonate and Supported Antimony Oxide Catalysts presented by Thomas John Curtis has been accepted towards fulfillment of the requirements for M. S . degree in Chemis try ajor professor 0-7639 MS U is an Affirmative Action/Equal Opponunity Institution » LIBRARY Michigan State University J PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or baton date due. DATE DUE DATE DUE DATE DUE H __J :1; ~ ‘1 Tl—T J MSU Is An Affirmative Action/Equal Opportunity Institution cmmma-m PREPARATION AND CHARACTERIZATION OF VANADIUM OXIDE/ALUMINUM ANTIMONATE AND SUPPORTED ANTIMONY OXIDE CATALYSTS By Thomas John Curtis A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Chemistry 1993 ABSTRACT PREPARATION AND CHARACTERIZATION OF VANADIUM OXIDE/ALUMINUM ANTIMONATE AND SUPPORTED ANTIMONY OXIDE CATALYSTS By Thomas John Curtis A variety of bulk and surface characterization techniques have been used to examine the structure and reactivity of novel V/AleO4, Sb/Al203, and Sb/Si02 catalysts. This work will establish the active phase(s) of these catalysts. V/AleO4 has been charaterized using XPS, XRD, IR, solid-state NMR, and EPR. XPS V 2pm binding energies for the catalysts indicate that V is in the +5 oxidation state. For V205 loading 5 20 wt.%, XPS data suggest that the V is highly dispersed over the Ale04 carrier. For loading 2 25 wt.%, XPS indicates that V is poorly dispersed. These results are consistent with XRD and IR data obtained. Propylene oxidation data shows that vanadium promoted sample shows a greater selectivity to acrolein (90%) for 5 wt.% vanadium than to that of the pure support (59%). The increase in selectivity of propylene to acrolein is due to V=O bond. Sb/A1203 and Sb/Si02 catalysts have been characterized using XPS and XRD. XPS Sb 3d3,2 binding energies indicate that the Sb is in the +5 oxidation state. XPS data indicates that Sb is highly dispersed over the alumina and silica supports. CO oxidation activity is highest for catalysts with active phase loadings of 10-20 wt.% Sb205 on both A1203 and SiOz. CO oxidation activity over Ale04 is greater than that measured for the Sb/Al203 catalysts. To my parents, Thomas and June, my sister, Denise, and my nephews, Nicholas and Alexander ACKNOWLEDGEMENTS I would like to thank Dr. Jeffrey Ledford for his guidance and support during the course of this work and for all the knowledge I obtained in our conversations. Without his support this project would never have been accomplished. I would like to thank my parents, my sister and my nephews for their support and understanding during the past few years. I would also like to thank the Ledford group members, Jeff R., Paul, Kathy, Greg, Mike, Mark and especially Ed and Per with their help with ChemDraw and Designer. 1 would like to thank the rest of the graduate students, especially Tim, Scott and Doug for their friendship during this time. Finally, I would like to thank all people who have provided assistance on this project. iv TABLE OF CONTENTS Page List of Tables - -- _ - -- -- - - .... _ _ vii List of Figures viii Chapter 1: Introduction to Selective Oxidation Catalysis..............-..- . -- 1 1.1. Overview of Selective Oxidation - - ........... - - - 1 1.2. Overview of Methane Conversion _ - - - - - - - -_ 8 1.3. Overview of CO Oxidation- ...... - - - 14 1.4. Focus of Research _ - - - -- _ - ..... 16 1.4.1. Selective Oxidation.--- - -- - 16 1.4.2. Methane and CO Oxidation _________ - 17 References -_ -- - ..... -- ...... 19 Chapter 2: Preparation and Characterization of Mixed Metal Oxide Catalysts ............... 22 2.1. Preparation of Catalysts - ......... - .... _ 22 2. 2. Characterization of Heterogenous Catalysts ...... - -- . 26 2. 2 1. Bulk Characterization -- - ........... _ . . ..... 26 2. 2. 2. X-ray Photoelectron Spectroscopy (XPS) - 28 2. 2. 3. Catalytic Activity--- - -_ - ..................... 34 References - -- -- - -- 38 Chapter 3: Vanadium Oxide/Aluminum Antimonate------- - 39 3.1. Introduction - -- ...... 39 3.2. Experimental - _ - - - - - - 40 3.3. Results - - _- - -- - - _ 45 3.4. Discussion - _- 59 3.5. Conclusions - - .......... - --.....64 References _ - -- - - ..... 65 Chapter 4: Antimony Oxide Supported Catalysts -- - -- _- - --_-67 4.1. Introduction - ........... 67 4.2. Experimental -- -_ ....... - - 69 4.3. Results 73 '4.4. Discussion _ - 83 4.5. Conclusions ..... -- - ............. . --86 References - _ ..... _ ...... 86 Chapter 5: Future Work- -- -- - - ................ - 89 5.1. Characterization __ _ - - _ -- 89 5.2. Vanadium Oxide/Aluminum Antimonate Catalysts- - - ..... _ - 90 5.3. Antimony Oxide/Supported Catalysts ......... - 93 5.4. Activation Energy ................. 94 References 95 Table Table 3.1. Table 3.2. Table 3.3 Table 4.1. Table 4.2. List of Tables Page BET surface area (mZ/g) of V205/AleO4 catalysts. ........... 46 Particle size V205/AISbO4 calculated from XRD data. _ -- -- - . ................ -..--51 Conversion and selectivity of V205/A1Sb04 for propylene oxidation at 200 °C and 100 cc/minute flow rate ...... - 58 BET surface area (mZ/g) of SbZOS/A1203 and SiOz.----- - - - . . --.-.-74 CO conversion for Sb205/A1203 and SbZOS/SiOZ at 450°C and 15 cc/minute flow rate. - - .......... 82 vii List of Figures Figure . Page Figure 1.1. Reaction mechanism for selective oxidation and ammoxidation of propylene.--- - ................ ........ - -3 Figure 1.2. Mechanism of selective ammoxidation (lefi) and oxidation (right) over bismuth molybdate. (FromJ. Catal, 87, 373, 1984).-- - ....... - - ........ -5 Figure 1.3. Reaction mechanism for selective oxidation and ammoxidation of propane. (From Ind Eng. Chem. Res., 31, 107, 1992). - ........... - . . 7 Figure 1.4. Reaction mechanism for methane conversion. (FromJ. Catal., 112, 168, 1988) ................................................................ 10 Figure 1.5. Reaction mechanism for non-catalytic coupling ofmethane. (FromJ. Carat, 121, 122, 1990). - _- - - - --12 Figure 1.6. Random mtile structure. (From Acta Chem. Scanda. A, 29(9), 804, 1975). - - . ................. 18 Figure 2.1. Various types of active oxide-support interaction. (From Preparation of Catalysts II, 439, 1979). — - 23 Figure 2.2. Reaction of metal alkoxide with support. (From .1. Carat, 101, 1, 1986). - - - - -- - - - --25 Figure 2.3 XPS process..- - -------- - - - - ---29 Figure 2.4. Model of the catalyst particle. (From .1. Phys. Chem, 83, 1612, 1979). - . - ------32 Figure 2.5. Protocol for study of heterogeneous catalysts by reactive gas treatment combined with XPS. (From Anal. Chem, 58(12), 1184 A, 1986).. ----- - - -- 35 Figure 2.6. Reaction chamber. - -------- - ........ 36 Figure 2.7. Catalytic reactor ------- . - - ..... - - -37 viii Figure 3.1. Figure 3.2. Figure 3.3. Figure 3.4. Figure 3.5 Figure 3.6. Figure 3.7. Figure 3.8. Figure 4.1. Figure 4.2. Figure 4.3. Figure 4.4. Figure 4.5. XRD of V205/AleO4 a) V25, b) vzo, and c) vo catalysts. ......... 47 27Al MAS NMR of a) V25 and b) V0 catalysts. 48 Semi-quantitative XRD for physical mixtures of V205 and AISDO4. - ------ 50 XPS spectra of V25 catalyst a) 1n propylene/Olee, b) 02/I-Ie, c) spray coated, and d) as powder--- - - 53 XPS intensity ratio of V 2p3,2/Al 2p as a function of V/Al atomic ratio ....................................................................................... 54 IR data of a) V205, b) V25, c) V20, d) V15, e) V10, and 1) V5 catalysts ............................................................ 55 Vanadium Wide-line NMR of a) V25 and b) V5 catalysts- . - 56 EPR V“ spectra measured for reacted a) V25 and b) V5 catalysts. Spectra were collected at room temperature. -- - ...-.57 XRD Of SD205/N203 a) SD40“, D) SD35AI, C) SD30AI, d) Sb25Al, e) Sb20Al, t) SblSAl, g) SblOAl, h) SbSAI, and i) SbOAl catalysts.----- 76 Semi-quantitative XRD for physical mixtures of Sb205 and A1203. 77 XRD of Sb205/Si02 a) Sb408i, b) Sb3SSi, c) Sb3OSi, d) SbZSSi, e) SbZOSi, f) SblSSi, g) SblOSi, h) SbSSi, and i) SbOSi catalysts. --------- 78 Semi-quantitative XRD for physical mixtures of Sb205 and Si02.- ------------------- 79 XPS intensity ratio of Sb 3d3,2/Al 2p as a fimction of Sb/AI atomic ratio. - 80 ix Figure 4.6. Figure 5.1. Figure 5.2. XPS intensity ratio of Sb 3d3,2/Si 2p as a function of Sb/Si atomic ratio. 81 In situ IR cell. - - - - - - - 91 Gas handling system to in situ IR cell. -- - -- - . .............. 92 Chapter 1 Introduction to. Selective Oxidation Catalysis 1.1. Overview of Selective Oxidation Reactions Selective oxidation of hydrocarbons is an important industrial process that uses transition metal oxide catalysts to convert less valuable hydrocarbons into more valuable compounds (e.g., acrolein and acrylonitrile). The most common catalysts used for these processes are based on metal oxides of vanadium and molybdenum. A common feature of these reactions is that the desired products are often not the most thermodynamically favorable ones. For reactions carried out in either air or oxygen, the most thermodynamically favored products are water and carbon dioxide. The desired products are alcohols, aldehydes, ketones, acids, anhydrides, or alkenes and dienes [1]. Thus, for each hydrocarbon, numerous products of various extents of oxidation are formed before carbon oxides are produced. Hence, it is a challenge to understand the factors that account for the selectivity and activity of a catalyst and to develop a practical catalyst. In selective oxidation reactions, the selectivity is usually dictated by the ability of the metal oxide to catalyze the formation of C-0 bonds while minimizing the breaking of C-C bonds. Selective oxidation reactions can be classified into two types: dehydrogenation and dehydrogenation followed by oxygen or ammonia insertion. In dehydrogenation reactions, aliphatic molecules are converted into alkenes by breaking C-H bonds and forming C=C bonds. Often oxygen is used as an oxidant to yield water as a byproduct. The oxygen additionally provides the thermodynamic driving force for this process. Other oxidants (e.g., 1, Br, and N20) can be used, but oxygen allows the reaction to be conducted at lower temperatures [1]. 2 In dehydrogenation and oxygen insertion processes, C-H bonds are broken and C-0 bonds are formed. Oxygen is used as an oxidant for the formation of oxygenates and in the formation of water in the dehydrogenation step. Exceptions to this general rule are the oxidation of ethylene to ethylene oxide in which no C-H bonds are broken and ammoxidation reactions (such as propylene to acrylonitrile) in which C-N bonds are formed. Oxidation of benzene to maleic anhydride is an example of a reaction that involves the breaking of C-C bonds and the insertion of oxygen. Catalysts used for selective oxidation and ammoxidation reactions contain several types of sites: olefin chemisorption sites, (Jr-hydrogen abstraction functions, oxygen or NH insertion sites, and redox couples associated with dioxygen dissociation. The olefin chemisorption site and the oxygen or NH insertion sites are often associated with elements in their highest oxidation state (e.g., Mo+6 or Sb+5). The (Jr-hydrogen abstraction site consists of metalloids with free electron pairs (e.g., Bi+3 or Sb”) that give radical character to the oxygen bond at the site. The presence of radical oxygen facilitates the abstraction of a-hydrogen as a radical from the chemisorbed olefin (an important step in the formation of the n-allyl intermediate). Oxygen or NH insertion fi'om maximum valency metal sites into the n-allyl species and forms the o-allyl surface intermediate. During this process, reduction of the metal (e.g., Mo+6 to Mo+4 in molybdate catalysts) occurs followed by desorption of the product. A redox couple with a reduction potential greater than that of the oxygen or NH insertion site may also be present to facilitate reoxidation of these elements back to their original active state. The redox element also serves as a site for 02 reduction [2]. Figure 1.1 shows the general reaction mechanism for the selective oxidation of propylene. CH CH=CH > ACTIVE\ Km 02 SITE 13!le nM[] REOXIDATION H ' 0 SITE :MQ/MZ/m f “[1 NH 1 3 CHZ=CHCHO + H20 C112=CHCN 4.31120 Figure 1.1. Reaction mechanism for selective oxidation and ammoxidation of propylene. 4 Among the most commercially important examples of selective oxidation are the oxidation of propylene to acrolein: CH3CH=CH2 +02-—>CH2 =CHCHO+H20 (l) and ammoxidation of propylene to acrylonitrile: CH3CH=CH2 +NH3+%_02—-)CH2 =CHCN+3H20 (2) which are used for fiirther reactions with hydrocarbons to produce valuable compounds. In 1948, Adams used cuprous oxide as the catalyst to transform propylene into acrolein, with a yield of approximately 50% [3]. In 1959, Idol improved the yield of propylene oxidation to acrolein by using a bismuth-molybdate catalyst [4,5]. The bismuth- molybdate catalyst was also found to convert a reactant mixture of propylene, air, and ammonia to acrylonitrile [6]. Commercial vapor-phase oxidation and ammoxidation was developed by Standard Oil of Ohio (801110). In 1965, Adams [7] found that the bismuth— molybdate catalyst could produce butadiene from butene. Also, in 1965, a more selective uranium-antimony catalyst was introduced for the same set of reactions. In order to eliminate the radiation concerns associated with the use of uranium, an iron-antimony catalyst was developed. This catalyst is still used in Japan [8-11]. In 1970, SOHIO introduced the first multicomponent catalysts [12,13]. These catalysts were composed of a variety of elements including Bi and Mo. Based on research by various workers, Burrington et a1. [14] proposed detailed mechanisms for both oxidation and ammoxidation over bismuth molybdate. The mechanisms for oxidation and ammoxidation are shown in Figure 1.2. Figure 1.2. Mechanism of selective ammoxidation (left) and oxidation (right) over bismuth molybdate. (FromJ. Catal, 87, 373, 1984). 6 In these mechanisms, the surface active site is a Bi-Mo pair site that is composed of a Bi-O group responsible for allyic H abstraction and a molybdenum dioxo group for nitrogen insertion. It is believed that propylene absorption occurs at the molybdenum cation. The oxidation reaction proceeds by dissociative adsorption of propylene to produce a rt-allyl species. It is not clear, at present, whether dissociation of propylene occurs upon adsorption or propylene is first adsorbed molecularly as a 1: complex and allylic H abstraction follows. The next step is the formation of C-0 bonds and a second hydrogen abstraction in the form of a (1,4) shift to produce adsorbed acrolein and Mo- OH. Finally, the product desorbs and the catalyst is reoxidized. In ammoxidation, a surface molybdenum diimido species is formed by the reaction of dioxo groups with ammonia. The activation of the propylene occurs in the same manner as the oxidation cycle. The reaction proceeds by dissociative adsorption of propylene, followed by formation of a C-N bond and two additional hydrogen abstractions producing acrylonitrile. The reduced surface site is reoxidized and then reconverted to the diimido species. Because of the reduced cost and the greater availability of alkane feedstock, significant effort has been devoted to understanding the reaction mechanism for the oxidation and ammoxidation of propane. At present, Cantani et a1. [15] suggest that the reaction mechanism occurs as shown in Figure 1.3. In this scheme, the propane is converted to the main product of propylene, which then further reacts to form acrylonitrile (ACN), acetonitrile (AcCN), hydrogen cyanide (HCN), ethane and ethylene (C2), and carbon oxides (COX). They suggest that the formation of acrylonitrile passes through the intermediate formation of propylene and the production of acrylonitrile is limited by the slow rate of propylene formation. ACN v Propane > Prop lene ’ cox NC / 9 HCN > C2 Figure 1.3. Reaction mechanism for selective oxidation and ammoxidation of propane. (From Ind. Eng. Chem. Res., 31, 107, 1992). 1.2. Overview of Methane Conversion Recently, there has been . increased interest in the oxidative coupling of hydrocarbons primarily using metal oxide, rare earth oxide, alkaline earth oxide and alkali metal-doped oxide catalysts. The oxidative coupling of methane to form ethylene and ethane has received considerable attention due to the abundance and low cost of methane. Interest in oxidative coupling has also increased due to the possibility of conversion to either gasoline, distillate or other products. The conversion over metal oxides is thought to occur as follows: xCH4+(x—1)MO-—>CXH2x+2 +(x—1)H20+(x-1)M (3) In 1969, Union Carbide investigated the feasibility of using methane coupling for ethylene manufacturing [16]. They performed extensive studies on different metal oxides for this process and found that most of the metal oxides of group IIIA, IVA, and VA exhibit both high activity and selectivity for methane coupling. In 1979, Mitchell and Waghome of Exxon [17] achieved 40-50% conversion of methane to mainly ethane and ethylene over a catalyst consisting of a group VIII noble metal, a group VIB metal oxide, and a group IIA metal. Between 1983 and 1986, Baems et al. [18] studied several lead oxide supported on ‘y-alumina, silica, and alumina/silica catalysts. Their work showed low acidity supports are preferred for high selectivity to Cz's. From 1984-1987, Arco [19] screened a large number of metal oxides supported on silica. They reported that manganese, indium, germanium, antimony, tin, bismuth, and lead oxides achieved approximately 15% conversion and give 10-50% selectivity to higher hydrocarbons. They also found that more acidic catalysts were less selective to C2 formation. Further, they reported that addition of alkali metal leads to an increase in Cz's selectivity [20,21]. 9 Both reducible and irreducible oxides have been used for methane coupling. Bhasin and Keller [22] found that oxides of Sn, Pb, Sb, Bi, T1, Cd, and Mn were the most active. They performed the reaction without the addition of 02 in the gas stream. They concluded that the common characteristic of the active metals was that they could cycle between at least two oxidation states. Thus, these reactions can be considered as oxide reduction. A number of irreducible metal oxides (e.g., MgO, Li/MgO, Na/CaO, Sr/La203, and Li/T 102) have been studied as methane coupling catalysts [23-26]. In these systems, oxygen in the gas stream is required to generate and maintain activity, but the oxidation state of the metal oxide does not change. These catalysts also have an advantage over the reducible catalyst because they are easier to control. The oxidative coupling reaction mechanism and the nature of catalytic coupling site over Li/MgO has been extensively studied by Lunsford et al. [27]. Oxidative coupling is proposed to proceed via the abstraction of hydrogen by lattice oxygen from the metal oxide followed by gas-phase coupling of methyl radicals. The methyl radicals can then combine to form ethane or react with oxygen to form products as shown in Figure 1.4. The pathway leading to complete oxidation is more complex. High selectivity to C02 is observed when the reaction is conducted in the presence of 02. This suggests a significant gas-phase contribution to COX production, possibly by methylperoxy radicals or adsorbed oxygen. In addition, a surface assisted route evident during the initial phases of the reaction in the absence of 02 can contribute to COx production [28]. In addition to methane coupling, ethane, and propane coupling have received considerable attention. Recently, Otsuka et al. [29] showed that at temperatures less than 650 K, Na202 partially oxides methane, ethane, and propane to ethane, butane, and hexane, respectively. Kinetic studies by Otsuka suggests that the activation of methane is initiated by diatomic oxygen such as 02", 02'2 or 02 which are adsorbed on the 10 02 (g) 1‘ N CH302 (g/a) MOx MOx - 1 \ Figure 1.4. Reaction mechanism for methane conversion. (From .1. Carat, 112, 168, 1988) 11 surface [30,31]. The proposed mechanism for these reactions is shown in Figure 1.5. In addition to the coupling of aliphatic hydrocarbons, there is interest in the coupling of aromatic hydrocarbons. such as methylbenzenes. The catalytic oxidation of methylbenzenes by metal oxides to the corresponding aldehyde and acid are well known industrial processes [32]. In a recent patent [33], anthraquinone was reported to be the major product of the oxidation of methylbenzene. King [34] recently found that under anaerobic conditions methyl-aryl coupling reactions occurred to produce methyldiphenylmethanes from the methyl-substituted benzenes. The corresponding aldehydes and acids were minor products. From this work, King concluded that methyl- aryl coupling is more favorable using more basic and ionic metal oxides formed from elements on the left side of the periodic table. A possible explanation for this finding is that a methylbenzene molecule may be polarized on the oxide and a proton from the methyl group abstracted to form OH‘ on the surface. If the multivalent metal oxidant (MN) is easily reducible, the benzyl anion may transfer two electrons to form a benzyl cation. The overall reaction involves reduction of the metal oxide and formation of water fi'om OH' and H". The reaction is believed to occur by the following mechanism: + Ar-CH3+Mn+ ——>Ar-CH;+M(““) (4) + Ar-CH;+M“+ ——+Ar-CH;+M("") +11+ (5) Ar-CH;+Ar—CH3—->Ar-CH2~Ar-CH3+H+ (6) For easily reducible metal oxides, the reaction occurs through the methylene group on the methylbenzenes. The first step of methylbenzene oxidation on the oxide is believed to involve the abstraction of (Jr-hydrogen to form water. The oxide at the surface may be reduced to its metallic form which has very weak interaction with the adsorbed organic 12 in3 H CH‘ + N3202 NaO 0N8 C2H6 + N8202 k2 N80 0N8 can, H CSHB 4’ N8202 A N80 0N8 CH3 k 2 N80 —4—> N8202 + C2H6 T211. 2 N80 Na202 + C4H10 T3117 II It 20 Na—7—u. Na20 + H20 Figure 1.5. Reaction mechanism for non-catalytic coupling of methane. (From J. Catal., 121, 122, 1990). 13 species: A proposed mechanism on PbO is as follows: ZAI-CH3+PbO——)2Ar-CH2'+PD+H20 (7) ZN-CHZ'fiN-CHz-CHz‘N (8) The formation of bibenzyl occurs through methyl-methyl coupling. In addition to oxidative coupling of methane to higher hydrocarbons, the deep oxidation of CH4 to C02 has potential for large scale application. Deep oxidation of methane was widely studied in the 1960's and 1970's when it was hoped that less costly alternatives to Pt/A1203 total combustion catalysts could be found [35]. The use of catalysts for total oxidation of hydrocarbons offers several advantages over more conventional destruction methods such as thermal incineration. For example, the formation of carbon and nitrogen oxides can be completely prevented by using appropriate catalyst formulations. The deep oxidation of methane is described by the following reaction: CH4 +202 -—)C02 +2H20 (9) In addition to Pt/Ale3, other catalysts such as Pd, Rh, and first row transition metal oxides have been used for this reaction [3 6]. Recently, Lacombe et al. [37] have studied the total oxidation of methane pathway in oxidative coupling of methane over lanthanum oxide catalysts. They found that total oxidation into C02 proceeds mainly through surface reactions by slow steps which combine with the initial slow step of methane activation. In addition to methane coupling and total oxidation, partial oxidation of methane to CO and hydrogen has received considerable attention. Among the most important 14 commercial methods of synthesis gas manufacturing is the steam reforming of natural gas and light hydrocarbons. For methane, the reaction is as follows [3 8]: CH4+H20-—-—)CO+3H2 (10) Typical catalyst used are CaO and/or K20 promoted Ni/or-alumina, MgO or MgA1204. Prettre et al. [39] were among the first to study synthesis gas formation by catalytic conversion of CH4/02 mixtures with 10 wt.% nickel/refractory oxide catalyst. They found that the compositions of the final CH4/CO/C02/Hz reaction mixture agreed with thermodynamic predictions. 1.3. Overview of CO Oxidation In recent years, there has been increased interest in the environment and the effect of technology on the environment. Air pollution is generated from emission sources such as electric power generation, refuse burning, industrial, and domestic fuel burning, industrial processes and transportation. CO is the most abundant air pollutant in the lower atmosphere. CO oxidation was first studied by Langmuir [40] in 1922 over group VIII metals. In general, the reaction occurs as follows: 2c0+o2 ——)co2 (11) Even though the overall reaction is very simple, the reaction kinetics are, in fact, rather complicated. For example, when performed on Pd, the rate dependencies of the reaction vary from negative first order to positive first order for each reactant [41,42]. In addition, 15 surfacedifl’usion of adsorbed CO [43,44] and metal surface restructuring under reaction conditions [45] also complicate the interpretation of kinetics data. CO oxidation has been extensively studied on single-crystal. Pd and supported Pd crystallites. Evidence suggests a surface reaction between an adsorbed CO molecule and an oxygen atom occurs. Most commercially employed catalysts for CO oxidation utilize noble metals. This is because they have high intrinsic activity for oxidation, are not greatly deactivated by sulfur in fuel at temperatures < 500°C, and have good thermal stability. However, at temperatures between SOC-900°C, the pure metal sinters rapidly. Furthermore, it has been found that the noble metal can disperse as oxides on supports at temperatures below the decomposition temperature of the oxides [51]. Although noble metal catalysts are more developed, base metal oxide catalysts are of importance because of the natural abundance and lower cost. Frontier research in this area focuses on the development of contaminant resistant and thermally stable base metal oxide catalysts for catalytic combustion [52]. In 1923, Jones and Taylor reported that copper is an active catalyst for the reaction between CO and 02 [47]. Ertl [48], Harbraken ct al. [49], and Arlow and Woodrufi‘ [50] have performed studies on Cu single-crystals and found that high oxygen coverage retards the reaction rate and that a surface reaction between an adsorbed CO molecule and an oxygen atom occurs. However, there is no agreement as to structural sensitivity of the reaction. Since oxygen adsorbs more strongly than CO on Cu, high oxygen coverage usually exists and reactions rates are lower. Thus, copper oxide requires higher temperatures to obtain similar activities per unit surface area in comparison with noble metal catalysts. Prokopowicz et al. [53] have studied CO oxidation over CuO/SiOz and observed first-order dependency on C0 (characteristic of base metals) and zero-order on 02 above 4% 02. Huang et al. [54] have reported a strong dependence of 16 pretreatment on the activity of CO oxidation on CuO/y-AIZO3 catalysts and concluded the reaction is structure sensitive. In general, high oxidation activity requires metal ions that can have more than one valence state and can participate in redox reactions. Among the most promising oxides are C0304, N10, Cr203, and CuO because they are refractory oxides and thus maintain their activity at high temperature. Mixtures of these oxides often exhibit greater stability and activity than the single oxide. One such mixture that has been shown to have activity near that of noble metal catalysts is the copper-chromium catalyst. In this system, it has been suggested that c0pper oxide is the active component and chromium oxide is a promoter. 1.4. Focus of Research 1.4.1. Selective Oxidation The discovery of selective oxidation and ammoxidation of alkenes has been the subject of much commercial interest. Recent research effort has focused on the development of selective oxidation and ammoxidation catalysts for the activation of less expensive feedstocks, such as propane and methane. The catalysts commonly used are prepared by mixing a hydrosol or gel of the carrier with a sluny of a solution containing the active phase(s) and the promoters such as W and/or Mo. This procedure can produce very complicated catalyst structures which makes comprehensive characterization difficult. Thus, the effect of surface structure on the properties of the catalyst and the active sites are not well known. In order to understand the influence of the catalyst structure on hydrocarbon activation, a simple catalytic system must be examined. 17 Several authors have reported that vanadium antimonate catalysts are promising systems for the ammoxidation or propylene [55] and propane [56] to acrylonitrile. Centi et al. [57] have investigated the. structure and activity for selective oxidation and ammoxidation of propane for VSbOx mixed oxides supported on SiOZ, A1203, and T102 and promoted with W or Mo. Their work suggests that VSbOx phase plus szOs is not the active phase for ammoxidation. They proposed that the random rutile phase of Ale04 (Figure 1.6) modified with vanadium is the active phase for the formation of propylene and vanadium or other redox elements having M=O bonds (W or M0) is needed to form acrolein or acrylonitrile. Although VSbOx/A1203 is not as complex as the commercial catalysts, further understanding of hydrocarbon activation would be obtained by studying the interaction of the pure phases believed to be responsible for selective oxidation and ammoxidation. The objective of this work is to study mixed oxide systems which model specific active phase-support phase interactions that have been proposed to play a role in selective oxidation. The work will focus on V oxide supported on aluminum antimonate. The nature of the V-AleO4 interaction as well as the influence of V dispersion and reactivity on the catalytic prOperties of the mixed oxide will be determined. 1.4.2. Methane and CO Oxidation The surface catalyzed coupling of methane has had limited success mainly because direct coupling of methane is a thermodynamically forbidden process at the reaction conditions of interest. However, if an oxidizing agent is used in the gas stream, either partial oxidation to methanol or oxidative coupling can take place. A number of researchers [58-62] have shown that it is possible to convert methane to higher hydrocarbons by reaction with metal oxides, thus coupling the methane through the 18 Figure 1.6. Random rutile structure. (From Acta Chem. Scanda. A, 29(9), 804, 1975). 19 abstraction of lattice oxygen and the formation of water. Antimony oxide has been used for methane coupling because the melting point of the compound lies in the low melting region of the periodic table and has been reported to give high C2 selectivity. In addition, Sb204 is a well known allylic oxidation catalyst promoter in propylene oxidation [63]. In addition to methane conversion, antimony oxide has also been used as a promoter for the oxidation of CO [64]. Ali-Zade et al. [65] found that the activity of Cu, Co, Cr, and Mn oxides supported on y—alumina is improved when Sb oxide is added. Antimony oxides have also been used in conjunction with oxides of Be, Mg, Zn, Al, Si and rare earth metals [66] and Sn-Mn-Pb oxides [67] for treatment of exhaust gases. The goal of this research is to investigate the surface structure and catalytic properties of antimony catalysts supported on A1203 and SiOz. This work will focus on catalyst preparation and characterization and will establish correlation between activity and structural information obtained from bulk and surface spectroscopies. CO oxidation will be used as a probe reaction to obtain information about the nature of active site(s) for the reaction. References 1. H.H. Kung, "Transition Metal Oxides: Surface Chemistry and Catalysis," Elsevier, Amsterdam, 1989. 2. JD. Burrtington, C.T. Hartisch, and R.K. Grasselli, J. Catal., 81, 489. 1983. 3. G.W. Heam and ML. Adams, US. Patent 2,451,485, 1948. 4. JD. Idol Jr., US. Patent 2,904,580, 1959. 5. F. Veacth, J.L. Callahan, EC. Milberger, and R.W. Foreman, "Proceedings, 2nd Int. Congr. Catal,” 1, 2647, Technip, Paris, 1960. 6. F. Veacth, J .L. Callahan, and EC Milberger, Chem. Eng. Process, 5b, 65, 1960. 7. CR. Adams, ”Proceedings, 3rd Int. Congr. Catal.," 1, 240, Amsterdam, 1965. 8. IL. Callahan and B. Sertisser, US. Patent 3,198,750, 1965. 9. G.K. Boreskov, S.A. Ven'yaminov, A. Dzis'ko, D.V. Tarasova, V.M. Dindoin, N.W. Sazanova, J.P. Olen'kova, and L.M. Kefeli, Kinet. Karat, 10, 1109, 1969. 10. ll. 12. 13. 14. 15. 16. 17. 18. 19. 20 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 20 T. Yoshino, S. Saito, J. Ishikara, T. Sasaki, and K. Sofugawa, Japanese Patent 7,102,802, 1971. T. Yoshino, S. Saito, and B. Sobukawa, Japanese Patent 7,103,438, 1971. RK. Grasselli, G.H. Heights, and IL Callahan, US. Patent 3,414,631, 1968. R.K. Grasselli, G.H. Heights, and HF. Hardman, US. Patent 3,642,930, 1972. JD. Buntington, C.T. Hartisch, and R.K. Grasselli, J. Catal., 87, 363, 1984. R Cantani, G. Centi, and F. Trifiro, Ind Eng. Chem. Res., 31, 107, 1992. M.M. Bhasin, "Methane Conversion," Elsevier Science Publishers, Amsterdam, p. 343, 1988. H.L. Mitchell III and RH. Waghorse, US. Patent 4,172,810, 1979. W. Bytyn and M. Baems, Appl. Catal., 28, 199, 1986. J.A. Sokfranko, J.J. Leonard, and CA. Jones, J. Catal., 103, 302, 1987. CA. Jones, 1]. Leonard, and J.A. Sokfi'anko, J. Catal., 103, 311, 1987. J.A. Sokfranko, 11 Leonard, and CA. Jones, Energy and Fuels, 1, 12, 1987. GE. Keller and M.M Bhasin, J. Carat, 73, 9, 1982. K.D. Campbell and J.H. Lunsford, J. Phys. Chem, 92, 5792, 1988. Y. Chen, H.T. Tohver, J. Narayan, and M.M. Abraham, J. Phys. Chem, 16(12), 5535, 1977. C-H. Chen, J-X. Wang, and J.H. Lunsford, J. Catal., 111, 302, 1988. J.M. DeBoy and RF. Hicks, J. Catal, 113, 517, 1988. CHH Lin, K.D. Campbell, and J.H. Lunsford, J. Phys. Chem, 90, 534, 1986. M.-Y. Lo, S.I(. Agarwal, and G. Marcelin, J. Catal., 112, 168, 1988 K. Otsuka, Y. Murakami, Y. Wada, A.A. Said, and A. Morikawa, J. Catal., 121, 122, 1990. K. Ostuka and K. Jinno, Inorg. Chim Acta, 121, 237, 1986. K. Ostuka, A.A Said, K. Jinno, and T. Komatsu, Chem. Lett, 77, 1987. Kirk-Othmer, ”Encyclopedia of Chemical Technology,” 3rd. ed., 3, p. 738, Wiley, New York, 1978. '1‘ K. Kojo, Japanese Patent 1,883,121,238, 1983. . S.T. King, J. Catal, 131, 215, 1991. Catalysts for the Control of Automotive Pollutants: Advances in Chemistry Series 143; JD. Mc Evory, ed.; American Chemical Society: Washington, DC, 1974. EH. Foules, J.H. Labinger, J.L. Beauchamp, and B. Fulty, J. Phys. Chem, 95, 7393, 1991 S. Lacombe, J.G. Sanchez, M.P. Delidore, H. Mozzanega, J.M. Tatibouet, and C. Mirodatos, Catal. Today, 13, 273, 1992. D. Dissanayake, M.P. Rosynek, K.C.C. Kharas, and J.H. Lunsford, J. Catal., 132, 117, 1991. M. Prettre, C. Eichner, and M. Perrin, Trans. Faraday Soc, 43, 335, 1946. I. Langmuir, Trans. Faraday Soc., 17, 672, 1922. T. Engel and G. Ertl, J. Phys. Chem, 69, 1267, 1978. PI. Berlowitz, HF. Peden, and D.W. Goodman, J. Phys. Chem, 92, 521, 1988. W.H. Weinberg, C.M. Comrie, and RM. Lambert, J. Catal, 41, 493, 1976. B. Mantolin and E. Gillet, Surf Sci., 166, L115, 1986. 11mm; ..1 h"! ht,” 1w- 45. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 65. 66. 67. 21 I. Keverekidis, L.D. Schmidt, and R.Aris, Surf Sci, 137, 151, 1984. HA. Jones and H.S.Taylor, J. Phys. Chem, 27, 623, 1923. G. Ertl, Surf Sci., 7, 309, 1967. F.P.H.M. Habraken, E. P. Keifer, and GA. Bootsma, Surf Sci, 83, 45, 1979. F.P.H.M. Habraken, C.M.A.M. Mestera, and GA. Bootsma, Surf Sci., 97, 264, 1980. IS Arlow and DP. Woodruff, Surf Sci, 180, 89, 1987. IR Prasad, L.A. Kennedy, and AB Ruckenstein, Cara]. Rev. Sci. Eng., 26, 1,1984. B.L. Yang, S.F. Chan, W.S. Chang, and Y2. Chen, J. Catal., 130, 52,1991. RA Prokopowicz, P.L. Silveston, R.R. Hudgins, and DE. Irish, React. Kinet. Cami. Lett., 37, 63, 1988. T-J. Huang,T-C. Yu, and S-H. Chang, App]. Carat, 52, 157, 1989. F]. Berry and ME. Brett, J. Catal., 88, 232,1984. G. Centi, D. Pesheva, and F. Trifiro, Appl. Caral., 33, 343, 1982. G. Centi, RK. Grasselli, E. Patane, and F. Trifiro,"New Developments in Selective Oxidation," Stud. Surf Sci. Catal., 55, 515, 1990. GE. Keller and M.M Bhasin, J. Catal., 73, 9, 1982. IT. Ali Emesh and Y. Amenaniya, J. Phys. Chem, 90, 4785, 1986. W. Hinsen, W. Bytyn, and M. Baems, "Proceedings 8th Intr. Congr. Catal., Berlin, 1984,” 3, p.581, Dechema, FrankfiJn-am-Maion, 1984. J. Driscoll, W. Martin, J.X. Wang, and J.H. Lunsford, J. Amer. Chem. Soc., 107, 58, 1985. T. Ito and J .H. Lunsford, Naturedondon), 314, 721, 1985. V. Fattore, Z.A. Fuhnnan, G. Manara, and BJ. Notari, J. Catal., 37, 215, 1975. IL Mikhailova, I.S. Sazonova, N.P. Keier, and TV. Belodtseva, Kinet. Karat, 9(3), 565, 1968. FM. Ali-Zade, N.M. Mardanova, PM. Sorgin, RM. Talshinskii, A.A. Medzhidov, and RG. Rizaev, U.S.S.R. Patent 49,052,169, 1991. S. Blumrich, R. Brand, B. Engler, W. Honnen, and E. Koberstein, European Patatent 410,440, 1991. Y. Ohno and K. Morishita, Japanese Patent 49,052,169, 1974. Chapter 2 Preparation and Characterization of Mixed Metal Oxide Catalysts 2.1. Preparation of Catalysts Recent advances in the preparation of catalysts illustrate the importance of carrier, impregnation technique, calcination temperature, and activation treatment on the structure and properties of heterogeneous catalysts [1-3]. A number of different types of interactions between an active metal oxide and the support are possible (Figure 2.1). Among them are the weak interaction between the supported oxide and the carrier, aggregates of deposited oxide electronically interacting with the carrier, metal oxide monolayers, covalently or ionically bound aggregates, solid solution of the active phase in the support, and new surface compounds. The objective of this research is to employ sophisticated synthetic procedures to control catalyst structure, thoroughly characterize the surface and bulk of the catalyst, and correlate catalyst structure with catalytic properties. The term impregnation generally refers to a procedure whereby a certain volume of solution containing a compound of the active element is contacted with a high surface area support. The most common methods of impregnation include wet impregnation, incipient wetness, grafting, chemical vapor deposition, and co-precipitation. Wet impregnation usually involves contacting a porous carrier with a solution of the active element to produce a suspension. The suspension is dried and the catalyst is activated by converting the dried phase to its active form through various physical and chemical changes (e.g., oxidation or reduction). Incipient wetness involves impregnation of the carrier with a solution whose volume corresponds to the total pore volume of the 22 23 ' 2 cl .....;9. r 5 , 379—19: i—ououo‘ ' , , . ° ‘, it 'a .1. o A m y I / ///;'// / .‘l // /,/_ ./I ,’ [0 'A O ‘ 0 A o l‘ 1 /1 1 1,1 1 1 1 1 1' ’ 1/1/ 1. 1 1'" 1‘ 1’ 1 1 1 A J 3 Weak iorces 13 Junction c Monolayer r ' . I . J i A 4 A 1 A A‘ ' A A 4 J a Chemically bound e Soho sotution t New compound aggregate 1n support with support Figure 2.1. Various types of active oxide-support interaction. (From Preparation of Catalysts II, 439, 1979). 24 carrier [8]. The loading of the active element is controlled by varying the concentration of the impregnating solution. Chemical vapor deposition. (CVD) involves the use of volatile inorganic or organometallic compounds to prepare the catalyst. This method has been found to produce well-dispersed active components on a variety of carriers [4—7]. The active and supporting oxides or their precursors may be co-precipitated from a solution containing compounds of each element [8]. This usually produces an intimate mixing of the active phase and the support. From this close proximity with the support, the active phase is dispersed throughout the bulk as well as the surface of the catalyst. Other less involved methods of impregnation include heating a physical mixture of the oxides to induce spreading of the active phase over the supporting oxide [9]. Recently, grafting reactions have been used to obtain monolayers of various active phases [10]. This preparation method involves a chemical reaction between the active element (often in the form of an alkoxide) and the hydroxyl groups of the support. One of the alkoxide ligands reacts with a surface hydroxyl of the support to produce a support- active phase bond and the corresponding alcohol. During calcination of the impregnated oxide, alkenes are formed from the destruction of the surface phase. The typical reaction used to produce vanadium oxide monolayers is shown in Figure 2.2. The grafting method usually involves wet impregnation of the support with a nonaqeous solution of the active element which is then filtered and the impregnated catalyst (filtrate) is washed. This impregnation procedure has been modified slightly for this work by employing incipient wetness impregnation of the appropriate solution. 25 i—C4H9—0 O—i—C4H9 1.1—OH .1—c,H9_o_v=o _. Al—O—Y=O . i- C4H90H i—C4H9—0 O—i—C4H9 Q-i-C4H9 OH Al-O-Y=0 AT Al—O—\:7=0 . ZC4H3 O-i—C4H9 OH o—i—cmg Al-O—V=O on / o-i—c.,H9 /Al—O—Y=0 0 AT 0 o +4C4H8+H20 9'i-C4H9 \Al-o—ir=o “TO—Yd) 0H o-i—c,H9 Figure 2.2. Reaction of metal alkoxide with support. (From J. Catal., 101, 1, 1986). 26 2.2. Characterization of Heterogeneous Catalysts 2.2.1. Bulk Characterization In order to understand the effect of preparation method, composition and pretreatment on the structure of heterogeneous catalysts, a variety of bulk characterization methods should be used. The bulk characterization methods that provide important information in this study are the Brunauer-Emmett-Teller (BET) method for measuring surface area, X-ray diffraction (XRD), nuclear magnetic resonance (NMR), infrared spectroscopy (IR), and electron paramagnetic resonance (EPR). Surface Area. Catalyst surface areas can be determined using the Brunauer- Emmett-Teller (BET) method. In this method, the Langmuir adsorption isotherm is extended to multilayer adsorption. The first layer is as assumed in the Langmuir adsorption isotherm. For subsequent layers, the rate of adsorption is assumed to be proportional to the fi'action of the lower layers which are vacant. The summation over an infinite number of adsorbed layers is defined by the following equation [11]: P _ 1 +(C-1) v(1>o —P) vmc vaPo (1) where C is a constant related to the heat of adsorption and liquefaction of the gas, Vm is the monolayer capacity (cm3 STP), V is the volume of the adsorbatc at STP (cm3/mole STP), P is the pressure of the adsorbatc, and P0 is the total pressure. X-ray Difi'raction. X-ray diffraction (XRD) can be used to identify crystalline phases with particle sizes greater than or equal to 3.0 nm. The particle size can be calculated from line broadening measurements using the Scherrer equation [12]: K}. E = BcosO (2) where d is the mean dimension of the particle, K is a constant (particle shape factor) taken as 0.9 (for cubic particles), [3 is the full width-half maximum of the diffraction line, 7» is the X-ray wavelength, and 9 is the angle between the atomic plane of both the incident and reflected beams. In addition, semi-quantitative XRD analysis can be performed in order to determine the weight percent of crystalline phases present in a catalyst series. In this procedure, a physical mixture of the phases are ground in order to produce a homogenous mixture. The areas of the peaks of interest are measured and used to generate a calibration plot. Comparison of intensity ratios measured for the catalyst and the physical mixtures can be used to determine the amount of crystalline phases present in the catalysts. This method requires the particle sizes of the active phase to be similar in the catalyst series and the physical mixture of components. Nuclear Magnetic Resonance. Solid-state NMR is able to provide qualitative information about chemically distinct sites, clarify physical states (coordination) [13], and also provide quantitative information [14]. Since only the local environment of the atom is probed, NMR is well suited for the structural analysis of systems that do not have long range order. The type of interactions involved in solid-state NMR are direct dipole-dipole coupling between nuclei, interaction of the nuclei with electrons in the environment creating chemical shifts, and quadrupolar interactions with electric field gradients. Infrared Spectroscopy. Infrared spectroscopy (IR) has been useful for obtaining information about the structure of catalysts and species adsorbed on the catalyst surface. IR has wide applicability in the study of dispersed metal or oxide catalysts. In addition, spectra can be readily be acquired at high temperature and pressure. Thus one may study 28 catalysts under realistic reaction conditions and subsequently obtain information about the nature of active sites in the catalysts. Electron Paramagnetic Resonance. Electron paramagnetic resonance (EPR) has been extensively used to study paramagnetic species of interest in catalysis such as supported metal ions, surface defects, and adsorbed molecules and ions. The extent of information obtainable from EPR data varies from a simple confirmation that an unknown paramagnetic species is present to a detailed description of the bonding and orientation of a surface complex. Factors such as spin-spin interactions, crystal field interaction and relaxation time have a significant effect on the spectrum [15]. The high sensitivity of EPR can be useful for the study of low concentrations of active phase species. 2.2.2. X-ray Photoelectron Spectroscopy (XPS) A variety of surface analytical methods have been used to study heterogeneous catalysts. X-ray photoelectron spectroscopy (XPS) is one of the most powerful surface sensitive techniques for the study of electronic structure of filled levels of surface atoms and adsorbates. The process occurring in XPS can be illustrated as follows [16]: 2 p /i i/ /i i/ /i i 2 s /i i/ h‘D 1 s 1 XPS Figure 2.3. XPS process A+hv——>A++e' (3) E(kinetic) = E(photon) - E(binding) (4) The kinetic energy of the photoemitted electron is essentially the difference between the energy of the incident photon and the binding energy of the electron. Due to a variety of processes in solids, equation 4 is more appropriately written as follows [17]: EKE =hv—Eb+r-:,+E, (5) 30 where E1. is the intra-atomic relaxation energy and E, is the extra-atomic relaxation energy associated with the solid environment. These relaxation shifts result when a core hole is created by photoionization. When this occurs, other electrons relax to lower energy states and partially screen the hole and thus make more energy available to the outgoing photoelectron. XPS can be used to obtain qualitative and quantitative information about surface species. Even though XPS is dominated by atomic rather than solid-state effects, the local electronic environment in the solid can influence the observed peak position and line shapes through both initial and final state effects. Initial state effects refer to shifts in the original binding energy due to changes in the electronic environment of the atom. Final state effects are shifts associated with intra- and extra-atomic relaxations [17]. The signal intensity, 15, of XPS can be related to the concentration by the following expression: 11 = 1071.1)(81 ”(81) (6) where 19 is the X-ray flux, m is the concentration, oi is the photoelectron cross section [18], D(ei) is the efficiency of the analyzer, and M6,) is the mean free path of the photoelectron. Inelastic mean free path or escape depth (3.) of low energy electrons in the solid is a key parameter in describing the surface sensitivity of XPS. To obtain the mean free path the following expression is used [19]: 8 ME) = [a(ln e + b)] (7) 31 where M8) is the mean free path of the excited electron with energy 8 and a and b depend on the electron concentration of the solid. In order to obtain quantitative information about catalysts, a number of models have been proposed to relate active phase/support peak intensity ratios to catalyst structure. It has been shown by Defosse et al. [20] that one can calculate the theoretical intensity ratio (Ip°/Is°) expected for a supported phase (p) atomically dispersed on a carrier (8). Kerkhof and Moulijn [21] extended the Defosse model and derived expressions based on a model catalyst that consists of sheets of support with cubic particles of active phases deposited between the support layers (Figure 2.4). For species dispersed as monolayers, the relationship is given by the following equation: 1° -13 p (.2) D(ep)op13,(1+e 2) (3) 17’- - s b D(e,)o,2(1-e'52) s monolayer where ID is the intensity of the electrons from element p in the supported phase, Is is the intensity of the electrons from element 3 in the carrier, GP and as are the photoelectron cross sections of the respective elements, D(e)p,s are the detector efficiencies, ens are the kinetic energies of the electrons, and [51.2 is the thickness (t) per mean escape depth of the photoelectrons from the support (its) or from the promoter (hp). For crystallites, the equation can be simplified to: . \\\\\\\ 2\\\\\\\\\\ S I 33 [11] = 1.0") 1L. (9) I 1° d s cm] J 3:. where (II’/196W is the experimentally determined intensity ratio, (Ip°/Is°) is the theoretical monolayer intensity ratio calculated using equation 9, d is the length of the edge of the cubic crystallites of the deposited phase, and XP is the mean escape depth of the photoelectron in the promoter [22]. The ability of XPS to determine active phase particle size independent of chemical state (metal or metal oxide) offers distinct advantages over more conventional methods of catalyst characterization. For example, chemisorption may be used to determine the average particle size of supported metals; however, the application of this technique to the study of supported metal oxides has been limited to a few select systems. Using XRD, it is possible to determine crystalline phases only if the particle size is 2 3.0 nm and the phase is present in sufficient quantities. An important limitation of XRD is that it cannot be used to study highly dispersed or amorphous phases. Electron microscopy may not be useful for the characterization of mixed oxides due to the limited contrast between the support phase and the canier. An important capability of surface spectroscopy is the ability to observe changes in the surface structure. Surface reactions provide useful information for understanding changes in the surface structure that result during reactions (Figure 2.5). Species which strongly interact with the support tend to undergo fewer chemical and structural changes during reactive gas treatment in comparison to weakly interacting phases. By choosing reactive gases that are involved in oxidation or ammoxidation reactions, it is possible to 34 study the reactivity of catalyst phases under conditions similar to those in actual reaction conditions. In situ surface characterization of catalysts reacted with reducing molecules such as H2, NH3, CO and hydrocarbons will be performed using a preparation chamber attached directly to the surface science instrument (Figure 2.6). 2.2.3. Catalytic Activity The catalytic activity experiments reported in this work were performed in a differential reactor as shown in Figure 2.7. As shown in the figure, the gases are passed through a zeolite bed to remove water. The flow of the gas is controlled by mass flow controllers. The gas mixture can either be routed to the reactor or can bypass the reactor for direct analysis of the gas stream prior to reaction. By correlating activity with the catalyst structure it will be possible to identifyI active sites of the catalysts. 35 II' from , /M(A) _"=_. M(A) reacts I. to form NP 1.13 E M(A) E "(3) CO I“ from M(B) we) M(B) reacts - to form M° all) <— Binding energy C) Figure 2.5. Protocol for study of heterogeneous catalysts by reactive gas treatment combined with XPS. (From Anal. Chem, 58(12), 1184 A, 1986). 36 Figure 2.6. Reaction chamber. h ”t 37 FURNACE La 59 3-WAY VALVE “it t arLJL z-zaoura Figure 2.7. Catalytic reactor. w-Massnowcomaorm m-GASOROMAM air-mm m-mromnonneracroa rep-Wm . fATr _.. tmtfl‘lfi' 38 References 1. F. Roozeboorn, D.D. Cordingley, and P]. Gellings, J. Catal., 68, 464, 1981. 2 LE. Wachs, R.Y. Saleh, S.S.' Chan, and CC. Chersich, App]. Catal., 15, 339, 1985. 3. PL. Villa, ”Catalyst Deactivation,” B.Delmon and GP. Fromont ed., Elsevier, Amsterdam, p. 103, 1980. 4. J.C.W. Chien, J. Amer. Chem. Soc, 93, 4675, 1971. 5. H. Praliaud and M.V. Mathieu, J. Chim. Phys, 73, 689, 1976. 6. V.A. Khalif, E.L. Aptekar’, O.V. Krylov, and G. Ohlmann, Kinet. Catal. (USSR), 18,867,1977. 7. F. Roozeboorn, T. Fransen, P. Mars, and PI. Gellings, Z. Anorg. Allg. Chem, 25, 449, 1979. 8. E.T.C. Vogt, A.J. van Dillen, J .W. Geus, and F .J.J .G. Janssen, Catal. Today, 2, 569, 1988. 9. S. Shan and D.Hoenicke, Chem-lng.-Tech., 61, 321, 1989. 10. A. Baiker and A. Wokaun, Naturwiss, 76, 168, 1989. 11. ON. Satterfield, "Heterogenous Catalysis in Practice," New York, McGraw-Hill, 1980. 12. HP. Klug and LE. Alexander, "X-ray Diffraction Procedures for Polycrystalline and Amorphous Materials," New York, Wiley-Interscience Publications, 1974. 13. C. Dybowski, Chemetech., March, 186, 1985. 14. BC. Gerstein, Anal. Chem, 55(7), 782A, 1983. 15. W.N. Delgass, G.L. Haller, R. Kellerrnan, and J .H. Lunsford, "Spectroscopy in Heterogeneous Catalysis,” Academic Press, 1979. 16. D.M. Hercules and SH. Hercules, J. Chem. Ed, 61(6), 483, 1984. 17. DP. Woodruff and TA. Delchar,:Modem Techniques of Surface Science," F Cambridge, Cambridge University Press, 1986. 18. J. H. Scofield, J. Electron Spectrosc. Relat. Phenom, 8, 129, 1979. 19. D. R. Penn, J. Electron Spectrosc. Relat. Phenom, 9, 29, 1976. 20. C. Defosse, P. Canesson, P. G. Rouxlet, and B. Delmon, J. Catal., 51, 269, 1979. 21. F. P. J. M. Kerkhof and J. A. Moulijn, J. Phys. Chem, 83, 1612, 1979. 22. M. Houalla, in "Preparation of Catalysts III", Elsevier, Amsterdam, 273, 1983. Chapter 3 Vanadium Oxide/Aluminum Antimonate 3.1. Introduction Considerable research effort has been focused on the development of selective oxidation and ammoxidation catalysts for the activation of less expensive feedstocks such as propane, propylene and methane. Several researchers have reported that vanadium antimonate catalysts are promising systems for ammoxidation of propylene [1] and propane [2] to acrylonitrile. Typical industrial catalysts are prepared by mixing a hydrosol or gel of the carrier with a shiny or solution containing the active phase(s) and the promoters, such as W and/or Mo. This procedure can produce very complicated catalyst structures that make comprehensive characterization difficult. Thus, the effect of surface structure on the properties of the catalyst and active site(s) are not well known. In order to understand the influence of the catalyst structure on alkane or alkene activation, a simple catalytic system must be examined. Centi et al. [3] have investigated the structure and selective oxidation and ammoxidation of propane activity for VSbOx mixed oxides supported on SiOz, A1203, and Ti02 and promoted with W or Mo. Their work suggests that VSbOx phase plus Sb205 is not the active phase for ammoxidation. Instead, they proposed that AleO4 rutile phase modified with V is the active phase for the formation of propylene and V or other redox elements having M=O bonds (W or Mo) are needed to form acrolein or acrylonitrile. Although VSbOx/A1203 is not as complex as the commercial catalysts, further understanding of propane and propylene activation would be obtained by studying the interaction of the pure phases believed to be responsible for selective oxidation and ammoxidation. 39 . ‘:'l.'1"a‘ 40 Recently, Volta et a1. [4,5] have examined the structure and propane oxidation activity of vanadium oxide on AINbO4. AleO4 was chosen because of the structural similarity between AleO4 and T102. These researchers showed that it is possible to control the structure and reactivity of vanadia monolayers by modifying the local structure of the oxide support. They also found that the Ale04 disorganized structure is present and has good selectivity for the oxidative dehydrogenation of propane into propylene. This work will correlate surface structure with propylene oxidation activity for V oxide catalysts supported on AleO4. The focus will be on defining the nature of the V- AleO4 interaction and influence of V dispersion and reactivity on the catalytic properties of the mixed oxides. In order to determine structure and reactivity of the species in the mixed oxide catalysts, X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), infrared spectroscopy (IR), solid-state nuclear magnetic resonance (NMR), and electron paramagnetic resonance (EPR) will be used. Structural information will be correlated with propylene oxidation activity measurements in order to establish the nature of propylene activation site(s) in V/AleO4 mixed oxide catalysts. 3.2. Experimental Catalyst Preparation. AleO4 was prepared by coprecipitation of a 1:1 atomic ratio of Al(NO3)3-9H20 (J.T. Baker, A.C.S. grade) dissolved in deionized water and SbC15 (Aldrich, 99%). SbC15 was added to a vigorously stirred solution of A1(NO3)3 over a 30 minute period. NH40H (J .T. Baker, A.C.S. grade) was added until the solution pH reached 6-7. The neutralized solution was allowed to stir for 1 hour. The precipitate formed was filtered and washed with deionized water, dried in air at 120°C for 12 hours, ground to produce a homogeneous mixture, and calcined in air at 750°C for 12 hours. The resulting AleO4 support had a pore volume of 0.1 ml/g and a surface area of 155 m2/g. Prior to impregnation, the support was calcined at 500°C for 12 hours. 41 .V/AleO4 catalysts were prepared by pore volume impregnation of Ale04 using vanadium triisopropoxide oxide (Alfa, 95-99%). The catalysts were dried under nitrogen for 48 hours at 25°C, dried in air for. 12 hours at 120°C and calcined in air at 500°C for 12 hours. The vanadium content varied from O to 25 wt % V205 (V/Al atomic ratio of 0 to 0.78). Catalysts will be designated V# where # is the weight percent vanadium as V205. Standard Materials. V02 (99.9%) and V203 (99%) were obtained fiom Aldrich. V205 was prepared by calcining NH4VO3 (Johnson Matthey, 99.99%) in air at 450°C for 4 hours. Sb205 was prepared by the addition of SbC15 to deionized water under vigorous stirring. The pH of the solution was adjusted to 6-7. The precipitate formed was filtered and washed with deionized water, dried in air at 120°C for 12 hours and calcined at 500°C for 12 hours. XRD patterns of all standard compounds matched the appropriate ASTM powder diffraction file. BET Surface Area. Surface area measurements were detemtined using a QuantaChrome Quantasorb Jr. Sorption System. Approximately 0.10 grams of the catalyst was outgased between 165°-170°C for 12 hours prior to absorption measurements. Measurements were made using relative pressures of N2 to He of 0.05, 0.08, and 0.15 (surface area of N2 = 0.162 nm2) at 77 K. Data were processed using a Macintosh computer. X-ray Drflraction. X-ray powder diffraction patterns were obtained using a Rigaku XRD diffractometer which employs Cu Kor radiation (1.541838 A). The x-ray was operated at 45 kilovolts and 100 milliamps. The patterns were scanned between 10°-75° (2 theta) at a scan rate of 0.1 °/min. with DS and SS = 0.5. Powdered samples were mounted on glass slides by pressing the powder in an indentation on one side of the slide. Particle sizes were determined from line broadening measurements using the Scherrer equation [6]: (1) where d is the mean dimension of the particle, K is a constant (particle shape factor) taken as 0.9 (for cubic particles), 6 is the hill width-half maximum of the V205 <001> difi‘raction line, 3. is the X-ray wavelength, and 0 is the angle between the atomic plane of both the incident and reflected beams. Semi-quantitative XRD has been performed using physical mixtures of V205 and Ale04 (0.2 wt %, 0.5 wt %, and 1.0 wt % V205). The areas of the V205 <001> and Ale04 <110> peaks were measured and used to generate a calibration curve that can be used to determine the amount of crystalline V205 present in the V/AleO4 catalysts. Peak areas were obtained using Googly software [7] and a linear background was assumed over the peak base. X-ray Photoelectron Spectroscopy. X-ray photoelectron spectroscopy (XPS) data were obtained using a Perkin-Elmer spectrometer that is equipped with a Mg (1253.6 eV)/Al (1486.6 eV) dual anode and a 10-360 hemispherical analyzer with an omnifocus small spot lens. Samples were mounted as powders on double-sided sticky tape or spray coated on glass slides. Spray coated samples were prepared by air brushing a 10% suspension of the catalyst (20% deionized water and 80% acetone) onto a glass slide heated at 60°C. XPS binding energies of the catalyst were referenced to the Al 2p (74.5 eV) peak. The binding energies for standard compounds were referenced to the C Is (284.6 eV). Peak areas and binding energies were obtained using Googly software. XPS binding energies were measured with a precision of d: 0.2 eV, or better. In situ XPS analyses of spray coated V/AleO4 were performed using a reactor attached directly to the spectrometer. The samples were initially analyzed, then reacted first with 100 cc/min mixture of 8.0% 02 (99.98%) and 92% He (99.9%) for 1 hour at 43 350°C and then with 100 cc/min 1.7% propylene (99.9%), 8.0% 02 and 90.3% He for 10 minutes at 200°C. Quantitative )G’S. A number of models have been proposed to relate active phase/support XPS peak intensity ratios to catalyst structure. It has been shown by Defosse et a1. [8] that one can calculate the theoretical intensity ratio (If/15°) expected for a supported phase (p) atomically dispersed on a carrier (5). Kerkhof and Moulijn [9] extended the Defosse model and derived expressions based on model catalysts that consist of sheets of support with cubic particles of active phase deposited between the support layers. The photoelectron cross-sections (0’) and mean escape depths (k) of the photoelectrons used in the calculations are taken from Scofield [10] and Penn [11], respectively. For species dispersed as monolayers, the relationship is given by the following equation: 1?, =(E) D(ep)ch,(1+e'l32) (2) I? b D(es)052(1-e-52) S monolayer where Ip° is the intensity of the electrons from element p in the supported phase, 15° is the intensity of electrons from element 5 in the carrier, GP and as are photoelectron cross sections of the respective elements, D(ep’s) are detector efficiencies, 8w are the kinetic energies of the electrons, and 61,2 is the thickness (t) per the mean escape depth of the photoelectron fi'om the support (3.5) or the promoter (hp). This model predicts a linear increase in the intensity ratio of the supported phase (p) to the carrier (5) as the (p/s) atomic ratio is increased. Infrared Spectroscopy. Infiared spectra were obtained using a Mattson 3020 FTIR spectrometer with a DTGS detector. The catalyst was suspended in a 80% acetone 44 20% water mixture and sprayed on a 13x2 mm KBr disk heated at 60°C. Spectra were obtained with 2 cm'1 resolution and both background and sample had 32 scans performed. Data acquisition were performed using a FIRST software package. Data manipulation were performed using Googly software. A linear background was subtracted to produce Figure 3.6. Nuclear Magnetic Resonance. Solid-state nuclear magnetic resonance experiments were performed using a Varian VXRS 400 spectrometer. The powdered catalysts were packed into a zirconium oxide roater. The samples were spun in a Bruker MAS probe at 4 kHz. The frequency used for aluminum catalysts was 104.230 MHz. Wide-line solid-state NMR experiments were also performed. The amplifier is fi'om American Microwave Technology and the probe is a standard Varian wide-line probe. The powder samples were placed in a quartz tube and then placed in the wide-line probe. The frequency used for the vanadium catalysts was 105.152 MHz. A SUN workstation was used to operate the spectrometer and process the data. Electron Paramagnetic Resonance. Electron paramagnetic resonance data for V/AleO4 catalysts were obtained using Bruker ER 300D spectrometer. A Hewlett Packard 5245L electronic counter with a 5255A frequency converter plug in and a Bruker ERO35M gaussmeter were used to measure microwave frequency and magnetic field, respectively. The cavity used was a rectangular TEroz- The samples were run as powders after reaction in 1.7% propylene, 8.0% 02 and 90.3% He. Propylene Oxidation Activity Measurements. Catalytic activity experiments were carried out in a differential type flow reactor at low conversion (<10%) with Oz/propylene/He (8.0/l.7/90.3%) at a temperature of 250°C. The catalysts were pretreated at 350°C for 3 hours with a 8.0% 02/92% He mixture. For both pretreatment and reactions, the flow rate was 100 cc/minute. Approximately 0.2 grams were used for each reaction. Reactions reached steady-state after 11 hours. The gas flow rates were held constant with Brooks 5850 and Porter 201 mass flow controllers. The reactor 45 temperature was controlled with an Omega CN 1200 controller. Product gases were analyzed with a Varian 3700 gas chromatograph equipped with a TCD (for C0, C02) and FID (for hydrocarbon, acrolein, and acetaldehyde). The column used for permanent gas separation was a 5 ft. 60/80 mesh Carboxen (Supelco) column. For hydrocarbon and acrolein separation a 5 it 80/ 100 Porapak QS column was used. The chromatograph was interfaced to a Hewlett-Packard 3394A integrator. The column temperature was programmed from 55°C to 190°C with analysis time of 30 minutes to allow complete separation of all components. Selectivities are reported for catalysts operated at similar conversion. 3.3. Results Ale04. The BET surface area of the Ale04 support is given in Table 3.1. The XRD powder pattern of the support (Figure 3.1) matches results reported in the literature [12]. The Ale04 particle size determined using x-ray diffraction line broadening calculations (equation 1) is 2.7 nm. No other peaks due to antimony oxides or aluminum oxides were observed. 2"Al MAS solid-state NMR of the aluminum antimonate support has been performed and peaks at 1, 34, and -31 ppm in the NMR spectrum were observed (Figure 3.2). The FTIR spectrum of the support was also obtained and no peaks were observed between approximately 920 cm'1 and 1050 cm'1 (the region of interest of the vanadium oxide). The XPS Sb 3d3,2 binding energy measured for the support was found to be 540.5 eV. The XPS binding energy for Sb205 was 540.5 eV. The XPS Sb3d3,2/Al 2p intensity ratio indicates that Sb/Al atomic ratio in the surface of the Ale04 support is 1.0. Vanadium/Aluminum Antimonate. Variation in the BET surface area of the V/AleO4 catalysts as a function of vanadium content is shown in Table 3.1. The surface area decreases from 140 to 108 m2/g as the WA] atomic ratio increases from 0.12 to 0.78 46 Table 3.1. BET surface area (mZ/g) of V205/N5b04 catalysts. Atomic Ratio V/Al Weight % V205 Surface Area (ngLI 0 0 155 0.12 5 140 0.26 10 1.32 0.41 15 123 0.58 20 1 11 0.78 25 108 47 1 1 m4 1 I <110>_ AleO4 <101> <001> a b c l l 1 I 10.0 20.0 30.0 40.0 DEGREE (2 THETA) Figure 3.1. XRD of V20,/AISb04 a) V25, b) V20, and c) V0 catalysts. 48 ITTTrf‘YVITYTY‘IWY'YIYVY rrfifrrrr . r . h 300 200 100 0 -100 -200 -300 1’1”“ Figure 3.2. 27Al MAS NMR of a) V25 and b) V0 catalysts. 49 (5 wt.% to 25 wt.% V205). However, corrections made for active phase loading indicate that the surface area of the support does not change. For V/AleO4 catalysts with V/Al atomic ratios of < 0.78, XRD patterns showed peaks characteristic only of the Ale04 carrier (Figure 3.1). The diffraction pattern measured for the V25 catalysts showed peaks characteristic of Ale04 and crystalline V205. For the V25 catalyst, XRD line broadening calculations (equation 1) indicated that the particle size of the V205 crystallites was 29 nm (Table 3.2). Figure 3.3 shows the variation of the V205 <001>/AISb04 <110> intensity ratios measured for 0.2, 0.5 and 1.0 wt.% V205/AleO4 physical mixtures. The value measured for the V25 catalyst with is also given. The results indicate that the amount of V205 crystallite in the catalyst is less than 0.1 wt.% V205. XPS V 2p3,2 binding energies measured for the catalysts, run as powders, were found to be 517.5 a: 0.2 eV. The XPS V 2pm binding energy measured for pure V205, V02 and V203 were 517.5 eV, 517.3 eV and 516.5 eV, respectively. The binding energies of the Sb 3d3,2 for the series of catalysts were found to be 540.4 :1: 0.2 eV. The binding energies for both vanadium and antimony are independent of vanadium loading. XPS V 2pm binding energies measured for the spray coated and propylene/OZ/He reacted catalysts were 517.3 :h 0.2 eV. For 02/He reacted catalysts the V 2pm binding energies were 517.0 i 0.3 eV. The Sb 3d3,2 for all the samples under each condition was 540.3 :1: 0.2 eV. Variation in the position of the vanadium peak as a function of reaction condition measured for the V25 catalyst is shown in Figure 3.4. Figure 3.5 shows the variation in the V 2p3,2/Al 2p XPS intensity ratio measured for powder, spray coated, oxidized, and reacted catalysts as a fimction of WA] atomic ratio. There is a linear increase in intensity ratios for the spray coated and reacted samples up to a V/Al atomic ratio of 0.58. Between an V/Al atomic ratio of 0.58 and 0.78, there is a sharp decrease in the WA] intensity ratio measured for the spray coated and reacted samples. For the powdered samples, there is a linear increase in the WA] intensity ratio as 50 0.08 -- I " Physical Mixtures , / — "' " Best Fit / / 0.06 " t 25 wt.% / ‘ V/AleO4 / / 7p / _o_ / V / v / 3 ) / 3 / S. 0 04 ‘* / A / g / n / c; / > / ' -‘-" / / / / 0.02 -- / / / / / / / A a / o ./ 1 1 1 t 1 0 0.2 0.4 0.6 0.8 1 Weight Percent V205 Figure 3.3. Semi-quantitative XRD for physical mixtures of V205 and Ale04. 51 Table 3.2. Particle size for V205/AleO4 calculated from XRD data. XRD Particle size (nm) V/Al Atomic Ratio XRD(nm) 0.12 J 0.26 _‘ 0.41 —" 0.58 J ‘ 0.78 29 ' - No V205 peaks observed. Thus, no V205 particle size can be calculated. 52 as the V/Al atomic ratio increases. For catalysts with V/Al atomic ratios < 0.58 and V25 catalyst the intensity ratios for the powder, spray coated, and reacted samples are similar, within experimental error. The V20 V/Al intensity ratios measured for spray coated and reacted catalysts were identical, within experimental error. However, the WA] intensity ratio measured for the powder catalyst was significantly lower than those obtained for the spray coated and reacted catalysts. IR. Figure 3.6 shows the IR spectra measured between approximately 920 cm'1 and 1050 cm"1 for the catalysts. The spectrum measured for the V205 is shown for comparison. The spectra contains a broad band centered at 985 cm"1 for all loadings. As the V/Al atomic ratio increases from 0.26 to 0.41, another band appears at 995 cm‘l. A band at 1022 cm"1 (which is due to crystalline V205 [13]) is observed in the spectrum measured for the V25 catalyst. Nil/IR. The 27Al MAS NMR spectra measured for the aluminum antimonate support and the vanadium catalyst series show one main peak at approximately 1.0 ppm and additional peaks at 34 ppm and -31 ppm (Figure 3.2). Figure 3.7 shows the wide-line solid-state NMR spectra obtained for the V5 and V25 catalysts measured at a frequency of 105.152 MHz. The vanadium peak was found to shift by 40 ppm from the 5 wt.% to the 25 wt.% catalysts. EPR EPR spectra were measured for the catalysts calcined at 500°C and after exposure to reaction conditions. The calcined catalyst showed no signal in the EPR spectrum. However after reaction, the catalyst showed a very broad signal due to V4+ with a g value of 1.98 (Figure 3.8). Hyperfine structure was observed in the spectrum measured for the V5 catalyst. However because of the concentration of the vanadium, detailed information could not be obtained. Dilution of the catalysts with support did not improve the spectra obtained. Propylene Oxidation. Table 3.3 shows the average percent conversion and selectivity to acrolein, acetaldehyde, C0, and C02 measured for the V/Ale04 catalysts. l l l 1 520.0 515.0 515.0 514.0 BINDING ENERGY (eV) Figure 3.4. XPS spectra of V25 catalyst a) in propylene/OZ/I-Ie, b) 02/He, c) spray coated, and d) powder. 54 6 .1.- ' Spray Coated e S '1'- . Ozme A PropyleneJ02/He Powder CE}- 4 .1 e 3 555 z a: . I N Z. 3 ‘ 2 .. l 0 1 1 1 1 1 1 1 a 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 V/Al Atomic Ratio Figure 3.5. XPS intensity ratio of V 2pm/A1 2p as a function of WA] atomic ratio. 55 __ 1‘) — _ l 1 1040.0 1WD 960.0 920.0 FREQUENCY (cm 4 ) Figure 3.6. IR data of a) V205, b) V25, c) V20, d) V15, e) V10, and 1) V5 catalysts. MT 1 Y Y I rY Y f1 1W 200 0 ~100 Figure 3.7. Vanadium Wide-line NMR of a) V25 and b) V5 catalysts. -200 l I I 1 2000.0 3000.0 4000.0 5000.0 FIELD (GAUSS) Figure 3.8. EPR V4+ spectra measured for reacted a) V25 and b) V5 catalysts. Spectra were collected at room temperature. 58 Table 3.3 Conversion and selectivity of V205/AleO4 for propylene oxidation at 200°C and 100 cc/minute flow rate. V/Al Percent Selectivity Selectivity C0 C02 Atomic Conversion to to Ratio Acetaldehyde 3 Acrolein b 0 0.5 37.1 59.5 3.0 <1.0 0.12 1.4 9.6 90.4 <0.1 0.1 0.26 2.2 13.6 86.2 <0.1 <0.1 0.41 3.3 17.3 82.3 <0.2 0.1 0.58 3.3 19.0 80.5 <0.2 <0.1 0.78 4.1 21.3 78.0 <0.5 <0.1 0.12 3.1 7.0 89.6 2.1 1.3 0.78 3.6 21.4 78.1 <0.4 <0.1 1' - Error bars of d: 20%. b - Error bars of i 10%. 59 The production of carbon oxides was the highest for the pure AleO4. As the amount of vanadium oxide increases, the percent conversion increases. The addition of 5 wt.% (V5 catalyst) vanadium oxide increases the selectivity to acrolein and decreases the amount of carbon oxides formed. For higher vanadium oxide loadings, the selectivity to acrolein decreases. When the conversion of the extremes of the catalysts were similar, the selectivity to acrolein for the V5 catalyst was 89.6% and for the V25 catalyst was 78.0%. The amount of C0 produced by the V5 catalyst increased from < 0.2% to 2.1% as the conversion increased. 3.4. Discussion Aluminum Antimonate. BET, XRD, and NMR results indicate that base hydrolysis of aluminum nitrate and antimony pentachloride followed by calcination in air at 750°C results in a pure, high surface area (155 m2/g) AleO4 support. The XPS Sb 3d3,2 binding energy measured for AleO4 carrier is consistent with the formation of Sb5+ expected for this compound. In addition, the Sb/Al atomic ratio determined fi’om XPS intensity ratios (1.0) suggests that no residual aluminum or antimony phases exist in the surface of the carrier. Volta et al. [4,5] have reported that coprecipitation followed by calcination at 750°C leads to the formation of crystalline AleO4. Structure of Calcined Vanadium/A luminum Antimonate Catalysts. For V/AleO4 catalysts with V/Al atomic ratio S 0.58, the absence of vanadium oxide peaks in the XRD pattern indicates that no crystalline vanadium oxide phases are formed in these catalysts. However, semi-quantitative XRD analysis suggests that the V25 catalyst contains small amounts of crystalline V205 (< 0.1 wt.% V205). The XPS V 2pm binding energies measured for the powder and spray coated V/Ale04 catalysts indicate that the vanadium is in the *5 oxidation state for all active 60 phase loadings. XPS Sb 3d3,2 binding energy data indicates that the antimony is in the +5 oxidation state for all vanadium loadings. The IR spectra measured for the catalysts suggest that monomeric forms of vanadium oxide are present in the catalysts because of the presence of a band centered at 985 cm"1 [13]. Polymeric forms of vanadium oxide are observed in catalyst with an V/Al atomic ratio 2 0.26 due to the presence of the band centered at 995 cm'1 [13]. For the V25, the observation of a peak at 1020 cm'1 (due to V205) [13] is consistent with XRD results that show the presence of V205 crystallites. Wide-line solid state NMR was used to determine the coordination of the vanadium in the catalysts. Eckert and Wachs [14] used wide-line NMR to study the local environments in two-dimensional vanadium(V) oxide surface layers on titania and alumina supports. Two main surface vanadium oxide species were detected with different bonding environments assigned to 4- and 6- coordinate V-O environments. Their results indicate that a marked dependence of the surface vanadium oxide structure on the metal oxide support material. For the V5 catalyst the vanadium is in a 4-coordinate environment. As the vanadium loading is increased to 25 wt.% (V25), the vanadium becomes 6-coordinate. This shift is as expected due to the change in the form of vanadium oxide from monomeric at low loadings to polymeric forms at high loadings as observed from IR data. MAS solid-state NMR on the V/Ale04 catalysts were identical to the spectrum measured for the support. The signal between 0-2 ppm is assigned to aluminum in the octahedral environment [15]. The peak at 34 ppm and -31 ppm has been attributed to pentacoordinated atoms or to aluminum in a tetrahedrally distorted coordination [16]. 2"Al MAS NMR spectra measured for both the support and catalysts show that there is no transformation of structure of the antimonate after the vanadium is added and the sample is calcined. Similar results were obtained by Volta et al. [17] for V0" on AleO4. The method of preparation used was designed to produce a single monolayer of V0x species on the Ale04 surface. In order to determine the theoretical monolayer 61 coverage, Roozeboom et al. [18] have defined 02.5 as the average area of supported surface which V015 unit occupies. For V205, this value is approximately 0.105 nmz. Thus, the monolayer capacity of the. AISb04 support is calculated to be 22.5 wt.% V205. From the XRD pattern we see that crystallite formation is observed between 0.58 and 0.78 atomic ratio of V205. This is as expected based on the estimated monolayer capacity. For V/AleO4 catalysts with V/Al atomic ratio 5 0.58, XPS and XRD data indicate that the vanadium is highly dispersed. XRD results show that large particles of V205 (~ 29 nm) are formed on catalysts with higher vanadium loadings. Figure 3.5 shows that the V 2pm/Al 2p intensity ratio increases linearly with vanadium loading up to a V/Al atomic ratio of 0.58. Between V/Al atomic ratio of 0.58 and 0.78 the intensity ratio decreases (observed for the spray coated and reacted samples). This suggests that the dispersion of the vanadium oxide decreases drastically at this level. In addition, this is the same loading for the formation of crystallites as observed in the XRD data and that estimated fi'om the monolayer capacity calculation. However, the powder sample of WA] of 0.58 does not follow the trend as seen for the spray coated and reacted samples. At present we can not determine the cause of the difference in WA] intensity ratios measured for the spray coated, reacted, and powder V20 catalysts. Samples with WA] 2 0.58 were prepared twice and the same trend was observed. IR results also suggest that the vanadium is highly dispersed due to the presence of bands characteristic of monomeric and polymeric vanadium oxide species. The band centered at 985 cm'1 is assigned to vanadate dispersed on the support as a monolayer. Distortion of the structure causes a shift of the V=0 stretching band fi'om 1020 to 985 cm'l [13]. The band centered at 995 cm'l is attributed to polymeric forms of vanadium on the support. The band at 1022 cm'1 is due to crystallite forms of V205. The absence of a band characteristic of vanadate (950 cm") suggests that no aluminum or antimony vanadates are present [17]. MAS NMR obtained for the support and the catalysts supports this finding. Nakagawa et a1. [13] studied monolayer species of V205/Ti02 and 62 found a shift fiom 1020 (crystallite V205) to 980 cm'1 for the V=0 stretching frequency. Their results lead to the conclusion that at low coverage V0x is amorphous and at high coverage both amorphous and crystalline V205 are observed. Similar results were obtained for V205 supported on T102, A1203, Zr02, and S102 deposited either fiom aqueous metavanadate or gaseous VOC13 precursors [19-21]. Bond et a1. [22] studied V205/T102 catalysts prepared by wet impregnation and grafting methods. They observed that as V205 concentration was increased, the XPS V 2p3,2/T i 2pm intensity ratio increased to approximately 7 wt.%, but after two monolayers were deposited the ratio remained approximately constant. These results suggest that once the V205 amount exceeds that necessary for formation of the first monolayer, the V205 is not dispersed uniformly, but crystallites of V205 cover a fraction of the monolayer surface. Based on XPS, XRD, IR and NMR vanadium oxide is present as monomeric species as the loading increase, polymeric forms of vanadium oxide are present and at 25 wt.% crystalline form of V205 are present. Structure of Reacted Vanadium/A luminum Antimonate Catalysts. The V/Al XPS intensity ratios for the reacted samples suggest that the dispersion of the vanadium does not change under reaction conditions. In order to determine if XPS can distinguish between V5+ and V4+ the binding energies of standard compounds have been obtained. Unfortunately, the binding energies of standard compounds have shown that it is difficult to distinguish between V5+ and V“. However, as seen from Figure 3.4, slight changes in the position of the vanadium peak for the powder, spray coated, propylene/Oz/He, and 02/He occurs, but the shifts observed are not as large as expected for a change in the vanadium oxidation state. The V 2pm binding energies reported in the literature for V205 vary fi'om 516.4 to 517.4 eV [23,24]. For V41“, represented by V204, the binding energies range from 515.4 to 515.7 eV [25,26]. Based on literature values, it appears that the vanadium in the catalyst is V5+. Unfortunately, XPS cannot unequivocally determine if a change in vanadium oxidation state occurs due to reaction. 63 EPR spectra clearly show that some of the vanadium is reduced from V5+ to V4+ under reaction conditions. Because of the high concentration of vanadium present in this system, it is not possible to obtain any useful information from the hyperfine splitting of the catalyst series. With conventional EPR, noise caused by the cavity was seen and at lower concentrations this signal was greater than that of the signal due to the vanadium. Thus, at present, we cannot obtain coordination of the vanadium as a firnction of loading and we cannot obtain quantitative information about the amount of reacted vanadium. Influence of Catalyst Structure on Activity and Selectivity. The V5 catalyst shows the greatest selectivity towards acrolein production. With the conversion extremes similar, the selectivity toward acrolein for the monomeric form is still greater than that with crystalline forms present. In addition, the pure aluminum antimonate support produced higher levels of carbon oxides than the promoted support. Activity measurements of the vanadium/aluminum antimonate catalyst suggest that monomeric forms of vanadium oxide are more selective in the production of acrolein than polymeric and crystalline forms of vanadium oxide. However, all forms of supported vanadium oxide catalysts showed greater selectivity toward acrolein than the aluminum antimonate support. By the addition of vanadium to the surface of the support, a redox element having a V=O bond is present. The presence of a terminal V=O site, the selectivity during the oxidation process has been shown to be effected [27]. It is generally believed that too weak M-O bonds results in non-discriminative C-O bond formation which leads to combustion. Ifthe M-O bond is too strong, lattice oxygen is unavailable to participate in the reaction [28]. In addition to M=O bonds and bond strength, geometric effects also play a role in that the active site must contain "the right amount" of oxygen [29]. In this system, the presence of the vanadium increases the selectivity to acrolein at all active phase loadings. However, as crystalline phases of V205 become present, an increase in CO" and the by-product, acetaldehyde is observed. This may be due to the reactive nature 64 of crystalline phases. In this system, as with MoO3-based mixed oxide catalysts, it is possible that the rate of allylic oxidation is controlled by the rate of allylic-intermediate formation [30]. Straguzzi et al. [31] studied the conversion of l-butene to butadiene on aluminum antimonate catalyst and found that the selectivity to butadiene was 75% at conversions less than 25% and showed activity of 1.0 mmol/mzh. Straguzzi suggests that due to the dimculty in reducing Al“, the activity of the aluminum antimonate can be attributed to Sb cations) However, the aluminum antimonate they prepared had a surface area of 51 m2/g and was higher in crystallinity. The aluminum antimonate that was produced in this work demonstrated a low selectivity to acrolein, 59.5%. The structure of the antimonate site for selective oxidation in uranium antimonate is composed of bridging oxygen species which connect two Sb5+ and two Sb3+ atoms. In this, the Sb3+-0 bridge serves as the H abstracting species and gives radical character to the oxygen bond at the site. This facilitates the H abstraction. The Sb5+-O moieties are used as olefin chemisorption sites and sites for the insertion of oxygen [32]. In order to determine which of these processes are occurring, firrther investigation is needed. 3.5. Conclusions The combined use of several techniques to investigate the structure and dispersion of V/AleO4 catalysts leads to the following conclusions: 1. XPS, XRD, IR, and NMR data indicates that Ale04 is a pure phase with no formation of aluminum or antimony oxide species present. 2. XPS, XRD, and IR data indicate that V is highly dispersed over the Ale04 carrier. XRD and XPS results suggest the trend of crystallite formation is consistent with estimated monolayer capacity calculations. 65 3. Based on XPS, data both Sb and V species are in the *5 oxidation state. XPS binding energies of V and Sb do not change as a firnction of loading or reaction conditions. . 4. IR and NMR data suggest that the vanadium oxide does not form a vanadate- type structure with the aluminum antimonate support. 5. EPR data indicates that vanadium is reduced fiom V5+ to V“. during propylene oxidation. 6. Activity data suggests that low loadings of vanadium oxide (V/Al S 0.26) increase selectivity to acrolein. Pure aluminum antimonate support shows lower selectivity to acrolein than vanadium oxide supported catalysts at all loadings. References fl F. J. Berry and M. E. Brett, J. Catal., 88, 232, (1984). 2. G. Centi, D. Pesheva, and F. Trifiro, App]. Catal., 33, 343, (1987). G. Centi, R. K. Grasselli, E. Patane, and F. Trifiro, in "New Developments in Selective Oxidation”, Stud Surf Sci. Catal., 55, (1990). 4. PG. Pries DE Oliveira, F. Lefebvre, M. Primet, J.G. Eon, and J .C. Volta, J. Cata], 130, 293, (1991). 5. PG. Pries DE Oliveira, J .G. Eon, and 1C. Volta, J. Catal, 137, 257, (1992). 6. H. P. Klug and L. E. Alexander,”X-ray Diffraction Procedures for Polycrystalline and Amorphous Materials", New York, N.Y., Wily-Interscience Publication, 1974. 7. A. Proctor, University of Pittsburgh. 8. C. Defosse, P. Canesson, P. G. Rouxlet, and B. Delmon, J. Catal., 51, 269, 1979. 9. F. P. J. M. Kerkhof and J. A. Moulijn, J. Phys. Chem, 83, 1612, 1979. 10. J. H. Scofield, J. Electron Spectrosc. Relat. Phenom, 8, 129, 1979. 11. D. R. Penn, J. Electron Spectrosc. Relat. Phenom, 9, 29, 1976. 12. D.V. Tarasova, V.A. Dzis'ko, I.P. Olen'kova, L.T. Tsikza, L.G. Karakchiev, and V.V. Molodtsova, Kin. i Kata, 14(2), 481, 1973. 13. Y. Nakagawa, T. Ono, H. Miyata, and Y. Kubokawa, J. Chem. Soc, Faraday Trans. 1, 79, 2929, 1983. 14. H. Eckert and LE. Wachs, J. Phys. Chem, 93, 6796, 1989. 15. G. Engelhardt and D. Michel, High-Resolution Solid-State WR of Silicates and Zeolites, Wiley, New York, 1987. 16. P. Massiani, B. Chauvin, F. Fajula, F. Figueras, and C. Gueguen, Appl. Catal., 42, 105, 1988. l” 17. 18. 19. 20. 21. 22. ' 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 66 PG. Pries de Oliveria, F.Lefebvre, M. Primet, J .G. Eon, and J .C. Volta, J. Chem. Soc, Chem. Commun, 1480, 1990. F. Roozeboom, T. Fransen, P. Mars, and PI. Gellings, Anorg. Allg. Chem, 449, 25, 1979. AA. Davydov, Zh. Fiz. Khim, 65,1803, 1991. H. Miyata, K. Fujii, T.Ono, Y.Kubokawa, T.Ohno, and F. Hatayama, J. Chem. Soc, Faraday Trans, 83(1), 675, 1987. Y. Kera and K. I-Iirata, J. Phys. Chem, 37(11), 3973, 1969. G. Bond and P. K6nig, J. Catal., 77, 309, 1982. QC. Bond, J.P. Zurita, and S. Flamerz, Appl. Catal., 27, 353, 1986. B. Blaauw, F. Leenhouts, F. van der Woude, and GA. Sawatsky, J. Phys. C., 8, 367, 1975. J. Mendialdua, Y. Barbaux, L. Gengembre, J .-P. Bonnelle, B. Grzybowska, and M. Gasior, Bull. Pol. Acad. Sci. Chem, 35, 213, 1987. B. Blaauw, F. Leenhouts, F. van der Woude, and GA. Sawatsky, J. Phys. C., 8, 367, 1975. K. Mori, Y. Yoda, N. Tamazoe ,and T. Seiyona, J. Phys. Chem, 84, 2564, 1984. W.M.H. Sachtler, G.J.H. Dorgels, V. Farenfort, and R.J.H. Voorhoeve, Proc. 4th Intern. Cong. Catal., 1, 355, 1970. RK. Grasselli, Appl. Catal., 15, 127, 1985. H.H. Voge and CR. Adams, ”Advances in Catalysis", Vol. 17, p. 151, Academic Press, New York, 1967. G.I. Straguazzi, K.B. Bischoff, TA. Koch, and G.C.A. Schuit, J. Catal., 104, 47,1987. J.D. Burrington, CT. Kartisek, and R.K. Grasselli, J. Catal., 87, 363, 1984. Chapter 4 Supported Antimony Oxide Catalysts 4.1. Introduction Antinomy oxide is often employed to promote the activity of catalysts used for selective oxidation [1-5] and for related reactions such as oxidative coupling [6]. Among the more common reactions are the oxidation and ammoxidation of propylene to acrolein and acrylonitrile, the conversion of isobutene to methacrolein, and the condensation of isobutene and formaldehyde to isopropene. Grasselli and co-workers [7] examined the relationship between structure and catalytic activity of uranium-antimonates for the synthesis of acroylnitrile fi'om propylene, ammonia, and air. They found that USb05 and USb3010 phases are catalytically active, but only USb3010 is selective. The selectivity has been attributed to the orthorhombic structure that has no adjacent U-O-U structures. In USb05, U-O-U-O-U structures present in the catalyst are centers for waste formation. In addition, work has been performed on iron antimonate catalysts which are used for the conversion of butene into butadiene [8-14]. Numerous hypotheses have been proposed to explain the reactivity of Fe-Sb-O systems based on surface geometry, electronic properties, and bulk structure [10,11,13]. Considerable work has also been performed on Sb oxides supported on rutile structures such as Sn02 [15-18] and Ti02 [19] for the selective oxidation of propylene to acrolein. It appears that in these systems, the catalytic performance for the oxidation of olefins is related to an enhancement of antimony at the surface of the rutile support. This enhancement is achieved by high temperature calcination which results in migration of the antimony through the bulk. 67 68 Lo et al. [20] reported the oxidative coupling of methane using a potassium- modified silica supported antimony oxide catalyst. This study found that of the Sb6013, KSbO3, and a-Sb204 phases most predominantly formed, only a-Sb204 was selective in methane oxidative coupling reactions and gave high selectivity to C2. Antimony oxide has also been used as a promoter for catalysts used in CO oxidation [21]. Ali-Zade et al. [22] found that the activity of Cu, Co, Cr, and Mn oxides supported on y-alumina is improved when Sb oxide is added. Antimony oxides have also been used in conjunction with oxides of Be, Mg, Zn, A1, Si and rare earth metals [23] and Sn-Mn-Pb oxides [24] for treatment of exhaust gases. Straguzzi et al. [8,10] examined C0 oxidation and selective oxidation of 1-butene to butadiene on iron, aluminum, cobalt, chromium, and rhodium antimonates. They found that these reactions have the same trend in activity, which suggests that the reactions occur on the same site and indicates that the antimony serves two functions: 1) to limit the size of the active oxygen ensemble and 2) to adsorb 02. In recent years, there has been an increased interest in preparation methods to produce highly dispersed metal oxides on supports such as Si02, A1203, and TiOz [25,26]. In general, these procedures are directed at producing monolayers with high thermal stability of the metal oxide coating. However, it has been found that the sorption [27,28] and catalytic properties [29-31] of the monolayer metal oxides are often different from those observed for the unsupported oxide phase. For example, Denofre et al. [28] found that Nb205/Si02 irreversibly adsorbs ascorbic acid forming a surface complex. Shirai et al. [29] reported that the activity obe205/8102 was 20 times greater than niobic acid bulk catalysts for esterification from ethanol and acetic acid. Recently, Benvenutti et al. [32] have reported the thermal stability and the acidic properties of Sb205 grafted on silica. Their results suggest that when antimony atoms are distant thermal sintering occurs formation of crystalline Sb205 is inhibited. The highly dispersed material contains Lewis acid sites due to unsaturated coordination of the Sb5+ 69 ion. The sites disappear upon thermal treatment at 500°C due to the extensive reticulation of Sb5’+ with the silica surface. In this chapter, the surface. structure of silica and alumina supported antimony oxide catalysts we will correlated with their activity for C0 oxidation. The focus will be on defining the nature of the Sb-support interaction and the influence of Sb dispersion and reactivity on the catalytic properties of the mixed oxides. In order to determine structure and reactivity of the species in the mixed oxide catalysts, X-ray diffraction (XRD) and X- ray photoelectron spectroscopy (XPS) will be used. The results obtained will be correlated with CO oxidation activity measurements in order to establish the nature of CO oxidation site(s) in the supported Sb oxide catalysts. 4.2. Experimental Cata]yst Preparation. The supported antimony oxide catalysts were prepared by pore volume impregnation of y-A1203 (Cyanamid; surface area: 204 mz/g; pore volume: 0.6 ml/g) and Si02 (Davison; surface area: 310 mZ/g; pore volume: 1.15 ml/g) which had been ground to 230 mesh. The supports were calcined in air at 500°C for 12 hours. The Si02 was allowed to sit in a desiccator for 12 days and A1203 for 7 days prior to impregnation. The catalysts were prepared by incipient wetness using a solution of antimony (III) n-butoxide (Strem, 99%) in dry n-hexane (Aldrich, 99.9%). The impregnated catalysts were dried at room temperature in a nitrogen atmosphere for 48 hours, dried in air at 120°C for 24 hours and then calcined in air at 500°C for 12 hours. Nominal loadings of the active element range from 0-40 wt.% as Sb205 (Sb/Al atomic ratio of 0 to 0.21 and Sb/Si atomic ratio of 0 to 0.25). Catalysts will be designated as Sb#A1 for the Sb/A1203 series and Sb#Si for the Sb/8102 series where # is the weight percent of antimony as Sb205. 70 Standard Materials. Sb205 was prepared by the addition of SbC15 to deionized water under vigorous stirring. The pH of the solution was then adjusted to a pH of 6-7. The precipitate formed was filtered and washed with deionized water, dried in air at 120°C for 12 hours and calcined at 500°C for 12 hours. Ale04 was prepared by coprecipitation of a 1:1 atomic ratio of AI(N03)3-9H20 (J .T. Baker, A.C.S. grade) dissolved in deionized water and SbCl5 (Aldrich, 99%). SbC15 was added to a vigorously stirred solution of A1(NO3)3 over a 30 minute period. NH4OH (J .T. Baker, A.C.S. grade) was added until the solution pH reached 6-7. The neutralized solution was allowed to stir for 1 hour. The precipitate formed was filtered and washed with deionized water; dried in air at 120°C for 12 hours, ground to produce a homogeneous mixture, and calcined at 750°C for 12 hours. XRD patterns of the standard compounds matched the appropriate ASTM powder dimaction file. BET Surface area. Surface area measurements were determined using a QuantaChrome Quantasorb Jr. Sorption System. Approximately 0.1 grams of the catalyst was outgased between 160°-165°C for 12 hours prior to adsorption measurements. The measurements were made using relative pressures of N2 to He of 0.05, 0.08, and 0.15 (surface area N2: 0.162 nmz) at 77 K. Data were processed using a Macintosh computer. X-ray Diflraction. XRD patterns were obtained using a Rigaku XRD dim'actometer which employs Cu Ka radiation (1.541838 A). The x-ray was Operated at 45 kilovolts and 100 milliamps. The patterns were scanned between 10°-75° (2 theta) for A1203 catalysts and 10°-45° (2 theta) Sb/Si02 catalysts. The scan rate was 0.1°/min. with DS and S5 = 0.5. Powdered samples were mounted on glass slides by pressing the powder in an indentation on one side of the slide. Semi-quantitative XRD has been performed by using physical mixtures of Sb205 and A1203 with 0.05 wt %, 0.1 wt %, 0.25 wt %, and 0.5 wt % Sb205. For Sb/Si02, physical mixtures of Sb205 and Si02 with 0.25 wt.%, 0.45 wt.%, and 0.7 wt.% Sb205 were prepared. The samples were ground in order to produce a homogenous mixture. The areas of the Sb205 <400> and the A1203 71 <400> or Si02 <101> peaks were measured and used to generate a calibration curve that can be used to determine the amount of crystalline Sb205 present in the Sb/A1203 or SblSiOz catalysts. Peak areas were obtained using Googly software [33] and a linear background was assumed over the peak base. X-ray Photoelectron Spectroscopy. X-ray photoelectron spectra (XPS) were obtained using a Perkin-Elmer spectrometer equipped with a Mg (1253.6 eV)/A1 (1486.6 eV) dual anode and a 10-360 hemispherical analyzer with an omnifocus small spot lens. Data were collected using a PC 137 interface board and a Zeos 386SX computer. Samples were mounted as powders on double-sided sticky tape or spray coated on glass slides or stainless steel plates. Spray coated samples were prepared by air brushing a 10% suspension of the catalyst (20% deionized water and 80% acetone) onto a glass slide or a stainless steel plate heated at 60°C. XPS binding energies of the Sb/A1203 catalysts were referenced to the A1 2p (74.5 eV) and SblSiOz catalysts to the Si 2p (103.5 eV) peak. The binding energies for standard compounds were referenced to the C 1s (284.6 eV) peak. XPS binding energies were measured with a precision of i 0.2 eV, or better. The data were processed using Googly software. In situ XPS experiments were performed on catalysts treated in a reaction chamber attached to the spectrometer and transferred directly to the analysis chamber without exposure to air. The antimony oxide catalysts were reacted with 100 cc/minute of 5% 02 and 95% He for 1 hour at 350°C, then reacted in 20 cc/minute of H2 at 450°C for 10 minutes and then for 10 minutes at 450°C with 20 cc/minute of 02. Quantitative ”S. A number of models have been proposed to relate active phase/support XPS peak intensity ratios to catalyst structure. It has been shown by Defosse et al. [34] that one can calculate the theoretical intensity ratio (If/15°) expected for a supported phase (p) atomically dispersed on a carrier (5). Kerkhof and Moulijn [35] extended the Defosse model and derived expressions based on model catalysts that consist of sheets of support with cubic particles of active phase deposited between the support 72 layers. . The photoelectron cross-sections (o) and mean escape depths (A) of the photoelectrons used in the calculations are taken from Scofield [36] and Penn [37], respectively. For species dispersed as monolayers, the relationship is given by the following equation: 1° -13 p =(p) D(ep)op01(l+e 2) (2) I; b D(es)052(1-e'32) S .; monolayer where 19° is the intensity of the electrons from element p in the supported phase, 15° is the intensity of electrons from element s in the carrier, GP and as are photoelectron cross sections of the respective elements, D(ep,s) are detector efficiencies, 6p,s are the kinetic energies of the electrons, and [31,2 is the thickness (t) per the mean escape depth of the photoelectron from the support (15) or the promoter (21,). This model predicts a linear increase in the supported phase (p) to the carrier (5) peak intensity ratio as the (p/s) atomic ratio is increased. CO Oxidation. C0 oxidation experiments were carried out in a differential type flow reactor at low conversions (<10%). Steady-state was reached between 1.5 and 2 hours for each catalyst series. The gas flow rates were held constant with Brooks 5850 and Porter 201 mass flow controllers. The reactor temperature was monitored with an Omega CN 1200 controller. The gas products for Sb/A1203 and 8102 were analyzed with a Varian 920 gas chromatograph equipped with a TCD. The chromatograph was interfaced to a Hewlett-Packard 3394A integrator. The column used for permanent gas separation was a 5 foot 60/80 mesh Carbosieve column. The catalysts were pretreated with 5% 02 (99.98%)/95% He (99.9%) at 350°C for 1 hour prior to reaction. The 73 reactorwas Operated at 450°C with a 15 cc/minute flow of 4.8% C0/9.8% 02/85.4% He (mixture from AGA with purity > 99.9%). 4.3. Results BET. Variation in the BET surface area of Sb/A1203 and Sb/SiOz catalysts as a function of antimony oxide content are shown in Table 4.1. For both series, the surface area decreases as the loading of the antimony increases. However, corrections made for the active phase loading indicate that the surface area of the support does not change. X-ray Diffraction (RD). XRD patterns measured for the Sb/A1203 catalysts showed lines characteristic of the alumina carrier (Figure 4.1). For Sb205 loadings S 15 wt.% only peaks of the alumina carrier were seen. For loadings 2 20 wt.% a new feature is observed between 20-40° (2 theta). The feature is centered at approximately 29° (2 theta) for all loadings. As the Sb205 loading increases, the intensity of this feature increases. At loadings of 35 and 40 wt.% Sb205, Sb205 <400> and Sb205 <111> peaks are observed. Figure 4.2 shows the variation of the Sb205 <400>/A1203 <400> intensity ratios measured for Sb205/A1203 physical mixtures. The value measured for the Sb40A1 catalyst is also given. The results indicate that the amount of Sb205 crystallite in the catalyst is less than 0.2 wt.% Sb205. XRD patterns measured for the Sb/Si02 catalysts show the broad peak centered at 22° (2 theta) characteristic of the silica carrier (Figure 4.3). For Sb205 loadings less than 10 wt.% only the peak from the silica support was observed. For Sb205 loadings > 10 wt.% a new feature between 20° and 40° (2 theta) is observed centered at approximately 29° (2 theta). As the loading of Sb205 increases, the intensity of this feature increases. At loadings of 25, 35, and 40 wt.% Sb205, Sb205 <400> and Sb205 <111> peaks are observed. Figure 4.4 shows the variation of the Sb205 <400>/S103 <101> intensity ratios measured for Sb205/Si02 physical mixtures. The results indicate that the amount of Table 4.1. BET surface area (mZ/g) of SbZOS/A1203 and SiOz. 74 Weight % Sb205 A1203 Si02 0 203.6 307.8 5 195.2 285.5 10 179.5 261.1 15 172.5 241.3 20 167.5 228.3 25 162.7 210.1 30 149.1 198.9 35 135.8 175.2 40 129.1 163.4 75 Sb205 crystallite in each catalyst is less than 0.5 wt.% Sb205. The presence of XRD peaks due to Sb205 observed for high Sb loadings can be attn'buted to errors in catalyst preparation. During the preparation of Sb rich catalysts (Sb205 > 20 wt.%) it is likely that very small amounts of the antimony alkoxide migrates out of the catalyst bed onto the walls of the crucible used to prepare the catalysts. During calcination the alkoxide is converted to Sb205 which subsequently falls into the catalyst bed. Unfortunately, the crystallites are not large enough to be separated by sieving. Thus, we believe that much of the crystalline Sb205 observed in the catalyst is due to this preparation error. Samples were prepared three times in an effort to eliminate this problem. Results are presented for the samples that gave the lowest Sb205 <400> XRD peak intensity. The XPS Sb 3d3,2 binding energies measured for the powder, spray coated and reacted Sb/A1203 and Sb/SiOz catalysts were 540.6 i 0.2 eV. The XPS Sb 3d3,2 binding energy measured for Sb205 was 540.5 eV. The XPS Sb 3d3/2 binding energy measured for Ale04 was 540.5 eV. Figure 4.5 shows the variation of the Sb 3d3,2/Al 2p intensity ratio as a firnction of the Sb/Al atomic ratio measured for the Sb/A1203 catalysts. For Sb205 loadings < 30 wt.% the intensity ratios increase linearly with Sb loading for powder, spray coated, and reacted samples. For loadings 2 35 wt.% the increase in Sb/Al intensity ratios with Sb loading is less pronounced. Variation of the Sb 3d3,2/Si 2p XPS intensity ratio as a function of the Sb/Si atomic ratio is shown for the Sb/S102 catalysts in Figure 4.6. There is a linear increase in Sb/Si intensity ratio as a fiinction of Sb loading for powder, spray coated, and reacted samples. CO Oxidation Activity. Table 4.2 shows the average percent conversion for Sb/A1203 and Sb/S102 catalyst series. For the Sb/A1203 catalysts, the percent CO conversion increases with Sb loading up to the SblSAl catalyst. The activity decreases as the Sb loading increases from 15 to 40 wt.% Sb205. For Sb loadings 2 25 wt.% Sb205, 76 1 1 1 1 1 10.0 20.0 33.0 40.0 50.0 DEGREE (2 THETA) Figure 4.1. XRD of Sb205/A1203 a) 51140111, 11) 51135111, c) 31130111, d) 31125.11, e) Sb20Al, f) SblSAl, g) SblOAl, h) SbSAl, and i) SbOAl catalysts. «sz0, <400>)/1(A1203 <400>) 77 0.03 T I ' Physical Mixtures / / "' ‘ - Best Fit / / 0 40 wt.% // Sb/A1203 / / / 0.02 ‘1- / / / / / / / / / / / / / 0.01 "i' / d / / / / / I /I / / / 0 9’ 1 1 1 1 a 0 0.1 0.2 0.3 0.4 0.5 Weight Percent Sb205 Figure 4.2. Semi-quantitative XRD physical mixtures of Sb205 and A1203. 78 59205 — — l l 1 0.0 20.0 30.0 40.0 DEGREE (2 THETA) Figure 4.3. XRD of Sb205/Si02 a) SMOSi, b) Sb3SSi, c) SbBOSi, d) SbZSSi, e) SbZOSi, 1) SblSSi, g) SblOSi, h) SbSSi, and i) SbOSi catalysts. 79 0.8 0.006 - ' Physical Mixtures . / — - - Best Fit / . / l 40 wt.% Sir/8102 / ' / ° 35 wt.% Sb/Sioz / A 0.004 -- // /\ § / / V r 5' e; ) / t: / "" / § / '1 / 2- / a / "' -_ / 0.002 / / / / / / / a / / a / / 0 1! 1 1 1 1 0 0.2 0.4 0.6 Weight Percent 511205 Figure 4.4. Semi-quantitative XRD physical mixtures of Sb205 and S102. 80 6 T- 5 a ' powder . I a 0 Hydrogen . z A Spray Coated A 4 we ' Oxygen a :5? ram/957.1% 3 57:1 3 - 1‘3" M .0 9’3 2 d l _1 0 i i l i 41 0 0.05 0.1 0.15 0.2 0.25 Sb/Al Atomic Ratio Figure 4.5. XPS intensity ratio of Sb 3dm/A1 2p as a function of Sb/Al atomic ratio. 81 4 q.- ' Powder 3 ° Hydrogen 1‘ Spray Coated 0 ' Oxygen . ‘ 5% 02/95% He . ’3. N 123, 1'5" (‘1 .D 9 1 .m- 0 1 i 1 1 i 0 0.05 0.1 0.15 0.2 0.25 Sb/Si Atomic Ratio Figure 4.6. XPS intensity ratio of Sb 3d3,2/Si 2p as a function of Sb/Si atomic ratio. Table 4.2. CO conversion for Sb205/A1203 and SbZOS/Sioz at 450°C and 15 cc/minute 82 flow rate. Weight % Percent Conversion Percent Conversion Sb205 for Sb/AIZOf for Sb/SiOza 0 2.7 0.4 5 5.1 1.1 10 7.0 2.0 15 7.3 1.3 20 4.0 1.9 25 2.2 0.9 30 1.8 0.9 35 1.7 0.8 40 1.0 0.8 AleO4 9.6 — 3 - Error bars oft 10%. 83 the activity is lower than the value obtained for the support. For the SblSiOz catalysts, the percent CO conversion increases with Sb loading up to the SbZOSi catalyst. The activity decreases as the Sb loading increases from 20 to 40 wt.% Sb205. As shown from the table the activity of antimony oxide supported on alumina is greater than that of the silica. However, the two series show similar trends in the region of greatest activity (between 10 and 20 wt.%. Sb205). CO oxidation was performed using the Ale04 catalyst. The average percent conversion at steady state is 9.6%. 4.4. Discussion Structure of Calcined Antimony Oxide/A lumina Catalysts. For Sb/A1203 catalysts with Sb205 loadings S 30 wt.%, absence of peaks that can be attributed to any antimony oxide phases indicates that no crystalline antimony oxide phases are formed in these catalysts. However, the observation of the Sb205 <400> peak for catalysts with 2 35 wt.% Sb205 indicates crystalline Sb205 is present. Semi-quantitative XRD analysis suggests that these catalysts contain less than 0.2% crystalline Sb205. In addition, the broad peak centered at approximately 29° (2 theta) is in the same region of the crystalline Sb205 <400>. This may be attributed to an amorphous form of antimony oxide. The XPS Sb 3d3,2 binding energies measured for Sb/A1203 catalyst series indicates that the antimony is in the *5 oxidation state and does not vary as a fiinction of antimony loading. The XPS Sb/Al intensity ratios measured for the alumina catalysts with S 30 wt.% Sb205 show a linear increase in intensity ratio as the Sb/Al atomic ratio increases. This suggests that the antimony phase is well dispersed over the alumina carrier up to 30 wt.%. For higher Sb loadings, the deviation from linearity indicates that the antimony oxide is not as highly dispersed over the alumina support. This is consistent with XRD results. 84 Further, the absence of any significant change in Sb/Al intensity ratio following reactions suggests that the antimony oxide phase is not mobile under reaction conditions. Structure of Calcined Antimony Oxide/Silica Catalysts. For Sb/Si02 catalysts with S 30 wt.% Sb205, the absence of peaks that can be attributed to antimony oxide phases indicate that no crystalline forms of antimony oxide are present. However, the observation of Sb205 <400> peak for catalysts with 25 wt.% and 2 35 wt.% Sb205 indicates that crystalline Sb205 is present. In addition, the broad peak centered at approximately 29° (2 theta) is in the same region of the crystalline Sb205 <400>. This result indicates that an amorphous form of antimony oxide is produced on the support. The XPS Sb 3d3,2 binding energies measured for Sb/Si02 catalyst series indicates that the antimony is in the +5 oxidation state and is independent of antimony loading. Binding energies for the series are similar to those reported previously [3 8,39]. For Sb/SiOz catalysts, the linear increase in Sb/Si XPS intensity ratio with increasing Sb loading suggests that antimony oxide is highly dispersed. This is consistent with XRD results which showed a broad feature that we have attributed to dispersed antimony oxide. In addition, no change is observed under reaction conditions. This indicates that the reactions do not affect the dispersion of the antimony. Structure ofReacted Antimony Oxide/A lamina and Silica Catalysts. The Sb 3d3,2 XPS binding energy for Sb205 is 5397-5405 eV [40]. The Sb 3d3,2 XPS binding energy measured for Sb203 is 539.5 eV. Thus, since the binding energy difference between Sb3+ and Sb5+ is approximately 0.7 eV, XPS cannot readily distinguish between these two species [41-44]. Since the difference in binding energies is small, a broad peak results when both peaks are present. Thus, curve fitting of the Sb XPS peak has some ambiguity. As a result, it is not possible to determine if changes are occurring as a function of reaction condition. Comparison of Antimony Oxide on Alumina and Silica. The concentration of the surface OH groups/surface area of the alumina support is 1.7x10'3 mmol/m2 and for the 85 silica support the value is 9.1x10'4 mmol/m2 [45]. Thus, the ratio of hydroxyls of the alumina support to the hydroxyls of the silica is approximately 2:1. If it is assumed that one Sb-alkoxide reacts with one hydroxyl on the support, monolayer capacity of the alumina is expected to be approximately 35 wt.% and of silica approximately 40 wt.% (taking the surface area of the support into effect). XPS and XRD data for the alumina and silica supported catalysts are in agreement with this prediction. Influence of Catalyst Structure on C O Oxidation Activity. CO oxidation reactions have shown that the most active catalysts contain 10 to 20 wt.% loadings of Sb205. As the loading increases (> 25 wt.%) in both series, the activity of the catalyst decreased. In the case of 30-40 wt.% Sb/A1203 the activity is lower than that of the support. As for SiOz, the activity decreases to approximately 0.8% which is higher than that of the pure support (0.4%). The drastic change in the support from that of the 40 wt.% catalyst may be due to the activity of Sb205 itself. This is shown by the activity of the higher loadings of the active phase. At the loadings of greatest activity, it is possible that a spillover, as seen with Pt catalysts, is occurring between the Sb and the support. The trend of observing a maximum then as the loading increases lends support to this explanation. Once the surface approaches saturation with the Sb phase, the interaction sites available between the support and the Sb decrease, which decreases the activity of the catalyst itself. Davydov [46] used CO to probe VzOs/A1203 catalysts and observed spillover of carbon-oxygen species from the active vanadium pentoxide component on the alumina carrier. CO oxidation on the aluminum antimonate support showed an increase in conversion over the supported Sb205/A1203. This suggests that the interaction between Sb and A1 is important in this reaction. Straguzzi et a1. [10] studied CO oxidation and found that at 460°C the conversions were less than 10% and showed activity of 1.0 mmol/mzh. Straguzzi suggests that because of the difficulty in reducing Al“, the activity of the aluminum antimonate is attributed to Sb cations. 86 4.5. Conclusions The use of surface and bulk spectroscopic techniques and CO oxidation activity data for Sb205/A1203 and Sb205/S102 catalysts leads to the following conclusions: 1. XRD and XPS data indicate that Sb is well dispersed over the A1203 and SiOz carriers. 2. Semi-quantitative XRD for A1203 suggests that less than 0.2 wt.% crystallites of Sb205 are present and less than 0.5 wt.% crystallites of Sb205 are present in the Si02 supported catalysts. We tentatively ascribe this to errors in catalyst preparation. 3. The binding energy of Sb indicates that the antimony is in the +5 oxidation state in both catalysts and is independent of Sb loading and reaction conditions. 4. CO oxidation suggests that the highest activity occurs with active phase loadings of 10-20 wt.% Sb205 on both A1203 and SiOz. As loading increases above this level, the activity decreases. This is tentatively ascribed to spillover. References 1. RK. Grasselli and JD. Burrington, Adv. Catal., 30, 133, 1981. 2. R.K. Grasselli, React. Kinet. Catal. Lett., 35, 327, 1981. 3. G. Centi and F. Trifirc‘), Catal. Rev. Sci. Eng, 28, 165, 1986. 4. 1.0 Vedrine, G. Coudurier, M. Forissier, and 1.0 Volta, Catal. Today, 1, 261, 1987. 5. F.J. Berry, Adv. Catal., 30, 97, 1981. 6. SE. Golunski and D. Jackson, App]. Catal., 48, 123, 1989. 7. RK. Grasselli and J.L. Callahan, J. Catal., 25, 273, 1972. 8. GI Straguzzi, K.B. Bischoff, TA. Koch, and 6.0 Schuit, App]. Catal., 25, 257, 1986. 9. G1. Straguzzi, K.B. Bischoff, T.A. Koch, and QC. Schuit, J. Catal., 103, 357, 1987. 10. G1. Straguazzi, K.B. Bischoff, T.A. Koch, and G.C.A. Schuit, J. Catal., 104, 47, 1987. 11. F. Sala and F. Trifirc‘), J. Catal., 34, 68,1974. 12. V. Fattore, Z.A. Furhman, G. Manara, and B. Notari, J. Catal., 37, 223, 1975. LEE-.ufl'. Airh- ' ..‘ fi.‘ .1“ -’ 13. 14 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 87 M. Burriesci, F. Garbassi, M. Petrisa, and G. Petrini, J. Chem. Soc, Faraday Trans 1, 78, 817, 1982. I. Aso, S. Furukawa, N. Yamazoe, and T. Seiyama, J. Catal., 64, 29, 1980. Y. Boudeville, F. Figueras, M. Foissier, J-L. Portefaix, and J.C. Vedrine, J. Catal., 58, 52, 1979. J.L. Barclay, 1.11. Bethell, J.°B. Bream, D.J. Hadley, RH. Jenkins, no. Stewart, and B. Wood, Brit. Patent 864,666. F.J. Berry and D]. Smith, J. Catal., 88, 107, 1984. J.C. Volta, B. Benaichouba, I. Mutin, and J.C. Vedrine, App]. Cata]., 8, 237. 1983. GA. Zenkovets, D.V. Tarasova, T.V. Andrushkevich, G.I. Aleshina, T.A. Nikoro, and RG. Ravilov, Kinet. Catal., 20, 380, 1979. M.-Y. Lo, S.K. Agarwal, and G. Marcelin, J. Catal., 112, 168, 1988. IL Mikhailova, I.S. Sazonova, N.P. Keier, and TV. Belodtseva, Kinet. Katal, 9(3), 565, 1968. FM. Ali-Zade, NM. Mardanova, PM. Sorgin, RM. Talshinskii, AA. Medzhidov, and RC. Rizaev, U.S.S.R. Patent 49,052,169, 1991. S. Blumrich, R. Brand, B. Engler, W. Honnen, and E. Koberstein, European Patatent 410,440, 1991. Y. Ohno and K. Morishita, Japanese Patent 49,052,169, 1974. G. Busa, Langmuir, 2, 577, 1986. GO Bond, S. Flamerz, and R. Shukri, Faraday Discuss. Chem. Soc, 87, 65, 1989. Y. Gushikem, C.R.M. Peixoto and LT. Kubota, New Developmemts in Ion Exchange, ICIE, 607, 1991. S. Denofie, Y. Gushikem, and C.U. Davanzo, Eur. J. Solid State Inorg. Chem, 28, 1295, 1991. M. Shirai, K. Asakura, and Y. Iwasawa, J. Phys. Chem, 95, 499, 1991. GO. Bond and K. Briickman, Faraday Discuss. Chem. Soc, 72, 235,198]. E.C. Alyea, K.F. Brown, and K.J. Fisher, J. Mol. Catal., 63, L11, 1990. E.V. Benvenutti, Y. Gushikem, C.U. Dananzo, 8.0 de Castro, and IL. Torriani, J. Chem. Soc. Faraday Trans, 88, 3193, 1992. A. Proctor, University of Pittsburgh. C. Defosse, P. Canesson, P. G. Rouxlet, and B. Delmon, J. Cata]., 51, 269, 1979. F. P. J. M. Kerkhof and J. A. Moulijn, J. Phys. Chem, 83, 1612, 1979. J. H. Scofield, J. Electron Spectrosc. Relat. Phenom, 8, 129, 1979. D. R. Penn, .1. Electron Spectrosc. Relat. Phenom, 9, 29, 1976. E.V. Benvenutti, Y. Gushikem, C.U. Dananzo, 8.0 de Castro, and IL. Torriani, J. Chem. Soc. Faraday Trans, 88, 3193, 1992. A. Li-Dun, J. Zhi-Cheng, and Y. Yuan-Gen, in ”Catalyst Deactivation 1987”, ed. B. Delmon and GP. Froment, 159, 1987. RA. Lemberanskii, F.M. Poladov, I.B. Annenkova, and AP. Shepelin, React. Kinet. Catal. Lett., 32(1), 103, 1986. WE. Morgan, W.J. Stec, and JR. Van Wazer, Inorg. Chem, 12, 953, 1973. 42. 43. 45. 46. 88 T. Birchall, J .A. Connor, and 1H. Hiller, J. Chem. Soc, Dalton Trans, 2003, 1975. AF. Orchard and G. Thornton, J. Chem. Soc.,Dalton Trans, 123 8, 1977. F. Garbassi, Surf Interface Anal, 2, 165, 1980. J. Kijenski, A. Baiker, M. Glinski, P. Dollenmeier, and A. Wokaun, J. Catal., 101, 1, 1986. ' AA Davydov, Zh. Fiz. Khim., 65,1803, 1991. Chapter 5 . Future Work 5.1. Characterization Characterization of Mixed Metal Oxide Catalysts. In addition to the use of XPS to determine the nature and distribution of the active sites in complex mixed oxides, ion scattering spectroscopy (188) can also provide information on the active sites(s) following various reactions. In the case of vanadium oxide and antimony oxide monolayer catalysts, reactions of the surface phase with the reducing molecules could lead to surface migration and subsequent multilayer formation. Monolayer-to-multilayer transitions can be difficult to follow using XPS due to the relatively large sampling depth of the technique. Thus, 188 can be useful to follow subtle changes in the geometry of the vanadium oxide and antimony oxide caused by surface reactions. Along with XPS, Raman spectroscopy can also provide information on the active site(s) under various reaction conditions. Raman spectroscopy can also be used to monitor differences in the chemical reactivity of surface species present. In addition to using EPR to identify V4+ in the reacted vanadium oxide on aluminum antimonate catalysts, it would also be useful to characterize the active site(s) by performing in situ reaction studies on both the V/AleO4 and Sb/SiOz and Sb/A1203 systems. An example of this use of EPR can be found by the double bond isomerization of butene over y-alumina [1]. Both the isomerization reactions and the butene-deuterium exchange reaction have been extensively studied with a view to understanding the nature of the active sites on alumina. NMR can also be used in much the same way as EPR to study active site(s), by using 13C NMR studies of adsorbed molecules. For example, Stejskal et al. [2] studied 89 sum.- .' 90 physically adsorbed C02 on zeolites. In addition, 13C NMR has been used to study chemisorbed molecules [3]. NMR is unique in its ability to view molecules on the time scale of a surface reaction. Since the motion of molecules and ions on the surface is an integral part of heterogeneous catalysis, these studies have aided in the development of a dynamic picture of adsorption and proton mobility. In situ IR will also provide information on the adsorption of probe molecules or reaction gases on the active site(s) of the catalysts. The IR cell shown in Figure 5.1 will allow the catalyst to be studied under controlled atmosphere and temperature. The gases can be delivered to the in situ IR cell by the system shown in Figure 5.2. 5.2. Vanadium Oxide/Aluminum Antimonate Catalysts Propane Oxidation and Ammoxidation. As mentioned earlier, there has been an increased interest in selective oxidation and ammoxidation of inexpensive feedstocks, such as propane and methane. The next area of investigation would be the study of propane oxidation and ammoxidation over the V/AleO4 catalyst. In order to learn more about the mechanism(s) for propane oxidation and ammoxidation, careful kinetics studies on the catalyst must be performed. In order to obtain this information, it will be necessary to study the effects of propane, oxygen and ammonia partial pressures on the rates of propylene, acrolein, acetaldehyde, acrylonitrile, acetonitrile and carbon oxides formation. Centi et a1. [4] reported that specific oxygen and ammonia concentrations exist which enhance the rates of propylene, acrylonitrile and acetonitrile formation over a V-Sb- W (1:5: 1) A1203 (70 wt.%) catalysts. The kinetic experiments proposed will determine if there is a similar ”optimum" ratio that exists for V/Ale04. If such a ratio exists, it will then be useful to perform IR and surface spectroscopic studies to determine the surface phases responsible for the enhanced catalytic properties. Influence of Excess SbOx on AleO,, supported V cata]ysts. It is well known that 91 Heater ports /\ Vent 8 Thermocouple Leads 1 Clip Thermocouple S'd Vi Top View I e ew Center Piece Figure 5.1. In situ IR cell. 92 Z-EOUTE W-MASSFLOWCON'I'ROUER I'm-FOURIERTRANSFORMW tic-mm Figure 5.2. Gas handling system to in situ IR cell. 93 crystalline and monolayer phases have different effects on activity. Recently, Beny and Brett [5-7] reported that biphasic interactions between VSbO4 and a-Sb204 are important for the selective oxidation of propylene. However, the influence of excess SbOx phase on the structure and activity of the catalysts is not well understood. The effects of excess SbOx phase on the chemical state and dispersion of the supported V oxide and subsequently on the activity and selectivity of the mixed oxide catalysts for propylene and propane oxidation and ammoxidation should be investigated. From the different phases produced, it will be possible to determine the effects of crystalline SbOx and highly dispersed SbOx phases on the properties of V/AleO4 catalyst. Future experiments will involve preparation and characterization of V/SbOx/AleO4 catalysts. 5.3. Antimony Oxide! Supported Catalysts Sb/Supported Catalysts. There has been considerable research effort in methane coupling and total oxidation of methane. As mentioned previously, Lo et al. [8] reported a-Sb204 was a selective methane coupling catalyst. Dang and Ding [9] found that Sb/Si02 catalysts that contain an Sb-SiOz surface compound (Sb loadings < 5 %) are active for condensation reactions. However, the condensation activity decreases significantly on the formation of a-Sb204 (Sb loadings > 5 %). Thus, the combination of methane coupling and condensation of isobutene and formaldehyde to is0prene may be useful reactions to probe the structure of supported Sb catalysts. It should be noted that regardless of the ability of the test reactions to probe the structure of the mixed oxides, valuable information about the nature of active site(s) in Sb supported catalysts will be obtained. By varying the partial pressures of methane and oxygen it will be possible to study the effects of the partial pressures on the rate of carbon oxides produced. In addition, by having a higher level of methane it may be possible to control the selectivity of methane 94 coupling over total oxidation of methane. In order to increase the selectivity of total oxidation it would be necessary to have a higher partial pressure of oxygen over methane. However, these kinetic experiments will determine if Optimum ratios exist for each reaction (i.e. methane coupling and total oxidation). If such a ratio exists, IR and surface spectroscopic studies will be used to determine the surface phases responsible for the enhanced catalytic properties. In addition, the effect of calcination temperature on the catalyst structure and activity should be examined. Sb205 decomposes to form B—Sb204 at 730°C and Ot-Sb204 at 950°C. By performing solid-state reactions, it will be possible to determine Optimum conditions for methane coupling and/or total oxidation of methane as a function of antimony oxides formed. 5.4. Activation Energy Activation Energy. It would also be useful to determine the activation energy for oxidation reactions of the V on aluminum antimonate and the Sb on alumina and silica, to improve our understanding of these reactions. Jonson et al. [10] have performed activation energy studies Of V on titania. From this, it was found that at least two different active sites are present with a active phase loading of 005% V; one for the support, which is the least active phase and one for the vanadium species, which is the more active phase. Also, at 2% V and above, the active part of the support surface is covered and changes in the Arrhenius parameters can be attributed to the presence in these catalysts of an additional V species at higher loadings. By performing such activity measures and correlating the data with that obtained from spectroscopic techniques, it will be possible to better define the nature Of metal oxide-support interaction and the influence of the metal oxide dispersion and reactivity on the catalytic properties of the mixed oxides. 95 References 1. J.H. Lunsford, L.W. Zingery, and MP. Rosynek, J. Catal., 38, 179, 1975. 2. E0. Stejskal, J. Schaefer, J.M.S. Henis, and MK. Tripodi, J. Chem. Phys, 61, 2351, 1974. 3. JJ. Chang, A. Pines, J.J. Fripiat, and HA. Resing, Surface Sci, 47, 661, 1975. 4. G. Centi, R.K. Grasselli, E. Patane, and F. Trifiro,"New Developments in Selective Oxidation,” Stud Surf Sci. Catal., 55, 515, 1990. 5. F.J. Berry, ME. Brett, and W.R. Patterson, J. Chem. Soc. Dalton, 1, 9, 1983. 6. F.J. Berry, M.E. Brett, and W.R. Patterson, J. Chem. Soc. Dalton, 1, 13, 1983. 7. F.J. Berry and ME. Brett, J. Catal, 88, 232, 1984. 8. M.-Y. Lo, S.K. Agarwal, and G. Marcelin, J. Cata]., 112, 168, 1988. 9. Z. Dang and S. Ding, Fenzi Cuihua, 1(3), 146, 1987. 10. B. Jonson, B. Rebenstrol, R. Larsson, and S.L.T. Andersson, J. Chem. Soc, Faraday Trans. 1, 84(10), 3547, 1988.