You are on page 1of 12
APPLIED ‘Applied Catalysis A: General 202 2000) 121-181 ‘wu leeviereomfoatlap The reaction performances and characterization of Fischer—Tropsch synthesis Co/SiO2 catalysts prepared from mixed cobalt salts Shouli Sun, Noritatsu Tsubaki*, Kaoru Fujimoto Deparment of Applied Chemis, School of Engineering, The Univers of Toyo, Honge 7-51, Bunkyo-tu, Toyo 113-8656, Japan ‘Received 14 October 199; ecelved in revised form 20 January 2000; cepted 21 January 2000 Abstract Catalysts prepared by mixed impregnation of cobalt) nitrate and cobal( IH) acetate displayed higher activity than the catalysts prepared from either mono-precursor at mild reaction conditions (1 MPa total pressure, Ha/CO=2, T=513K) of, Fischer-Tropsch synthesis (PTS). X-ray diffraction (XRD) indicated that highly dispersed eoball metal provided the main active sites onthe catalyst prepared by mixed impregnation method. Through the mixed impregnation, different cobalt species ‘were formed and their reduction performances were detected by the temperature-programmed reduction (TPR) and thermal zgzavimetic analysis. Transmission electronic microscopy (TEM) and FT-IR spectoscopy of adsorbed CO as probe molecule revealed that the presence of different sites associated with cobalt after the reduction of the catalysts with hydrogen at 673 K. Jtwas assumed that the metal readily reduced from cobalt nitrate promoted the reduction of Co* to metallic state in cobalt acetate by Hg spillover mechanism during catalyst reduction process. The reduced cobalt from cobalt acetate was highly ‘ispersed and remarkably enhanced the catalytic activity. © 2000 Elsevier Science B.V, All ights reserved. Keywords ColSiOg catalysts; Fischer-Tropsch symibess; Preparation method: Cyt preeusor 1. Introduction Cobalt has been one of the most widely used met- als for Fischer-Tropsch catalysts since the 1930s, specially when high-chain-growth probability and low branching probability are required [1-3], which fre very important for the production of liquid synthetic fuels. The FTS catalytic activity can be eased at high temperature, but selectivities of unfavorable products such as CH or CO are also enhanced sharply. So catalysts showing high a tivity at low temperature are needed. The catalytic Corresponding author. Tel: +813-5841-8579; fax: +81-3-5841737 Exmait adress: fan@appchem.w-ioky02¢jp (N. Tsuba’) activity of FTS is closely related to cobalt dis- persion values at typical FTS operation conditions 3]. The formation of highly dispersed catalysts re- uires strong interaction of support and the cobalt precursor. But such interactions tend to interfere with the reduction of such precursors 10 cobalt metal, Here, we report that using different cobalt salts as precursor has a great effect on the cobalt catalyst activity and selectivity. Highly active cobalt cata- Iysts prepared from the mixture of cobalt nitrate and cobalt acetate and their performances in the hydro~ genation of carbon monoxide in slurry phase will be introduced. XRD, TPR, TEM, FT-IR and chemisorp- tion techniques were used to characterize the catalysts. (826-860x00§ ~ see front maner ©2000 Elsevier Science BV. Allright resered, Pi: $0926-860x(00)00435-5 mm 2. Experimental 2.1. Catalyst preparation and its nomenclature ‘The nomenclature for catalysts inthis study consists of two parts: [xX/y¥] or [+X+y¥]. The first part is 2X, )¥:X or ¥ refers tothe type of used cobalt metal salt: ‘N’ indicates cobalt) nitrate and ‘A’ indicates cobalt(Dacetate; x refers to the weight percent of X and y refers to the weight percent of Y. The second part gives the order of impregnation of metal salts: [XY] means that X was impregnated into the support previous to Y, and [X+Y] means that X and Y were coimpregnated. ColSiO2 catalysts having different loading amounts of cobalt and N/A ratios were prepared (Table 1). ‘The supported cobalt catalysts were prepared by incipient-wetness impregnating cobalt{}) salts on a commercially available silica gel (ID gel, Fuji Da son, specific surface area: 270m? g-, pore volume 1.22cm3 g and average pore diameter 8.7 nm), dry- ing in air at 393K for 12h. Then, the catalyst was calcined in air by raising its temperature to 723K with a ramping rate of 3Kmin=! and holding that temperature for 2h [4,5]. After calcination, catalysts were activated in flowing hydrogen at 673K for 10h, the temperature ramping rate was 3K min“. The cat- alysts were passivated by 1% oxygen in nitrogen: For the sequentially impregnated catalysts, after the fist ‘impregnation step, the catalyst precursors were dried in an oven for 12h at 393K. Then, each was calcined in air at 723K for 2h and used for the second salt '. Sun etal. /Applied Catalysis A: General 202 (2000) 121-131 22. Catalyst characterization 2.2.1. Temperature-programmed reduction (TPR) A quartz microreactor was used for performing TPR experiments. The gas stream, & mixture of 5% Hh diluted by argon as reducing gas, was fed with a ‘mass flow controller. After the reactor, the effluent 5 was led via a 3 A molecular sieve trap to remove the produced water. The effluent gas was determined by gas chromatography with a thermal conductivity detector (TCD), where argon was used as a reference, CCalcined catalyst (0.2 g) was initially flushed with ar- ‘gon at 323K for 10min. Then, the reducing ges was switched on at SOml min“, and the temperature was raised at a rate of 7K min! until 1073K. 2.22. X-ray diffraction ‘An X-ray diffractometer (XRD, Rigaku) was used ‘with monochromatized Cu Ka radiation and a scan rate of 2.0° min~!, The catalyst samples for XRD measurement were at oxide state. The average Co304 particle sizes were calculated using the half-width at half-height of the most intense peak of the dffrac- tion pattern and the Scherrer equation. The diameter of a given CosOy particle could be used to calculate the diameter of metallic Co erystallite by the formula below [6] ay(Co°) ).75d(CosO4) a 2.2.3. Chemisorption analysis ‘The active surface metal of the catalyst was mea- ‘impregnation, repeating the procedure, ‘sured by chemisorption in a conventional volumetric ‘nile 1 Preparation duis for CoSiO2 cays, Coast wee wi Amie rao Step I Sep ee ee aod Sotaton Meiod Sotaion iw iO Tocipicnt weress Aqueous CotN) A 3s Tncpient wetness Aqueous Co(N) Incipient wetness Aqueous CoCA) BSA SSD Incipient wetness Aqueous CotN) ‘Incipient wemess__Aquevus CX(A) MBA T3823 Incipient wetness Aqueous Co(N) Incipient wemess Aqueous COA) SAN SSD Incipient wemess Aqbeous Co(A) Incipient wemess_Aguenas CotN) SN 5S Tncipien wemess Aqueous Co+A) m0 oo Incipient wemness Aqueous Co(A) "Following this wep the catyst was did aan ove for 12h a0 373K and then caleoed for 7h t TB 5. Sun et el. /Appled Catalysis A General 202 (2000) 121-131 3 adsorption system, as described in detail elsewhere (7). The chemisorbed uptake (at 373 K) was measured using the double-isotherms method as described by Bartholomew et al. [8,9]. Bartholomew etal. pointed ‘out that Ha chemisorption at 373K was most reliable for silica-supported cobalt catalysts. Dispersion per- centage (D%) was calculated according to the equa- tion D% = ak 2) WF @ Where X is the total Hy uptake in micromoles per gram of catalyst, W the weight percentage of cobalt, and f the fraction of cobalt reduced tothe metal determined from O- titration. It should be noted here tht, with the introduction of reduction degree f, the unreduced cobalt was not included in the calculation of D%. ‘Average crystallite diameters in nanometers were calculated from D% assuming spherical metal crys- lallites of uniform diameter d with a site density of 14.64toms/nm?. Thus, 96 = am ® here D® isthe percentage dispersion. The Os titration was carried out at 673K accord- to the method of Bartholomew and Farrauto (10), where the reduction of reduced cobalt with oxygen ‘was considered to form Co304 (1). 224, Thermal gravimetric analysis (TGA) ‘TG measurements were conducted using Shinku-Riko MTS 9000 instrument. The catalyst pre- ‘cursors after calcination were placed in the furnace and heated in a stream of Hy to the desired tempera ture at a heating rate of 18°Cmin=? and then held at the final temperature for 10 min, 22.5. FT-IR measurements ‘The FT-IR spectra were recorded using a Nico- Jet Magna 550 spectrophotometer. For each sample, 30-4Omg was pressed into a disk with a diameter cof 10mm. The disk was placed in an infrared cell ‘equipped with CaF windows. CO was adsorbed on the catalysts at room temperature (Pco=200 Tor) for 10min. Then, the cell was evacuated for 5 min (Pco: 10°8Tort). The detailed procedure of the measure- ‘ment was given elsewhere (12). 2.26. Transmission electronic microscopy (TEM) ‘A catalyst pellet was embedded into resin and cut bby microtome method (Leica Ultracut UCT). The ob- servation was implemented on JEOL TEM-2010, 2.3. FTS Reaction Catalysts were tested in a semi-batch slurry-phase reactor with a volume of 80m. 20ml of n-CigHse was used as the liquid medium. A cooling trap was set between the reactor exit and the pressure regula- tor to collect the water and liquid hydrocarbons. The standard reaction conditions were P(total)=1.0MPa, COM,=1/2, WIF(CO+H2)=5 ghmol!, T=513K. ‘Argon was employed as an internal standard with ‘concentration of 3% in the feed gas. Gaseous products were analyzed with on-line gas chromatographs. CO and COz were analyzed by us- ‘ng an active charcoal column equipped with a ther- ‘mal conductivity detector (TCD). Light hydrocarbons were analyzed using a Porapak-Q column with aflame ionization detector (FID). Hydrocarbons dissolved in the solvent and cooled in the trap were analyzed using a Silicone SE-30 column with FID. 3. Results 3.1. Effect of metal loading amount on Co/SiOn catalysis Table 2 shows the reaction performances in the slurry-phase on the supported cobalt catalysts different loading amounts. The catalyst precursor was ‘cobalt nitrate only. As cobalt loading increased from 2 to 40 wt.%, CO conversion increased from 1.38 to 56.84%. But the selectivity of methane did not show ‘8 meaningful change. ‘tis demonstrated in Fig. 1 that the CO conversion increased linearly with increasing the amount of sur- face metal. This result indicates that FTS on cobalt catalyst was a structure-insensitive reaction. This ob- servation well coincided with what reported by Iglsi et al [13]. The chain growth probability slightly in- creased with the increasing size of supported metallic crystallite [3,4] Iti clear that in Fig. 1, the line did not pass through the origin. It seems thatthe metallic 6 ‘Tate 2 '. Sun et aL. /Appied Catalysis A: General 202 (2000) 121-131 Sury-phase reaction performances of Co/SiO; eataysts with diferent Co loadings* ‘Cobalt coment (wa) CO conversion (6) Cis selectvy ‘CO, selec 7 2 138 338 08 083 5 316 40 on 086 10 2051 353 03s 0s 2» 33 430 088 090 25 3435 316 099 089 35 on 42 134 090 0 5684 603 235 ost crystalline wes very small in the case of very low metal surface concentration. This kind of small crystalline ‘was too weak to decompose the adsorbed CO, as CO bbond was readily cleaved by large metalic erystalline Mm. 3.2, Catalytic activity of Co-based catalysts Prepared from mixed cobalt salts Table 3 shows the catalytic activity and selectivity of the catalysts with different N/A ratio for CO hy- drogenation. For the catalysts 10A, 3N/7A, SN/SA, ‘TNIBA and 10N, whose total loadings were 10 wt.%, the reduction degree and the particle size determined from Ha adsorption increased with increasing N/A ratio, a8 shown in Fig, 2. But the CO conversion has the largest value at N/A=1. This phenomenon oo 100 «180 ‘Amount sutaca meat Hele 7300 Fig. 1. The relation of CO comersion with surfice ‘etal Co amouat Reaction conditions: T=S03K, Pal.OMPa, WF=S ghmol, COM=12. aS ghmol"!, LOMPa, steady sate. should be attributed to the fact that the Co catalyst activity depended on the number of active sites as discussed above. As will be discussed below, cobalt oxide species from cobalt acetate had high dispersion but were difficult to be reduced. On the other hand, cobalt oxide species from cobalt nitrate were easy to be reduced but had low dispersion. ‘The XRD spectra of various catalysts are shown in Fig. 3. This clearly displays thatthe catalysts pre- ppared from mixed salts had smaller particle sizes than that of the catalyst prepared from cobalt nitrate only. In Table 4, the metal particle size calculated from XRD data is generally in accordance with that from Hp chemisorption In case ofthe catalysts where cobalt acetate was the precursor, no cobalt peak was detected in XRD, indicating the fine dispersion of cobalt oxides from cobalt acetate. Pare Sto nm ° Nm ‘00 Fig, 2. The relation of CO conversion, reduction degree and particle sine with cobalt ivate percentage: redvction depre;P: particle size; © CO coaverion S. Sum et ob. /Applied Catalysis A: General 202 (2000) 121-131 ns “Tate 3 CCauiytic behaviors of various ColSiOs catalyst for CO hysrogeaton® Coast 0 convertion () CH sees (C0, seecviy = iow 28 638 085 3NnA 213 625 084 SNA as 67 036 NBA 308 608 08s SAN 313 665 08s SAHIN 33 339 08 10a 62 162 ~ * Reaction condions: SISK, L0MPa, Hy/CO=2, steady sate, WiFaS peat Wma, ~: ot aval ‘Table « ‘The catalyst charcesization results (Causa chemisorption Reduction degree) ‘Reduction drree! Oz XRD Average dy TOF x100s-* ‘TPR 373-1000 K(®) nation at 673K() of CoP (am) Hy wpake __Avenge dy of CoD mol Halg Co (am)" iN 35 204 an w 155 79 3NTA 2B 2 13657 a 10 SNSA 31 1s 636 76 3 101 27 NBA 34 m9 53677 ™ a4 SAIN 36 7 731 6+ 56 ma 80 SASSN 34 m0 799 31 2 62 94 1A 7. 2 46715 13 82 In Table 4, except catalyst SNISA, all catalysts showed very similar TOF value, varying from 0.07 to 0.09s~!. It seems that FTS reaction here. was ‘generally still stracture-nsensitive, ‘The TGA results in Fig. 4 show that the catalyst JON began to be reduced under Hs stream at around ev teat pe) ig. 3. The Xray diftacton specta of oxide stat CoN), COUN, ‘COQNA), CO(AAN) and Co{A) on silica. The total cobalt lading ‘amount Yas IO WLS. x Cop04, © SiO3. ‘SSOK and catalyst 10A began to be reduced just at ‘bout 1000. This phenomenon may be due to the fact that Co? from cobalt acetate had strong interac- tion with the support, and the Co-O bonds were still kept under severe reduction conditions. The catalysts, SNISA, SA/SN and SA+5N, had higher reduction de- ‘grees than catalyst 10A and lower ones than catalyst 10N. ‘The theoretical reduction degree of the catalysts prepared from mixed salts can be calculated from the equation: R=NXRN+AXRA ® where R means the theoretic reduction degree of the catalysts prepared from mixed salts; N and A means ‘molar percentage of cobalt nitrate and cobalt acetate respectively; Ry and Ra were the reduction degree of catalysts 10N and 10A obtained from Op titration at ‘673K. The results calculated from Eq. (4) are listed in Table 5. The results calculated from Eq. (4) were ‘obviously smaller than those obtained from Oe titra- tion. This comparison indicated thatthe reduction de- agree of any catalyst prepared from mixed precursors 26 . Sun et a./Appied Catalysis A General 202 (2000) 121-131 eight Change 96 4 won aon wo woe te Temper Fig. 4. The 1G curves of catalysts Cot), CotWA), COLAMN), CO{AN), COCA) and pute SiO; in 5% Hy diluted by argon stmospher. Te total loading amount was 1OwtS. ‘was enhanced due to the interaction between different cobalt species. 33. The TPR behaviors of co-based catalysts Prepared from mixed cobalt salts ‘The calcinated catalysts exhibited only one de- tectable bulk phase of cobalt oxide, CosO4, from XRD spectra, except catalyst Co(AY/SiOg. No Co304 XRD peaks ‘were observed for Co(AY/SiO2, ir cating that the crystalline size of CosOg was very small. Fig. 5 shows the TPR spectra of impregnated ‘Table 5 ‘The Ha spillover effet on various catalysts (Caalyt Measured reduction Calculated reduction degree! () eps? (%) ion 8 ® NBA 4 a SNSA ST 4s GNA n 10a, 13 3 SAN 50 s SAHSN 56 “6 * Messed rom O; Walon st 673K. Catal from Ba, (0. Reduction Dogro!% Fig. 5. The TPR profiles of diferent catalysts. ColSiO> catalysts starting from different metal salts and the mixture of cobalt nitrate and cobalt acetate, ‘There were three regions in the TPR profile for the reduction of catalyst ION, The first region was from 473 to 600K, the second was between 600 and 800K and the third, a broad region, was located above 800K. Catalyst ION had peaks at 580 and 640K in the first and the second region respectively, as previ- ously found by Rosynek et al. (14). The two peaks have been identified as conversion of Co* to Co™ followed by the conversion of Co™ to Co, and the broad peaks indicate the existence of several species reduced at approximately the same temperature. For the TPR profile of catalyst 10A, it also has ‘wo small peaks at 580 and 840K. It is important to note thatthe first peaks had the same temperature but different intensities for catalysts 10N and 10A. But the second peaks had the different temperature and the intensity was also remarkably different. Some species ‘could only be reduced at above 1000K for catalyst 10A. This means that cobalt acetate with silica support had strong interactions and was hard to be reduced. ‘The reduction degree of catalyst 10A and catalyst 10N in Table 4 were 1.32 and 88.0% from oxygen titration, 15 and 79% from TPR, respectively. These data, as well as TGA results in Fig. 4, also demonstrated that catalyst 10A was hard to be reduced. 5. Sun et oh. /Appled Catalysis A: General 202 (2000) 121-131 m ‘The TPR spectra of sequentially impregnated cat- alysts N/A changed with the N/A ratio, For catalysts 3N/TA, SNISA, TN/3A, all the first peaks appeared at 613K, but the second peak temperatures decreased withthe increase of N/A ratio, For the catalyst SA/SN, the first peak appeared atthe same temperature as N/A catalysts, but the intensity of the second peak was very weak. The TPR profile of catalyst SAA-SN was similar to that of catalyst SA/SN. ‘Unsupported CosO4 exhibited also two peaks [15]. ‘The first one was ascribed to the Co™ to Co™ and the second was due to Co’ to Co metal. The TG results showed that unsupported cobalt nitrate and acetate finished decomposition at 473 and 513K re- spectively, therefore the first peaks (540-660) in Fig. 5 were impossible to be reasoned to the decom- position of bulk cobalt nitrate and cobalt acetate. It ‘can be concluded thatthe first peak at 540-660.K was ‘due to the reduction of trivalent to divalent cobalt, with the interaction from SiOz. However, not all trivalent-cobalt existed inthe form of Co3Og, because the corresponding hydrogen consumption for the se- {quent reduction of CoO to Co did not exhibit only one kind of species responsible forthe peaks in the second region. ‘The second peak (660-1000 K) has tobe ascribed to several different species. Low temperature parts could ‘be attributed to the reduction of divalent cobalt, having weak interaction with the silica, to metallic cobalt. ‘The species reduced at high temperature might have ‘4 somewhat stronger interaction with silica surface. Species reduced at above 1000K, the third region, had stronger interaction with the support, and could not be reduced at 673K, the reduction temperature in the catalyst preparation in this study. So they were probably inactive in CO hydrogenation. 3.4. FT-IR studies on the co-based catalysts prepared from mixed cobalt salts ‘The transmission IR spectra of the adsorbed CO for the reduced catalysts are compared in Fig. 6. For catalyst 10N, an intense peak at 2001 em"! with two shoulder peaks at 2058 and 1937cm™! and a weak peak at 1813em™! were observed. The 2001 cm™! peak was assigned to CO adsorbed on cobalt metal in linear geometry (16-19). The 2058cm~ shoulder peak can be assigned to the surface carbonyl species, ln am [$042 Mey a ea In Ke eee A veme "avenumber car) Fig. 6. PIR spectra of CO adsorbed on the exaysts reduced at oak. ‘Co(CO)x (where X>1), which readily happened to ‘comer sites on the cobalt metal, and the 1937em™! shoulder peak was due to the bridged CO on Co sites (18,20). The 1813cm™! weak peak was assigned to an adsorbed CO onto multifold sites [21]. For catalyst 10A, the peak at 2181em~! was as- signed to the CO adsorbed on Co" (n=2,3) species {21}, indicating that surface cobalt oxide species from supported cobalt acetate were difficult to be reduced. ‘The band of CO adsorbed on Co metal in linear mode shifted from 2001 em on catalyst 10N to 2015em~! here. In general, the blue shift of this peak indicated that the C-O band became strong if the CO adsorbed ‘onto small cobalt particles. Similarly, peak 2065 em! ‘derived from surface carbonyl species Co (CO)x and 1943. cm=! was from CO in bridged adsorption state. All of these peaks (2065, 2015 and 1943cm™!) were so weak thatthe reducibility of catalyst 10A was very low. Fig. 7 shows the TEM photos of catalyst 10N, SNISA and SA/SN. It is clear that the average size of the cobalt crystalline obeyed the series: LON>SNISA>SA/SN, which was the same as the Te~ sult of chemisorption in Table 4. More interestingly, the cobalt crystallin size of catalysts LON and SN/SA ‘was uniform but that of catalyst SA/SN was not. The surface of catalyst SA/SN was divided into several domains where large crystalline and small crystalline existed separately. vs S. Sun eta. /Applied Catalysis A: General 202 (2000) 121-131 Fig. 7. TEM photoraps of extlyts IN, SNISA and SA/SN. mile: 6000 1: 1ON; 2: SNISA: 3: SAISN, 4, Discussion From Table 2 and Fig, 1, it is clear that the CO conversion increased linearly with loading amount ‘of cobalt and metal surface. It is considered that increasing the metal loading just increased the active site number. This should be the prerequisite of the structure-insensitive reaction here. This was also gen- erally true for the catalysts prepared from the mixed precursors, as shown in Table 4 Table 3 and Fig. 2 clarified the relationship be- tween reduction degree, average particle size and ‘overall activity. It is clear that high N/A ratio led to high reduction degree, and large particle size but low activity as the metal dispersion decreased. On the con- trary, low N/A ratio gave small particle size as well as low reduction degree, resulting in low activity. As a compromise effect, N/A ratio of | exhibited the high- est activity while the total loading amount was the Figs. 3 and 4, as well as Table 5 clearly indicated the promotional role of cobalt metal from cobalt nitrate 10 cobalt acetate for the catalysts prepared from mixed precursors. Matsuzaki et al. reported that the added noble metals improved the reduction of cobalt oxide ‘to metal effectively (22,23). The reason was atributed ‘S. Sun etal. /Applied Catalysis A: Generel 202 (2000) 121-131 9 to the spillover effect of hydrogen from noble metal Similarly, considering the obviously different re- ucibility ofthe supported cobalt nitrate and acetate, it is inferred that reduction of cobalt acetate species could be greatly promoted by spillover hydrogen ‘activated on the cobalt metal from cobalt nitrate. The reduced finely-dispersed cobalt acetate contributed to CO conversion effectively Fig. 5 compares the TPR profiles of various cat- alysts. With the increase of NA ratio, the tem- perature of the second peak of the catalyst N/A series decreased, indicating higher reducibility. Cat- alysts SA+SN and SA/SN had large peak area at high-temperature side, showing a similar trend to catalyst 10A to some extent. From FT-IR information in Fig. 6, catalyst ‘SNV/SA exhibited similar peaks to catalyst 10N. Peak 2001 cm", the linearly adsorbed CO, was the main peak. The absorbance in the 1813em~' (multifold) disappeared, an additional band at 1873cm™! was seen which was likely to be related to another kind of bridge-type adsorbed CO. Furthermore, peaks 1937 and 1873cm™! were very strong and formed a ‘road peak-like area between 1800 and 2000cm~', It is clear that the cobalt crystallite size increased and metal dispersion was low, while the reduc- tion degree was still high for this catalyst. From chemisorption data in Table 4, the cobalt crystallite size changed a5 ION>SN/SADSA/SN>SA+SN>10A, while reduction degree series by oxygen titration was IONDSNISADSAISNDSA+SN>IOA. It is clear that IR observation here is in good accordance with the chemisorption result. The formation of bridge-type adsorbed CO was favored for the SN/SA and SA/SN catalysts, whereas the linear type CO was preva- Jent on other catalysts, As has been pointed out, the bridge-type CO is much more active than linear CO [12}. On catalyst SIN/SA, 2064 em! peak also almost disappeared; this led to the conclusion that Co(CO)x ‘on the comer hardly existed and metal average size ‘was enhanced. Although half of the cobelt was from cobalt acetate, it was completely reduced and no peaks (of CO on cobalt oxide were observed. It is referred that impregnated cobalt nitrate first decomposed on silica and formed a layer to weaken the strong inter- action between silica and cobalt acetate, the secondly impregnated species, making the cobalt acetate easily reduced. Very high reducibility of this catalyst SISA ‘was derived from the smoothly-operated spillover effect, by which hydrogen from metal formed from cobalt nitrate activated the reduction of cobalt acetate. Based on the reasons above, this catalyst exhibited the highest activity among all 10wt.% catalysts. Concerning catalyst SA+5N, the absorbance in the 2015cm™! region was much more intense than that ‘of catalyst 10A; this indicated that Ha spillover might have been the main reason for this. Even if the re- duction degree of this catalyst was lower than that of the other mixing-precursor-derived catalysts, it was remarkably higher than that of catalyst 10A. As this catalyst was prepared by coimpregnation of cobalt ac- tate and cobalt nitrate, the cobalt acetate could more easily combine with silica due to the high affinity be- tween silica and the acetate. This competitive adsorp- tion led to a cobalt-acetate-like surface of this catalyst ‘whose IR peaks were similar to those of catalyst 10A. ‘The biue shift of tinearly-adsorbed CO to 2015 em~! and the existence of a peak at 2181.cm™! means that the metallic particle size was small and a part of oxi- ized species stil existe. For catalyst SA/SN, the peak location of linearly- adsorbed CO appeared at 2001 cm™!, due to the red shift from 2015cm™!, But the intensity of this peak ‘was weaker than that of the same peak of catalyst 'SN/SA or catalyst 10N. Similarly, from the chemisorp- tion results mentioned above, the average cobalt size ‘of this catalyst was smaller than that of catalyst SN/SA. (or 1ON, but higher than that of SA-+SN or 10A. Also Co(CO)x species at 2064 em", bridge-type adsorbed CO at 1937, 1873cm~! and multifold-type adsorbed CO at 1813cm~? indicated that cobalt-nitrate-derived cobalt sites existed. It is believed that, due to the red shift effect, the peak of 2035em™! of catalyst SAISN was from the peaks of 2065 em™! of catalyst 10A. Because the cobalt acetate was impregnated to silica previous to cobalt nitrate, spillover effect from cobalt-nitrate-derived metal was not enough to completely reduce all cobalt acetate, which strongly and directly combined with silica. Consequently, its reduction degree was lower than that of catalyst ION or SN/SA. Its surface structure seemed to be a mixture of cobalt-nitrate-derived metallic sites and cobalt-acetate-derived small crystallites. as was also observed by TEM in Fig. 7. Different from catalyst 10N and catalyst SN/SA, catalyst SA/SN exhibited a different cobalt crystalline 10 S. Sun e ah “Applied Catalysis A: General 202 (2000) 121-131 sg 40 2s a C(OAchAHO 373.579 HSK, sagorax, Comets! G3 * COI” WOH” lh ca coz > SBigpegos me se ee = te . 4 Sel : ':cobalaitaelebaljacente size distribution as in Fig. 7. Similar phenomenon ap- peared on SN+SA. It is considered that large crys- tallines were from cobalt nitrate while small ones were from acetate. Itis speculated that the distance between the large crystalline and the small crystalline made the spillover effect dificult 24]. This resulted in the lower reduction degree of this catalyst than that of catalyst SNISA. The uniform crystalline size distribution of the catalyst SN/SA and more CO adsorbed in bridge-type ‘were the reasons of its high TOF as in Table 4. ‘Based on the studies above, a comparison of the cat ‘lyst reduction process during the catalyst preparation ‘of Co(A) and Co(N/A) was elucidated in Scheme 1. 8. Conclusions Using a mixture of cobalt nitrate and cobalt acetate for impregnation resulted in highly active cobalt sil- catalyst. The supported cobalt acetate had stronger interaction with SiO than cobalt nitrate and was dif- ficult to be reduced. It is considered that the reduc- tion of cobalt acetate species could be promoted by spillover hydrogen activated on the cobalt metal from cobalt nitrate, The N/A ratio had a great effect on cat- alyst reduction degree and metal particle size. High reduction degree and high cobalt metal dispersion are ‘wo important factors for FTS catalysts. The synthe- sis of highly dispersed cobalt catalysts required the Cometst on, yi), - "SiO Scheme 1. Proposal reduction procedure of Cotpacetate, Co(tnitateCo(facetate on siica during catalyst preparation, Coaaceate; initial formation of very small CoO or Co304 erys- tallites. The formation of small oxide clusters needs strong interactions between the support and the cobalt precursor. But oo strong interaction between SiOz and small CoO clusters would suppress reduction of, these CoOx clusters; thus, assistance from other co- existing sites, such as cobalt-nitrate-derived metallic sites, was necessary. Using a suitable N/A ratio of pre~ cursors appeared to be a promising synthesis route of ‘active cobalt silica catalysts. Acknowledgements ‘This work was partly supported by New Energy De~ ‘velopment Organization (contract number: 9710005) and JSPS Research for the Future Program (contract ‘number: ISPS-RFTF98PO1001) of Japan, References 11) HH, Storch, N. Golombic, RB. Anderson, Fischer Trpsch and Related Synthesis, Wiley, New York, 1951. (2) H. Sehul, Evan Sten, M. Clays, Sto, Sut. Se. Cal. 81 1994) 204, (3) B.Ieern, Appl. Ca. 161 (1997) 89. (4) K. Fujimoto, M. Kajioka, Bul. Chem. Soe. Jpn. 60 (1987) 237 (5) K Fujimoto, K. Odan, K. Yoshi, Japan Patet No. 03181293, 1952 04138534, 193 and 04138535, 1993 S. Sun et ah. /Appled Catalysis A: General 202 (2000) 121-131 1 (6 D. Schante, 8. Vado, B.A. Blektan, AM. Hilmen, A. Hoff, ‘A, Holmen, J. Catal 156 (1995) 85. (71 RB. Pannell, KS. Chung, CH. Bartholomew, J. Catal. 46 (1993) 340. (8) JM. Zomtak, CH. Bartholomew, J. Catal. 83 (1983) 107. {19} RC. Reve, CH. Bartholomew, J. Ca. 85 (1988) 63. (00) CH. Barhlomen, RJ. Fara, J. Catal. 45 (1978) 4 1] RL. Chin, DIM. Hercules, J. Phys. Chem. 86 (1982) 360. (02) K.Fajimot, M. Kareyama,T: Kunugi, J. Catal 61 (1980) 7 (03) E Iplesia, SL. Sole, RA. lato, J. Catal 137 (1992) 22 (04) MP. Rosyuek, C.A. Polansky, Appl Cal. 73 1991) 97. (5) B. Viswanathan, R. Gopslsishnan, J. Catal. 99 (1986) 30. 16) K. Sato, ¥. Inowe, 1. Kojima, E. Miyajai, 1. Yasumod ‘Chem, Soc, Faraday Tans. 1 (80) (1984) 841. U7] KR Mobana, G. Spor, A. Zeechina, J. Catal 113 (1988) 466 (08) Ma. Dees, T. Sebi, ¥, Iwazawa, V, Pores J. Catal 124 (1980) 530. (193 A. Lapidus, A. Kiylova, V. Kazanski, V. Borovkon, A. Zaisev, J. Ratousky, A. Zukal, M, Janalkova, Appl. Catal 73 (991 6. (20) MJ, Heal, EC. Leisepang, RG. Tomington, J. Catal. $1 978) 314 {211 S. Ici, ¥. Ono, B. Viswanathan, Sut Sei 161 (1985) 349. [22] T Matsuzaki, K. Takeuchi, T Hanaoka, H. Arakawa, Y. Sug, Cau. Today 28 (1996) 251. (23) T: Matai, K.Takewehi, T: Hanaoka H. Arakawa, Y. Sug, ‘Appl. Catal. 105 (1993) 159. (24K. Pajimot, 1. Nakamura, Std. Surf, Si. Catal. 112 (1997) 2.

You might also like