Вы находитесь на странице: 1из 10

F U E L P R O CE SS I NG T EC H NOL O G Y 9 0 (2 0 0 9 ) 2 3 7 2 4 6

a v a i l a b l e a t w w w. s c i e n c e d i r e c t . c o m

w w w. e l s e v i e r. c o m / l o c a t e / f u p r o c

Ruthenium promotion of Co/SBA-15 catalysts with high cobalt


loading for FischerTropsch synthesis
Haifeng Xiong a , Yuhua Zhang a , Kongyong Liew b , Jinlin Li a,b,
a

School of Chemistry and Chemical Engineering, Suzhou university, Suzhou, China


Key Laboratory of Catalysis and Materials Science of the State Ethnic Affairs Commission & Ministry of Education,
South-central University for Nationalities, Wuhan, China
b

AR TIC LE D ATA

ABSTR ACT

Article history:

30 wt.%Co/SBA-15 catalysts with different ruthenium contents (0.050.5 wt.%) were

Received 19 June 2008

prepared by incipient wetness impregnation and characterized by diffuse reflectance

Received in revised form

infrared fourier transform spectroscopy, N2 adsorption-desorption, X-ray diffractometry,

19 August 2008

temperature-programmed reduction and H2 desorption, oxygen titration as well as X-ray

Accepted 29 August 2008

photoelectron spectroscopy. The addition of a small amount of Ru promoter to Co/SBA-15


shifted the reduction temperature of both steps (Co3O4 CoO and CoO Co0) to lower

Keywords:

temperatures and suppressed the formation of Co2+ species. After reduction, ruthenium

FischerTropsch synthesis

atoms were encapsulated partially with cobalt cluster. There was no strong electronic

Cobalt catalysts

interaction between metal cobalt and ruthenium, however, hydrogen spillover from

SBA-15

ruthenium to cobalt oxide clusters occurred. With increasing ruthenium content, catalyst

Ruthenium promoter

reducibility increased and the surface was enriched in cobalt atoms. Moreover, the peak
intensities of both the linear and bridge types CO adsorption increased with the increase of
ruthenium content, enhancing the catalytic activity on FischerTropsch synthesis.
2008 Elsevier B.V. All rights reserved.

1.

Introduction

For gas-to-liquid fuel technology, cobalt catalysts are preferred in FischerTropsch synthesis (FTS) due to the low
activity of the water-gas shift reaction and are believed to
deactivate less rapidly and yield a high fraction of linear
paraffin [1]. The dispersion of active cobalt on the surface of
support requires some interaction between cobalt and support. The existence of such interaction can stabilize the
catalyst against aggregation of active cobalt and deactivation
under FTS conditions. However, strong interaction between
cobalt and support decreases both the reducibility and activity
of cobalt catalyst. Some noble metal promoters such as Ru, Pt
and Re have been employed in cobalt catalysts in order to
increase the reducibility and dispersion. In the study of Iglesia
et al. [2], it is found that the presence of Ru in supported cobalt

catalysts not only enhanced the reduction of cobalt oxide but


also inhibited carbon deposition during FTS. Panpranot and
coworkers [3] studied CO hydrogenation on MCM-41 and SiO2supported Ru-promoted Co catalysts using steady-state isotopic transient kinetic analysis. They concluded that the
CoRu/MCM-41 may be potentially useful for FTS as well as
for other catalytic applications. Kogelbauer et al. [4] studied
the effect of Ru on the Co/Al2O3 catalysts which were prepared
by different preparation method and found that Ru acted as a
reduction promoter for the catalyst by increasing the reducibility and dispersion of the cobalt particles. Tsubaki et al. [5]
found that small amounts of different noble metal (Ru, Pt, Pd)
on Co/SiO2 catalyst have different effects on the reducibility,
turnover frequency (TOF), catalytic activity and CH4 selectivity. For the promotion of Re on the reduction property of Co/
Al2O3 catalysts, Jacobs et al. [6] have done a detailed

Corresponding author. Key Laboratory of Catalysis and Materials Science of the State Ethnic Affairs Commission & Ministry of Education,
South-central University for Nationalities, Wuhan, China. Tel.: +86 27 67843016; fax: +86 27 67842752.
E-mail address: lij@scuec.edu.cn (J. Li).
0378-3820/$ see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.fuproc.2008.08.014

238

F U E L P RO CE SS I NG T EC H NOL O G Y 9 0 (2 0 0 9) 2 37 2 4 6

investigation by in situ EXAFS/XANES and XPS techniques.


They reported that the addition of small amounts of Re
catalyze the reduction of cobalt oxide only by shifting the
reduction temperature of the second step to lower temperature. Thus, the promotion effects of different promoters are
complex [7]. These noble metal might enhance the rate of
reduction of Co species, however, the nature of the promotion
as well as the location of the promoter are unclear. Some
authors have implied that contact of the promoter with the
cluster is not necessary to achieve spillover of hydrogen to
produce cobalt metal, while others believe that intimate
association of the noble metal with the cobalt is important.
SBA-15 is a silica-based mesoporous material with uniform hexagonal channels ranging from 3 to 30 nm and very
narrow pore size distribution [8]. It has enhanced hydrothermal stability as compared to other ordered mesoporous
silicas such as MCM-41 and large surface area (6001000 m2/
g), allowing for the dispersion of a large number of catalytically active species. Besides, the reducibility is favorable for
SiO2 supported Co catalyst because the strength of interaction between the cobalt and support is lower than other
support [9].
The utilization of SBA-15 as a support for preparing cobaltbased FTS catalyst has been recently explored. Wang et al.
[10] prepared highly dispersed Co/SBA-15 catalysts from
cobalt acetate and cobalt acetylacetonate precursor and
found that the obtained catalysts exhibited low FTS activity,
which was ascribed to a low reducibility of the Co species.
Nevertheless, Khodakov et al. [11] found that 5 wt.%Co/SBA15 catalysts were more active and selective toward C5+
hydrocarbons than 5 wt.%Co/MCM-41. Our preliminary
investigation about 5%Co/SBA-15 showed that the catalyst
reducibility, FTS activity and C5+ hydrocarbons selectivity
increased with increasing Ru content [12]. However, Martnez
et al. [13] investigated the influence of cobalt precursor and
loading on the physicochemical and catalytic properties of
Co/SBA-15 catalysts for FischerTropsch reaction. The maximum CO conversion was found at ca 30 wt.% Co loading. We
recently investigated the FTS activity of 30 wt.%Co/SBA-15
with different pore sizes and found that the FTS activity of
30 wt.%Co/SBA-15 of pore size with ca.6.4 nm is the highest in
the range of pore size of 3.413.3 nm [14]. For the preparation
of Co/SBA-15 catalyst, the low activity reported can be
remedied by the addition of noble metal promoter, which
has not been investigated.
In the present study, in situ diffuse reflectance infrared
spectroscopy (DRIFTS) and X-ray photoelectron spectroscopy
(XPS) were used to probe the surface structure of 30 wt.%Co/
SBA-15 catalysts promoted with ruthenium. In addition, in
situ X-ray diffraction (XRD), temperature-programmed reduction (TPR), temperature-programmed desorption, oxygen
titration and N2 adsorptiondesorption were also used to
characterize these catalysts. FischerTropsch synthesis was
performed in a fixed bed reactor. The location of ruthenium on
the Ru-promoted cobalt based catalyst and the effect of
ruthenium on the physical-chemical properties and the
performance of FTS over Co/SBA-15 catalysts have been
elucidated. This investigation will provide the basis for
preparation of highly active commercial FTS Co/SBA-15
catalyst.

2.

Experimental

2.1.

Sample preparation

SBA-15 with pore size of 6.4 nm was obtained using Pluronic


P123 (EO20PO70EO20, MAV = 5800, BASF) and tetraethyl orthosilicate (TEOS, AR) under acidic conditions following the method
reported [9]. The detailed synthesis route was as follow: P123
(10 g) and 2 M HCl (350 mL) were mixed at 35 C to obtain a clear
solution. Subsequently, TEOS (21 g) was gradually added to the
solution with stirring for 48 h. The materials were obtained by
filtration and washing with deionized water. Subsequently,
the precursor was dried in flowing air at 100 C for 10 h. Finally,
the obtained materials were transferred to a muffle furnace
and calcined at 500 C for 5 h.
Co/SBA-15 catalysts with different cobalt loading (5 wt.%
and 30 wt.%) were prepared by incipient wetness impregnation of relative SBA-15 with the desired amount of aqueous
cobalt nitrate. The precursor was dried at 120 C for 12 h,
followed by calcinating at 400 C for 5 h. Co/Ru/SBA-15
catalysts with different Ru contents (0.050.5 wt.%) were
prepared by co-impregnation of SBA-15 using an aqueous
solution containing the desired amount of cobalt nitrate and
ruthenium nitrosyl nitrate. The catalysts were dried overnight
in an oven at 120 C and calcined at 400 C in an air flow for 5 h.
In this study, the Co-loaded samples were labeled as x%Co/
SBA-15 with x standing for the cobalt weight percent in the
samples. The Co catalysts with the addition of ruthenium
promoter were labeled as x%Co/yRu/SBA-15 with y standing
for the ruthenium weight percent in the samples.

2.2.

Catalyst characterization

A Brukers D8 powder X-ray diffractometer with monochromatic


CuK radiation and Ni filter equipped with a VANTEC-1 detector
was used for the XRD measurement. A reactor cell (Anton Paar)
mounted on a goniometer was used for in-situ measurement
under different gas atmosphere. Prior to the XRD measurement,
the calcined catalysts were flushed with high purity argon at
150 C for 1 h, to drive away the water and other impurities, and
then, cooled down to 30 C. The diffractograms were recorded
from 20 to 80 with 0.016 steps. Crystallite phases were
determined by comparing the diffraction patterns with those
in the standard powder XRD files (JCPDS) published by the
International Center for Diffraction Data. After catalyst was
reduction in 450 C for 8 h, Co particle
 size was calculated using
180from the most intense Co
the Scherrer equation d 0:89k
Bcosh  p
peak (2 = 44.2). Where d is the cobalt particle diameter, is the
X-ray wave length (1.54056 ), and B is the full width half
maximum (FWHM) of Co diffraction peak.
N2 adsorptiondesorption experiment was conducted at
193 C using a Quantachrome Autosorb-1. Prior to the
experiment, the sample was outgassed at 200 C for 6 h. The
surface area was obtained using BET model for adsorption
data in a relative pressure ranged from 0.05 to 0.30. The total
pore volumes were calculated from the amount of N2 vapor
adsorbed at a relative pressure of 0.99. The pore size
distribution were evaluated from the desorption branches of
the isotherms using the BarrettJoynerHalenda (BJH) method.

F U E L P R O CE SS I NG T EC H NOL O G Y 9 0 (2 0 0 9 ) 2 3 7 2 4 6

TPR experiment was carried out with a Zeton Altamira AMI200 unit. The catalyst (ca. 0.06 g) was placed in a quartz tubular
reactor, fitted with a thermocouple for continuous temperature
measurement. The reactor was heated with a furnace designed
and built to stabilize the temperature gradient and minimize the
temperature error. Prior to the hydrogen temperature programmed reduction measurement, the calcined catalysts were
flushed with high purity argon at 150 C for 1 h, to drive away the
water or impurities, and then, cooled down to 50 C. Then 10%
H2/Ar was switched on and the temperature was raised at a rate
of 10 C/min from 50 to 800 C (hold 30 min). The gas flow rate
through the reactor was controlled by three Brooks mass flow
controllers and was always 30 cm3 min 1. The H2 consumption
(TCD signal) was recorded automatically by a PC.
Hydrogen temperature programmed desorption was also
carried out in a U-tube quartz reactor with the Zeton Altamira
AMI-200 unit. The sample weight was about 0.100 g. The catalyst
was reduced at 450 C for 12 h using a flow of high purity
hydrogen and then cooled to 100 C under hydrogen stream. The
sample was held at 100 C for 1 h under flowing argon to remove
weakly bound physisorbed species prior to increasing the
temperature slowly to 450 C. At that temperature, the catalyst
was held under flowing argon to desorb the remaining
chemisorbed hydrogen and the TCD began to record the signal
till the signal returned to the baseline. The TPD spectrum was
integrated and the amount of desorbed hydrogen was determined by comparing to the mean areas of calibrated hydrogen
pulses. Prior to the experiments, the sample loop was calibrated
with pulses of nitrogen in helium flow, comparing with the
signal produced from a gas tight syringe injection (100 L) of
nitrogen under helium flow. O2 titration was also performed
with the Zeton Altamira AMI-200 unit. The extent of cobalt
reduction was determined by O2 titration of reduced samples at
450 C. After reduction under the conditions (as described above
for H2-TPD), the catalysts were kept in flowing Ar at 450 C and
the sample was reoxidized by injecting pulses of high purity
oxygen in argon. The extent of reduction was calculated by
assuming metal Co was converted to CoO [11,15]. All flow rates
were set to 30 cm3 min 1. The uncorrected dispersions and
cluster size are based on the assumption of complete reduction,
and the corrected dispersions and cluster size are reported by
percentage reduction. The formula for the calculation has been
shown in previous studies [9,16].
The surface composition of the catalysts were determined
from X-ray photoelectron spectroscopy (XPS), performed by a
Vacuum Generator Mutilab 2000 spectrometer with a monochromatized Al K source (1486.6 eV). The system consisted of an
analysis chamber and a preparation chamber. The analysis
chamber was evacuated to b1 10 8 mbar and the preparation
chamber was evacuated to b1 10 7 mbar. Sample were grinded
and pressed as pellets (c.a.40 mg) and positioned into a high
pressure gas cell placed in the preparation chamber for in situ
hydrogen reduction (60 mL/min), and then, the cell was
evacuated and the sample was transferred to the analysis
chamber through the preparation chamber without exposure
to air. Spectra were collected with a passing energy of 30 eV. All
binding energies (BEs) were corrected referencing to the C 1s
(284.6 eV) peak of the contamination carbon as an internal
standard. The binding energy is estimated to be accurate
within 0.2 eV. The Co 2p binding energy of the core level was

239

determined by computer fitting of the measured spectra.


Surface compositions were calculated from photoelectron
peak areas after correction by the photoionization crosssections and electron escape depth [17]. The quantitative
analysis of surface compositions was involved in a background
subtraction of linear integral profile. The relative atomic
concentration ratio can be obtained by the following formula:


nCo ICo aSi ESi 0:75


nSi
ISi aCo ECo
Where nCo and nSi are the numbers of cobalt and silica atoms at
the surface, ICo and ISi are the intensities of cobalt and silica,
respectively, Co and Si are the photoionisation cross-sections
of the respective elements [18], ECo and ESi are the kinetic
energies of the respective elements.
DRIFTS spectra were recorded with a Nicolet NEXUS 6700
FTIR spectrometer supplied with a MCT detector and a diffuse
reflectance attachment using a spectral resolution of 4 cm 1.
The high purity carbon monoxide (N99.999%) was used as the
probe gas. Helium and hydrogen (N99.999%) were used as the
flushing gas and the reducing gas, respectively. In order to
compare the behavior of the intensities of IR bands among the
studied catalysts, the amount of sample used was the same
for all the experiments. The catalyst (ca. 5 mg) was placed in
an infrared cell with ZnSe windows. The catalyst was reduced
in situ for 12 h under atmospheric pressure with a stream of
hydrogen at 450 C at a flow of 20 cm3 min 1. For the spectra of
CO chemisorbed on catalyst, the system was cooled down to
30 C and the background spectrum was collected. After
introduction of carbon monoxide (30 cm3 min 1) for 2 h, the
catalyst was purged with He (flow rate = 20 cm3 min 1) for
30 min to remove gaseous carbon monoxide before the IR
spectra were recorded.

2.3.

Catalytic evaluation

FischerTropsch synthesis was performed in a tubular fixed bed


reactor (id = 12 mm). The 30%Co/SBA-15 sample (0.5 g, 210 m)
was mixed with 5 g carborundum and reduced in high purity H2
(6 SL h 1 g 1) at atmosphere pressure. The reactor temperature
was increased from ambient to 100 C (hold 60 min), then,
increased to 450 C in 2 h and held at that temperature for 10 h.
Subsequently, the reactor was cooled down to 150 C. Then, the
syngas (CO/H2 = 1:2, 8 SL h 1 g 1) was switched on and the
pressure was increased to 2.0 MPa. The reactor temperature was
raised to 210 C at 1 C/min and the reaction was carried out at
210 C. The products were collected in a hot trap (130 C) and a
cold trap (2 C) in sequence. The outlet gases were analyzed
online by an Agilent 3000 GC and the oil collected at 2 C was
analyzed by an Agilent 6890 GC. Analysis of solid wax collected
at 130 C was performed by an Agilent 4890 GC.

3.

Results and discussion

3.1.

X-ray diffraction

The XRD profiles of low and high diffraction angles of the


calcined samples are displayed in Figs. 1 and 2, respectively.

240

F U E L P RO CE SS I NG T EC H NOL O G Y 9 0 (2 0 0 9) 2 37 2 4 6

Table 1 N2 adsorption desorption data of samples


Sample

SBA-15
30%Co/SBA-15
30%Co/0.5Ru/SBA-15
30%Co/0.1Ru/SBA-15
30%Co/0.5Ru/SBA-15

Surface
area
(m2/g)

897
473
476
467
471

Total
pore
volume
(cm3/g)

Pore
size
(nm)

1.37
0.69
0.67
0.68
0.67

6.46
6.17
6.15
6.15
6.16

Elemental
content
(mass %)
Co

Ru

29.8
29.7
29.8
30.1

0.052
0.094
0.488

high diffraction peaks at low angles. When impregnating 30%


Co to support, the XRD profiles exhibited decreased diffraction
peaks. The decrease of intensity of diffraction peaks suggested
that the mesoporous structure of the catalyst may be partially
destroyed. It is different from the investigation of Khodakov
and coworkers [19]. In their reports, although aqueous
impregnation had an impact on the long range ordering and

Fig. 1 Low angle XRD profiles (a) and pore size distribution
(b) of SBA-15 and catalysts.

As can be seen from Fig. 1, SBA-15 gave three distinct


diffraction lines, indicating p6mm hexagonal symmetry
structure, as reported earlier [8]. After impregnation of 5%Co,
the XRD profiles were almost unchanged and exhibited the

Fig. 2 X-ray diffraction profiles of catalysts.

Fig. 3 In situ X-ray diffraction of 30Co/SBA-15 catalysts with


Ru promoter in high purity H2 (30 mL/min); a: 30%Co/SBA-15,
b: 30%Co/0.5%Ru/SBA-15; : Co3O4, #: CoO, *: Co0 (cubic), :
Co0 (hexagonal).

241

F U E L P R O CE SS I NG T EC H NOL O G Y 9 0 (2 0 0 9 ) 2 3 7 2 4 6

Co/0.5Ru/SBA-15 in H2 at 450 C for 8 h. This suggests that the


addition of ruthenium can promote the formation of hexagonal phase of cobalt.

3.2.

Fig. 4 TPR spectra of 30%Co/SBA-15 catalysts with different


ruthenium content.
mesoporous structure of the support, the impact of aqueous
impregnation on SBA-15 was very small. Some cobalt species
have entered the pore of support. This is evidenced by the
decrease of support pore volume revealed by N2 adsorption
desorption experiment (Table 1). In Fig. 2, the diffraction peaks
at 2 = 31.3, 36.9, 45.1, 59.4 and 65.4 indicated that cobalt
species was mainly in the form of crystalline Co3O4 spinel on
all catalysts [20].
The in situ XRD patterns of 30%Co/SBA-15 and 30%Co/
0.5Ru/SBA-15 catalysts reduced under H2 are shown in Fig. 3.
For the two catalysts, reduction with H2 for 8 h resulted in the
complete reduction of the Co3O4 to CoO and to metal Co. When
reduction temperature reached 450 C, the diffraction peaks of
CoO phase for 30%Co/SBA-15 catalyst can still be found, while
that for 30%Co/0.5Ru/SBA-15 catalyst have disappeared completely when reduction temperature reached 424 C. This
suggested that the addition of ruthenium promoter led to
much easier reduction of 30%Co/SBA-15 catalyst. It should be
mentioned that the diffraction peak of metal ruthenium is not
detected during the reduction process. Tsubaki et al. [5]
reported that the Ru-promoted Co/SiO2 catalyst exhibited a
narrow and intense peak for metal cobalt after reduction
compared with unpromoted catalyst, which was not found in
the present study. It was found that cobalt metal (fcc) formed
after reducing 30%Co/SBA-15 in H2 at 450 C for 8 h, while both
hexagonal and cubic phase of Co0 formed when reducing 30%

Temperature-programmed reduction

The TPR spectra of 30%Co/SBA-15 catalysts with different


ruthenium loadings are shown in Fig. 4. For 30%Co/SBA-15
catalyst, three reduction peaks are shown, located at 311, 465
and 677 C, respectively. Correlated with the in-situ XRD data,
the first reduction peak at 311 C can be attributed to the first
reduction step of Co3O4 (Co3O4 CoO). The second reduction
peak at 465 C is attributed to the reduction of intermediate CoO
phase (CoO Co0). The high temperature reduction peak at
677 C is attributed to reduction of barely reducible cobalt
silicates (Co2SiO4) formed due to a strong interaction between
cobalt and the siliceous support [13]. With the addition of 0.05%
Ru to 30%Co/SBA-15, the peak position of the first reduction of
Co3O4 shifts from 311 to 273 C. However, the peak position of
the second reduction is not shifted significantly. Further, the
reduction peak at high temperature disappeared on adding
0.05%Ru to 30%Co/SBA-15 catalyst due to the suppression of
interaction between cobalt and support on addition of ruthenium. The two remaining reduction peaks shift to lower
reduction temperature with increasing ruthenium loading.
The TPR spectrum of 0.5%Ru/SBA-15 has also been collected
and a small reduction peak at 173 C is observed, assignable to
the reduction of ruthenium oxide to metal Ru.
The TPR result suggests that the promotion is a result of
hydrogen spillover. The ruthenium oxide on 30%Co/Ru/SBA15 catalysts was first reduced and then catalyzed the
reduction of cobalt oxide to metal cobalt subsequently. Jacobs
et al. [6] studied the promotion of Re on the reduction of Co/
Al2O3 catalysts. In their study, the Re2O7 of Re-added Co/Al2O3
catalyst was reduced at the same temperature as the
reduction of Co3O4 to CoO and thus Re affected only the
second reduction step of cobalt oxide (CoO Co0).

3.3.

H2 chemisorption and oxygen titration

The data for hydrogen chemisorption and O2 titration are


present in Table 2. As can be seen, the addition of Ru promoter
significantly affected catalyst dispersion and reducibility. It
is found that the catalyst reducibility increased with increasing
ruthenium content, consistent with the TPR results. The
metal cobalt particle size and dispersion remains relatively

Table 2 H2-TPD data and O2 titration data of prepared 30%Co/SBA-15 catalysts with ruthenium promoter
Sample
30%Co/SBA-15
30%Co/0.05Ru/SBA-15
30%Co/0.1Ru/SBA-15
30%Co/0.05Ru/SBA-15
a
b
c
d
e

H2 desorbed
(mol g 1)

Duncorr a
(%)

duncorr b
(nm)

O2 uptaked
(mol g 1)

Dcorr c
(%)

Reducibility
(%)

dcorr d
(nm)

d' e
(nm)

49.5
48.8
55.9
56.8

1.95
1.92
2.20
2.24

53
53.8
46.9
46.2

1694
1822
1921
2071

3.9
3.6
3.9
3.7

49.7
53.4
56.3
60.7

26.3
28.7
26.4
28.0

12.3
13.0
12.1
11.5

The uncorrected catalyst dispersion.


The uncorrected metal cluster diameter.
The corrected catalyst dispersion.
The corrected metal cluster diameter.
The cobalt metal particle diameter from XRD.

242

F U E L P RO CE SS I NG T EC H NOL O G Y 9 0 (2 0 0 9) 2 37 2 4 6

unchanged with increasing ruthenium content (00.5%Ru).


However, the cobalt crystallite diameter decreased slightly
with increasing Ru loading for low cobalt loading (5 wt.%)
catalyst [12]. It should be mentioned that the particle size based
on H2 chemisorption is larger than the crystal size of cubic Co
determined via XRD. This can be explained by moderate
aggregation of the crystallites and/or cluster formation by the
amorphous fraction [21].

3.4.

X-ray photoelectron spectroscopy

The catalysts were investigated by X-ray photoelectron spectroscopy (XPS). Typical Co 2p spectra are shown in Figs. 5 and 6. For
Co 2p spectra of pure Co3O4, an intense main peak (Co 2p3/2)
located at ca.780.3 eV is observed and weak shoulder
was located at the high binding energy side of the main peak,
which can be ascribed to the shake-up process of Co2+ in the
high spin state [22,23]. Similar to the pure Co3O4, 30%Co/SBA-15
and 30%Co/0.5Ru/SBA-15 catalysts showed the intense Co 2p
peak at 780.3 eV and the weak shake-up satellite peak. This
suggests that the main phase on the catalysts surface is Co3O4,
in agreement with the XRD results. When the catalyst
was reduced in H2 at 450 C for 10 h, the Co 2p3/2 spectra splited
with formation of a shoulder at 777.5 eV. This is due to the

Fig. 5 The Co 2p XP spectra (a) of pure Co3O4 and Ru 3p


spectra (b) of 0.5Ru/SBA-15.

Fig. 6 Co 2p XP spectra of a: 30%Co/SBA-15 catalyst, b: after


reducing 30%Co/SBA-15 at 450 C for 10 h in H2 and
subsequent evacuation, c: 30%Co/0.5Ru/SBA-15, d: after
reducing 30%Co/0.5Ru/SBA-15 at 450 C for 10 h in H2 and
subsequent evacuation.

Co3O4 Co0 transition [6,7]. We have not detected the occurrence of Co2SiO4 and it needed further studies in the future. It
should be mentioned that the XPS spectra of Ru for 30%Co/
0.5Ru/SBA-15 catalyst was not detected, while the Ru 3p spectra
of 0.5%Ru/SBA-15 can be detected (Fig. 5b) [24]. It possibly
suggests that the ruthenium was encapsulated in cobalt cluster
and escaped XPS detection. Recently, Shannon et al. [25]
reported that Ru was present in regions of high concentration
within Co particles by aberration-corrected scanning transmission electron microscopy.
To investigate the changes in the surface composition after
the addition of ruthenium promoter and reduction, the atomic
ratio of Co/Si was calculated. The obtained quantitative results
are displayed in Table 3. For fresh catalyst, the Co/Si ratio
increases with the addition of ruthenium promoter. This
suggests that the ruthenium addition gave rise to increased
metal dispersion. The increased metal dispersion is possibly
due to the effect of ruthenium on the nucleation of cobalt
cluster [26]. For all catalysts, the reduction in H2 led to the
increase of Co/Si ratio. Previous study has reported that the
Co/Ti ratio for Co/TiO2 catalyst changed with reduction
treatment. This is possibly due to the changes in the atomic
densities occurring upon the phase transition from Co3O4 to
Co0 and to a decrease in the inelastic mean free path of the
electrons in Co0 with respect to Co3O4 [27,28]. The present
increase of Co/Si ratio after reduction and re-oxidation should
be assigned to the formation of surface compound between
cobalt and support [28].

F U E L P R O CE SS I NG T EC H NOL O G Y 9 0 (2 0 0 9 ) 2 3 7 2 4 6

Table 3 Atomic ratio on the surface of 30%Co/SBA-15 and


30%Co/0.5Ru/SBA-15 catalysts after different treatment
process derived from XPS measurements
Treatment

As-prepared
After reduction a
After re-oxidation b
a
b

Surface atomic ration


30%Co/SBA-15

30%Co/0.05Ru/SBA-15

Co/Si

Co/Si

0.0141
0.0164
0.0169

0.0213
0.0231
0.0234

Reduced at 450 C for 6 h.


Calcined at 300 C in airflow for 2 h.

3.5.
Diffuse reflectance infrared Fourier
transform spectroscopy (DRIFTS)
The spectra of CO chemisorbed on the surface of 30%Co/Ru/
SBA-15 catalysts with different Ru loading after reduction with
H2 at 450 C for 12 h are shown in Fig. 7. For the 30%Co/SBA-15
catalyst at room temperature, five bands located at 2171, 2120,
2061, 2036 and 1936 cm 1 are clearly resolved, as previously
reported. In addition, the intensity of a band at 1978 cm 1 is
too weak to be discerned clearly [13]. The bands at 2171 and
2120 cm 1 can be assigned to CO adsorbed on relatively high
oxidation state cations (Co2+, Si2+) with -type electron backdonation coordination bond [23]. The band at 2061 cm 1 can be
assigned to CO species on metallic cobalt with weak electrondonor properties (with partial positive charge Co+) [29]. The
band at 2036 cm 1 can be assigned to CO adsorbed on
zerovalent Co0 sites in a linear geometry [30,31]. The peaks
at 1978 and 1936 cm 1 can be assigned to the bridge-type CO
adsorbed on metal cobalt. Song et al. [32] have reported the
speculation that the peaks at around 1937 and 2035 cm 1 were
attributed to CO bands adsorbed on the hexagonal cobalt
phase, while the peaks at around 1974 and 2054 cm 1 could be

Fig. 7 FTIR spectra collected following adsorption of CO at


room temperature on 30%Co/Ru/SBA-15 pretreated in H2 at
450 C and subsequent desorption in He at room temperature
for 30 min; a: 30%Co/SBA-15, b: 30%Co/0.05Ru/SBA-15, c: 30%
Co/0.1Ru/SBA-15, d: 30%Co/0.5Ru/SBA-15.

243

due to CO bands adsorbed on cubic cobalt. We have found that


the band intensity at 1978 cm 1 decreased with the increase of
catalyst pore size, while the band intensity at 1936 cm 1
increased. This suggested that two bands (1978 and 1936 cm 1)
were derived from different bridge-type CO species.
With the addition of ruthenium promoter, it is noted that
the peak intensities of the bands at 2171 and 2120 cm 1
decreased sharply. This indicates that the amount of CO
molecules adsorbed on the high oxidation state cation (Co2+)
decreased. Our TPR and O2 titration results revealed that the
catalyst reducibility increased on addition of ruthenium
promoter. Thus, less Co2+ and Co3+ cations were dispersed
on the catalyst surface. With increasing ruthenium content,
the peak intensities of CO adsorption both the linear and
bridge types increased, indicating that more metal cobalt was
present on the surface of 30%Co/Ru/SBA-15, compared with
30%Co/SBA-15. This is in agreement with the results obtained
from in-situ XRD, TPR and O2 titration.
CO adsorption spectra for 5%Co/SBA-15 are shown in Fig. 8.
Compared with 30%Co/SBA-15, the bands intensity of both
linear and bridged types CO adsorbed on metallic cobalt
decreased for all the low metal loading samples, including
those containing Ru as promoter. It is understandable as lesser
amount of cobalt species was reduced providing less metal
cobalt sites on the support surface. For Ru-added 5%Co/SBA-15
catalyst, the most prominent character is that the band
intensity at 2061 cm 1 increased remarkably on increasing Ru
content from 0 to 0.5%. However, the increased intensity of this
band is not directly related to the catalytic activity because the
intensity of the band for 5%Co/0.5Ru/SBA-15 is stronger than
that for 30%Co/SBA-15 while the activity of the former is lower
than that of the later.
Little investigation on the effect of ruthenium on the DRIFTS
spectra of cobalt catalyst has been reported. Tsubaki et al. [5]
observed that the addition of Ru enhanced the intensity of the
bridge-type adsorbed CO. They concluded that Ru was enriched
at the Co particle surface. In this case, the addition of ruthenium
led to the increase of the peak intensities of CO adsorption both
the linear and bridge types and no frequency shift has been

Fig. 8 FTIR spectra collected following adsorption of CO at


room temperature on Co/SBA-15 with Ru promoter
pretreated in H2 at 450 C for 12 h and subsequent desorption
in He at room temperature for 30 min.

244

F U E L P RO CE SS I NG T EC H NOL O G Y 9 0 (2 0 0 9) 2 37 2 4 6

observed due to the addition of Ru promoter, indicating there is


no strong electronic interaction between metal cobalt and
ruthenium, in agreement with the spillover theory [31].
Similar to the Co/Ru/SBA-15 catalysts, the DRIFTS spectra
of 0.5%Ru/SBA-15 was recorded, as shown in Fig. 9. In contrast
with the DRIFTS of CO on Co/Ru/SBA-15 catalyst, large
intensity peaks of CO adsorbed on Ru have been found at
room temperature [24]. The bands at 2138 and 2074 cm 1 have
been observed on well-dispersed Ru catalysts or supported
catalysts with large particle size and are assigned to the
stretching vibration of tricarbonyl species adsorbed on partially oxidized Ru sites (i.e.Run+(CO)3) [33,34]. The band at
2036 cm 1 may be assigned to CO linearly adsorbed on Ru0
[35,36]. Compared with the above results, no adsorbed CO peak
on Ru for Co/Ru/SBA-15 is observed, possibly because the Ru
atoms were largely encapsulated within Co particles, as
evidenced by the XPS result. Simultaneously, some of the
ruthenium atoms might be located on the edge of cobalt
clusters and took part in hydrogen spillover according to the
TPR results. Thus, some ruthenium atoms are also dispersed
on the support surface after reduction.

3.6.

FischerTropsch synthesis

The changes of FTS activity of these catalysts with time onstream are shown in Fig. 10. Steady state was generally
reached after 45 h TOS for all catalysts. The data of FTS activity
and product selectivity for the catalysts after reaching steady
state are listed in Table 3. As can be seen, CO conversion and
selectivity to C5+ hydrocarbons increases while methane
selectivity decreases with increasing ruthenium content.
The addition of ruthenium to 30%Co/SBA-15 catalysts
decreased the amount of Co2+ and Co3+ species and led to
the enrichment of more cobalt active sites at the surface of
SBA-15. The dispersed cobalt active sites were available for CO
adsorption. It is believed that the bridge-type CO was more
easily formed on large Co particle and was much more active
than linear-type CO because it has a weaker CO bond and

Fig. 10 FTS activity with time on-stream on 30%Co/Ru/SBA15 catalysts (205 C, 2.0 MPa, 8000 h 1, H2/CO = 2.0).

thus can be more easily dissociated to carbon and oxygen,


which is one of the crucial processes in FTS. Our DRIFTS
results demonstrated that the addition of ruthenium led to the
increase of the peak intensities of CO adsorption both the
linear and bridge types and no frequency shift has been
observed due to the addition of Ru promoter, indicating there
is no strong electronic interaction between metal cobalt and
ruthenium. Thus, the increased intensity of CO bands of both
the linear and bridge types led to the increase of CO
conversion. In FTS, the presence of unreduced cobalt oxides
(Co2+ or Co3+ species) can catalyze the water-gas shift reaction
and thus increase the H2/CO ratio [22]. The increase of the H2/
CO ratio near the surface Co0 sites would favor hydrogenation
of the adsorbed species leading to higher methane selectivity.
The decrease of CH4 selectivity can be evidenced by the
decrease of CO2 selectivity in the Table 4. The Ru-promoted
cobalt catalysts (30%Co/0.05%Ru/SBA-15, 30%Co/0.1%Ru/SBA15) showed decreased CO2 selectivity as compared to unpromoted catalyst except that 30%Co/0.5%Ru/SBA-15 catalyst
showed the highest CO2 selectivity. The increase of CO2
selectivity for 30%Co/0.5%Ru/SBA-15 is due to the formation
of cobalt oxide on high water pressure, described below.
Khodakov et al. [11] found an inverse relationship between
methane selectivity and the overall extent of Co reduction in a
series of cobalt-supported mesoporous silicas. However, these

Table 4 Performances of Co/SBA-15 with ruthenium


promoter in a fixed bed reactor
Samples

Fig. 9 FTIR spectra collected following adsorption of CO at


room temperature on 0.5%Ru/SBA-15 pretreated in H2 at
450 C for 12 h and subsequent desorption in He at room
temperature for 30 min.

5%Co/SBA-15
5%Co/0.05Ru/SBA-15
5%Co/0.1Ru/SBA-15
5%Co/0.05Ru/SBA-15
30%Co/SBA-15
30%Co/0.05Ru/SBA-15
30%Co/0.1Ru/SBA-15
30%Co/0.05Ru/SBA-15

XCO
(%)
5.3
6.9
7.0
8.9
29.1
33.4
35.4
37.7

Products selectivity (mol%)


CH4

C2

C3

C4

C5+

CO2

44.8
39.6
34.1
33.9
9.61
8.47
6.83
7.23

1.2
0.8
0.6
0.7
1.2
1.1
0.9
0.9

2.4
2.2
2.4
1.9
5.0
4.4
4.1
4.2

2.4
2.6
2.3
2.2
5.1
4.3
3.9
3.9

47.6
53.6
59.5
60.7
79.0
81.7
84.2
82.9

1.6
1.2
1.1
0.6
0.1
0.09
0.02
0.8

Reaction conditions: 2.0 MPa C, CO/H2=1:2, GHSV=8000 h 1.

F U E L P R O CE SS I NG T EC H NOL O G Y 9 0 (2 0 0 9 ) 2 3 7 2 4 6

authors did not observe significant formation of CO2 in their


catalysts and thus the higher methane selectivity were
ascribed to the presence of either unreduced cobalt species
or to small cobalt particles. By contrast, Martnez et al. [12]
observed that a general parallelism between the selectivity to
methane and CO2, suggesting that the higher methane
selectivity displayed by well-dispersed low-reducible catalysts
could be due, at least in part, to a higher extent of the WGS
reaction occurring on unreduced Co species. We observed the
formation of CO2 among the reaction products and our data
exhibited that the addition of ruthenium to Co/SBA-15 catalyst
suppressed the formation of Co2+ cation on the surface of
catalyst. Then, CH4 selectivity decreased with the addition
of ruthenium.
It is observed that the initial activity of 30%Co/0.5Ru/SBA15 decreased sharply with time on-stream. One would expect
to see a significant loss of activity with TOS if the pores were
blocked during reaction so that most of the Co surfaces within
the pores were inaccessible for reaction [3]. This did not
happen in our study because the other cobalt catalyst
remained stably active under the same FischerTropsch
synthesis conditions. As we know, cobalt particles were
found to be distributed on both the exterior and interior
surfaces of support SBA-15 after reduction [13,14]. The larger
Co particles are present on the external surface while the
cobalt particles within the SBA-15 pores are small (b6.4 nm).
The addition of ruthenium promoter led to high initial
catalytic activity. It is known that very small cobalt particles
can undergo reoxidation in high water pressures [37]. Thus, it
is believed that the deactivation was resulted from the
formation of cobalt oxide under the high water vapor obtained
from high initial activity for 30%Co/0.5Ru/SBA-15.

4.

Conclusions

The addition of a small amount of Ru promoter to 30 wt.%Co/


SBA-15 shifted the reduction temperature of both steps
(Co3O4 CoO and CoO Co0) to lower temperatures and
decreased the amounts of Co2+ species. After reduction,
ruthenium atoms were partially encapsulated within cobalt
clusters with ruthenium in direct contact with cobalt atoms.
A part of Ru atoms took part in hydrogen spillover from
ruthenium to cobalt oxide clusters and strong electronic
interaction between metal cobalt and ruthenium has not
been observed. With increasing ruthenium content, catalyst
reducibility increased and cobalt atom became enriched at
the surface of support. Moreover, the peak intensities of both
the linear and bridge types CO adsorption increase with the
increase of ruthenium content, leading to the increased
activity. The decreased CH4 selectivity is due to the decrease
of Co2+ and Co3+ species at the catalyst surface with the
addition of ruthenium.

Acknowledgement
This work was supported by National Natural Science
foundation of China (20590360, 20773166).

245

REFERENCES

[1] A.Y. Khodakov, W. Chu, P. Fongarland, Advances in the


development of novel cobalt FischerTropsch catalysts for
synthesis of long-chain hydrocarbons and clean fuels, Chem.
Rev. 107 (2007) 16921744.
[2] E. Iglesia, S.L. Soled, R.A. Fiato, US Patent No. 4794099, 1988.
[3] J. Panpranot, J.G. Goodwin Jr., A. Sayari, CO hydrogenation on
Ru-promoted Co/MCM-41 catalysts, J. Catal. 211 (2002) 530539.
[4] A. Kogelbauer, J.G. Goodwin Jr., R. Oukaci, Ruthenium
promotion of Co/Al2O3 Fischer-Tropsch catalysts, J. Catal. 160
(1996) 125133.
[5] N. Tsubaki, S. Sun, K. Fujimoto, Different functions of the
noble metals added to cobalt catalysts for FischerTropsch
synthesis, J. Catal. 199 (2001) 236246.
[6] G. Jacobs, J.A. Chaney, P.M. Patterson, T.K. Das, B.H. Davis,
FischerTropsch synthesis: study of the promotion of Re on
the reduction property of Co/Al2O3 catalysts by in situ EXAFS/
XANES of Co K and Re LIII edges and XPS, Appl. Catal., A Gen.
264 (2004) 203212.
[7] W. Chu, P.A. Chernavskii, L. Gengembre, G.A. Pankina, P.
Fongarland, A.Y. Khodakov, Cobalt species in promoted cobalt
alumina-supported FischerTropsch catalysts, J. Catal. 252
(2007) 215230.
[8] D.Y. Zhao, J. Feng, Q. Huo, M. Melosh, G.H. Fredrickson, B.F.
Chmelka, G.D. Stucky, Triblock copolymer syntheses of
mesoporous silica with periodic 50 to 300 A pores, Science. 279
(1998) 548552.
[9] G. Jacobs, T. Das, Y.Q. Zhang, J.L. Li, G. Racoillet, B.H. Davis,
FischerTropsch synthesis: support, loading, and promoter
effects on the reducibility of cobalt catalysts, Appl. Catal., A
Gen. 233 (2002) 263281.
[10] Y. Wang, M. Noguchi, Y. Takahashi, Y. Ohtsuka, Synthesis of
SBA-15 with different pore sizes and the utilization as
supports of high loading of cobalt catalysts, Catal. Today 68
(2001) 39.
[11] A.Y. Khodakov, A. Griboval-Constant, R. Bechara, V.L.
Zholobenko, Pore size effects in FischerTropsch synthesis
over cobalt-supported mesoporous silicas, J. Catal. 206 (2002)
230241.
[12] Q. Cai, J.L. Li, Catalytic properties of the Ru promoted
Co/SBA-15 catalysts for FischerTropsch synthesis, Catal.
Commu. 9 (2008) 20032006.
[13] A. Martnez, C. Lpez, F. Mrquez, I. Daz, FischerTropsch
synthesis of hydrocarbons over mesoporous Co/SBA-15
catalysts: the influence of metal loading, cobalt precursor,
and promoters, J. Catal. 220 (2003) 486499.
[14] H.F. Xiong, Y.H. Zhang, K.Y. Liew, J.L. Li, Fischer-Tropsch
synthesis: the role of pore size for Co/SBA-15 catalysts, J. Mol.
Catal. A Chem. (in press). doi:10.1016/j.molcata.2008.08.017.
[15] . Borg, S. Eri, E.A. Blekkan, S. Storster, H. Wigum, E. Rytter,
A. Holmen, FischerTropsch synthesis over -aluminasupported cobalt catalysts: effect of support variables, J. Catal.
248 (2007) 89100.
[16] H.F. Xiong, Y.H. Zhang, J.L. Li, Y.Y. Gu, Effect of cobalt loading
on reducibility, dispersion and crystallite size of Co/Al2O3
FischerTropsch catalyst, J. Cent. S. Univ. Technol. 11 (2004)
414418.
[17] T. Baird, K.C. Campbell, P.J. Holliman, R.W. Hoyle, D. Stirling,
B.P. Williams, M. Morris, Characterisation of cobaltzinc
hydroxycarbonates and their products of decomposition, J.
Mater. Chem. 7 (1997) 319330.
[18] J.H. Schofield, HartreeSlater subshell photoionization
cross-sections at 1254 and 1487 eV, J. Electron. Spectrosc.
Relat. Phenom. 8 (1976) 129137.
[19] A.Y. Khodakov, V.L. Zholobenko, R. Bechara, D. Durand,
Impact of aqueous impregantion on the long-range ordering

246

[20]

[21]

[22]

[23]

[24]

[25]

[26]

[27]

[28]

F U E L P RO CE SS I NG T EC H NOL O G Y 9 0 (2 0 0 9) 2 37 2 4 6

and mesoporous structure of cobalt MCM-41 and SBA-15


materials, Micropor. Mesopor. Mater. 79 (2005) 2939.
S. Hinchiranan, Y. Zhang, S. Nagamori, T. Vitidsant, N.
Tsubaki, TiO2 promoted Co/SiO2 catalysts for FischerTropsch
synthesis, Fuel Process Technol. 89 (2008) 455459.
R. Riva, H. Miessner, R. Vitali, G.D. Piero, Metal-support
interaction in Co/SiO2 and Co/TiO2, Appl. Catal., A Gen. 196
(2000) 111123.
H.F. Xiong, Y.H. Zhang, K.Y. Liew, J.L. Li, Catalytic
performance of zirconium-modified Co/Al2O3 for
FischerTropsch synthesis, J. Mol. Catal., A Chem. 231 (2005)
145151.
L. Ji, J. Lin, H.C. Zeng, Metal-support interactions in Co/Al2O3
catalysts: a comparative study on reactivity of support, J.
Phys. Chem., B. 104 (2000) 17831790.
C. Elmasides, D.I. Kondarides, W. Grnert, X.E. Verykios, XPS
and FTIR study of Ru/Al2O3 and Ru/TiO2 catalysts: reduction
characteristics and interaction with a methane-oxygen
mixture, J. Phys. Chem., B. 103 (1999) 52275239.
M.D. Shannon, C.M. Lok, J.L. Casci, Imaging promoter atoms in
FischerTropsch cobalt catalysts by aberration-corrected
scanning transmission electron microscopy, J. Catal. 249
(2007) 4151.
P. Li, J. Liu, N. Nag, P.A. Crozier, In situ synthesis and
characterization of Ru promoted Co/Al2O3 FischerTropsch
catalysts, Appl. Catal. A Gen. 307 (2006) 212221.
F. Morales, F.M.F. de Groot, O.L.J. Gijzeman, A. Mens, O.
Stephan, Mn promotion effects in Co/TiO2 FischerTropsch
catalysts as investigated by XPS and STEM-EELS, J. Catal. 230
(2005) 301308.
R. Riva, H. Miessner, R. Vitali, G. Del Piero, Metalsupport
interaction in Co/SiO2 and Co/TiO2, Appl. Catal., A Gen. 196
(2000) 111123.

[29] M.J. Heal, E.C. Leisegang, R.G. Torington, Infrared studies of


carbon monoxide and hydrogen adsorbed on silica-supported
iron and cobalt catalysts, J. Catal. 51 (1978) 314325.
[30] J.L. Li, N.J. Coville, The effect of boron on the catalyst
reducibility and activity of Co/TiO2 FischerTropsch catalysts,
Appl. Catal., A Gen. 181 (1999) 201208.
[31] L.E.S. Rygh, C.J. Nielsen, Infrared study of CO adsorbed on a
Co/Re/-Al2O3-based FischerTropsch catalyst, J. Catal. 194
(2000) 401409.
[32] D.C. Song, J.L. Li, Q. Cai, In situ diffuse reflectance FTIR study
of CO adsorbed on a cobalt catalyst supported by silica with
different pore sizes, J. Phys. Chem. C. 111 (2007) 1897018979.
[33] K. Hadjiivanov, J.-C. Lavalley, J. Lamotte, F. Maug, J.
Saint-Just, M. Che, FTIR study of CO interaction with Ru/TiO2
catalysts, J. Catal. 176 (1998) 415425.
[34] G.-D. Lei, L. Kevan, Characterization of ruthenium species
generated in H-X zeolite: interaction with carbon monoxide,
nitric oxide, oxygen, and water, J. Phys. Chem. 95 (1991)
45064514.
[35] H. Pfnur, D. Menzel, F.M. Hoffmann, A. Ortega, A.M. Bradshaw,
High resolution vibrational spectroscopy of CO on Ru(0 0 1):
the importance of lateral interactions, Surf. Sci. 93 (1980)
431452.
[36] T. Zubkov, G.A. Morgan Jr., J.T. Yates Jr., O. Khlert, M.
Lisowski, R. Schillinger, D. Fick, H.J. Jnsch, The effect of
atomic steps on adsorption and desorption of CO on Ru(1 0 9),
Surf. Sci. 526 (2003) 5771.
[37] E. van Steen, M. Claeys, M.E. Dry, J. van de Loosdrecht, E.L.
Viljoen, J.L. Visagie, Stability of nanocrystals: thermodynamic
analysis of oxidation and Re-reduction of cobalt in water/
hydrogen mixtures, J. Phys. Chem., B. 109 (2005) 35753577.

Вам также может понравиться