You are on page 1of 21

Fuel 128 (2014) 120–140

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Propylene production via propane oxidative dehydrogenation over


VOx/c-Al2O3 catalyst
Sameer A. Al-Ghamdi a,b, Hugo I. de Lasa a,⇑
a
Chemical Reactor Engineering Centre, University of Western Ontario, London, Ontario, Canada
b
Research & Development Center, Saudi Aramco Oil Company, Dhahran, Saudi Arabia

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 We study propane oxidative


dehydrogenation in a fluidized-bed Propane CH
reactor under O2-free conditions. H3C CH2
Propylene
 We use fludizable VOx/c-Al2O3 CH2 100

catalyst prepared with close control CH3 CH3 5% VOx /Al2O3


7% VOx /Al2O3

of Brønsted acidity. 80 10% VOx /Al2O3

 We employ VOx supported on c-Al2O3

Selectivity (%)
60
catalysts with 5–10 wt.% vanadium
loading. Isolated VO4 surface species Polymeric VOx surface species C3 H6
40
(monovanadate) (polyvanadate)
 We employ NH3-TPD and pyridine
O O O O
FTIR to assess the effect of vanadium
O O ODH
V V V V V 20 COx

loading on acidity. O O O O O O O O O O O

 Reactivity tests gave 11.7–15.1% γ-Al2O3 Support 0


10 11 12 13 14 15 16 17 18
propane conversion and 67.6–85.9% Increasing vanadium loading (5-10 wt.%) C3H8 Conversion (%)

propylene selectivity.

a r t i c l e i n f o a b s t r a c t

Article history: This study contributes with a new vanadium oxide catalyst supported on c-Al2O3 especially designed for
Received 18 November 2013 propane ODH under oxygen-free conditions. This catalyst is prepared with different vanadium loadings
Received in revised form 12 February 2014 (5–10 wt.%) via wet impregnation using ammonium metavanadate as a precursor. The prepared catalyst
Accepted 15 February 2014
is characterized using BET surface area, H2-TPR, NH3-TPD, O2 chemisorption, Laser Raman Spectroscopy,
Available online 11 March 2014
pyridine FT-IR and XRD. The characterization of the supported VOx/c-Al2O3 catalysts by Raman spectros-
copy reveals that monomeric VOx species are dominant at low vanadium loadings. Polymeric VOx species
Keywords:
increase, however, at higher loadings until surface monolayer coverage is reached. Pyridine FTIR and
Propane ODH
Oxidative dehydrogenation
NH3-TPD reveal that this VOx/c-Al2O3 catalyst shows progressive acidity reduction with vanadium addi-
Vanadium oxide tion while compared to bare alumina. This controlled acidity reduction is achieved with a higher Brønsted
Lattice oxygen acidity. The performance of the prepared catalysts is established having in mind possible industrial pro-
CREC Riser Simulator pane ODH applications. This is accomplished using a CREC Fluidized Bed Riser Simulator at 475–550 °C
and at atmospheric pressure. Successive propane injections for ODH experiments (without catalyst
regeneration) over a partially reduced catalyst show that the prepared catalysts display promising pro-
pane conversions (11.73–15.11%) and propylene selectivities (67.65–85.89%) at 475–550 °C. Propylene
selectivity increases while that of COx decreases as the degree of catalyst reduction augments with
consecutive propane injections. Under such oxygen-free conditions oxygen from the catalyst lattice is
consumed via propane ODH. This demonstrates that a controlled degree of catalyst reduction is very
desirable for high propylene selectivity in ODH.
Ó 2014 Elsevier Ltd. All rights reserved.

⇑ Corresponding author. Tel.: +1 519 661 2144; fax: +1 519 850 2931.
E-mail address: hdelasa@eng.uwo.ca (H.I. de Lasa).

http://dx.doi.org/10.1016/j.fuel.2014.02.033
0016-2361/Ó 2014 Elsevier Ltd. All rights reserved.
S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140 121

Nomenclature

ri rate of reaction i (mol/g s)


VOx vanadium oxide surface species Abbreviations
XC3H8 propane conversion (%) ODH oxidative dehydrogenation
Si selectivity of component i (%) MW molecular weight
Yi Yield of product i CREC Chemical Reactor Engineering Center
SBET Brunauer–Emmet–Teller specific surface area (cm2/g) XRD X-Ray Diffraction
Dp particle diameter (nm) TPR Temperature-Programed Reduction
Vatom number of vanadium atoms TPO temperature-programed oxidation
Nopropane initial moles of propane (mols) TPD Temperature-Programed Desorption
Wcat. weight of catalyst (g) LRS Laser Raman Spectroscopy
i representing all chemical species including coke XPS X-ray photoelectron spectroscopy
FID flame ionization detector
Greek symbols TCD thermal conductivity detector
b rate of temperature increase (°C/min) AOS Average Oxidation State

1. Introduction While there is significant research on ODH catalysts [34–38],


there is limited information that deals with light paraffins ODH
The conversion of light alkanes to their corresponding olefins in dense fluidized beds, risers or downers reactors. These studies
has been the subject of intensive research over the last two dec- are however confined to butane ODH [39–42].
ades [1–3]. This has been driven by a worldwide steady increase Propane ODH offers new opportunities for innovative fluidized
on the demand for light olefins [4–20]. Currently, light olefins are bed technology using riser and downer reactors. In this regard, eth-
produced via steam cracking of naphtha or catalytic cracking/dehy- ane ODH over VOx/c-Al2O3 has been successfully produced in the
drogenation. However, these processes suffer from several limita- CREC Riser Simulator in the absence of molecular gas oxygen.
tions related to their high energy requirements, coke formation, Promising ethane conversions and ethylene selectivities have al-
selectivity control and thermodynamic constraints [21–27]. Oxida- ready been demonstrated by CREC researchers [43,44]. Thus, it is
tive dehydrogenation (ODH) has been investigated as an alterna- in the interest of the present study to extend our research and
tive technology for olefins production. ODH does not suffer in explore propane ODH in the CREC Riser Simulator over a
principle from the drawbacks of traditional methods. Oxidative VOx/c-Al2O3 catalyst in an oxygen-free atmosphere.
dehydrogenation of propane is of particular importance with pro- For propane conversion to propylene via ODH, supported vana-
pane being a main component of natural gas. This makes propane dium oxide catalysts have been reported as the most active and
a preferable raw material, to be a substitute of naphtha in the selective catalysts [1,32,45–55]. This is due to the vanadium ability
manufacturing of propylene [28]. to provide lattice oxygen for hydrogen removal from alkanes
Compared to the current olefin production technologies, pro- [2,56]. Several studies in the literature have shown that supported
pane oxidative dehydrogenation has the advantages of being a vanadium catalysts are more active in ODH reactions with surface
low temperature process and having an enhanced catalyst lifetime VOx coverage in the sub-monolayer region [32,48,57–60]. How-
due to the presence of oxygen that limits coke formation. More- ever, due to the higher reactivity of propylene while compared to
over, as an exothermic process, propane ODH can overcome the propane, high propylene selectivities are only attainable at low
thermodynamic limitations of non-oxidative dehydrogenation propane conversions. Thus, a catalyst for ODH should be designed
through the formation of water product. However, the low selec- in such a way that it can selectively convert propane to propylene
tivity toward the desired propylene product is still a main issue without further oxidation of propylene to carbon oxides.
in propane ODH process. This has been in fact, a major ODH draw- The catalytic behavior of supported vanadia catalysts in ODH
back limiting ODH opportunities for reactor scale up. Propylene reactions is generally controlled by three factors: (i) the structure
higher reactivity than propane reactivity in ODH favors a compet- of the VOx surface species; (ii) the redox properties of the VOx sur-
itive propylene combustion with the undesired formation of addi- face species; and (iii) the acid–base character of the catalysts and
tional amounts of carbon oxides [2,29–33]. support material. These three factors can be strongly influenced
Regarding ODH, new reactor concepts might also help to make by the type of support and the vanadium loading [32,57,61–63].
oxidative dehydrogenation a feasible alternative for olefin produc- The structure of alumina-supported vanadia has already been con-
tion. Fluidized bed reactors offer some outstanding advantages in sidered in several studies [1,8,32,48,57,64–68]. At low vanadium
comparison to conventional reactor systems, namely: controlled loading, isolated VOx surface species (monovanadates) are exclu-
isothermal conditions, uniform residence time distributions and sively present. As VOx surface density increases with loading, sur-
the absence of mass transfer limitations. Thus, one can circumvent, face structures evolve from isolated monovanadates to polymeric
with fluidized beds, the potential issues with hot spots in fixed bed polyvanadates until monolayer coverage is reached. A crystalline
catalytic reactors that may interfere with reactor performance and V2O5 phase was reported at high vanadium loading in addition to
damage the catalyst. Moreover, fluidized-bed reactors with peri- amorphous vanadia phases. In this respect, variations of lattice
odic catalyst re-oxidation offer the means that enable the transport oxygen binding strength in different VOx surface species formed
of reduced catalyst species from the oxidative dehydrogenation on the surface of metal oxide supports have been suggested. This
unit (zone) to the re-oxidation unit (zone). Therefore, the operation appears to be a main factor affecting the catalytic performance of
of twin reactors, one for the ODH and one for catalyst regeneration, alumina-supported vanadia catalysts in ODH reactions [66,69]. In
appears to be an attractive alternative for the implementation of fact, it is generally accepted for alkane ODH that isolated tetrahe-
this technology at the industrial scale. dral VOx species (which are obtained at low V-loadings) are more
122 S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140

selective but less active than polymeric VOx species [70]. On the acid (Sigma Aldrich, 99%) in a 1:2 weight ratio at a pH of 2. This
other hand, there is a relationship between catalyst reducibility solution was prepared by mixing ammonium metavanadate
and VOx surface structures on a given metal oxide support. In this (NH4VO3) with oxalic acid, heating and stirring the mixture until
respect, the reducibility of the surface VOx species increases with clear before impregnating the support. Vanadium concentrations
surface VOx coverage. Thus, the following trend for the reducibility in the solution were varied in order to ensure the desired VOx con-
of the different supported vanadia species can be considered: poly- centration in the catalyst. The impregnation was performed at
meric surface VOx > isolated surface VOx > crystalline V2O5 nano- 70 °C with continuous stirring. The solvent was then evaporated
particles [49,57,71–73]. Moreover, the extent of the reducibility under vacuum and the resulting cake was dried in air at 120 °C
of supported vanadium oxide catalysts is affected by the type of for 16 h. Finally, the particles were calcined under a gas chroma-
support used. This variation of vanadium oxide reducibility on dif- tography quality air stream at 600 °C for 4 h. Once the catalyst
ferent metal oxides supports could be related to the reducibility of samples prepared, the bulk amount of vanadium in the catalyst
the different VAO-Support bonds existing on different support samples were determined using Inductively Coupled Plasma (ICP).
types [50,74,1,57].
Several studies in the technical literature have shown that the 2.2. Catalyst characterization
acid–base support character when using vanadium oxide catalysts
for ODH plays an important role in determining olefin selectivity 2.2.1. BET surface area
[57,62,70,73,75–78]. This was explained due to the fact that the The specific surface area and pore size distribution of the bare
acid–base character of the support can control the vanadium cata- c-Al2O3 support and the VOx/c-Al2O3 catalysts were determined
lyst reactivity/selectivity. This occurs as a result that the acid–base in an ASAP 2010 Analyser manufactured by Micromeritics using
support properties influence both the reactants and product Nitrogen adsorption–desorption at 77 K. Before each measure-
adsorption/desorption. Thus, acidic catalysts should favor basic ment, a 0.2–0.30 g catalyst sample was degassed at 250 °C for 2 h
reactant adsorption and acidic product desorption, thereby pro- in order to ensure a clean and dry surface. The adsorption iso-
tecting these chemical species from further oxidation to carbon therms were measured in 106 to 1 pressure ranges and the surface
oxides [79]. area and pore diameters were calculated using the method of
For instance, higher selectivities to ethylene have been obtained Brunauer, Emmett and Teller (BET).
over acid catalysts like the VOx/c-Al2O3 catalysts. These findings
were justified as a result of the fact that higher support acidity de- 2.2.2. X-ray Diffraction (XRD)
creases the interaction between the ethylene product and the cat- Powder X-ray Diffraction (XRD) measurements were considered
alyst [59]. In line with this, the significance of Lewis acidity for to identify the crystallographic structure of the prepared
ethane ODH was shown in our previous studies [43,44]. In these VOx/c-Al2O3 catalyst. The XRD diffraction patterns of the catalyst
studies, ethane ODH was reported using a VOx catalyst supported samples and the support material were obtained using an Ultima
on a chi-type alumina (c-Al2O3) support with controlled Lewis IV instrument from Rigaku Instruments with a monochromatic
acidity. For propane ODH reactions, low acidity with dominant Cu Ka radiation source (wavelength = 0.15406 nm, 40 kV, 40 mA)
Brønsted acid sites is desirable to facilitate rapid propylene desorp- and a normal scan rate of 2° scan per minute within the 2h range
tion, thus, avoiding its further oxidation to COx. However, the acid- from 10° to 90° with a step size of 0.02.
ity has to be kept at levels that can favor propane adsorption,
yielding, in this manner, acceptable overall catalytic activity [73]. 2.2.3. Temperature Programed Reduction (TPR)
Thus, there is a delicate balance in ODH between the support Temperature Programed Reduction (TPR) experiments on the
acid–base character and the support acid–base strength. VOx/c-Al2O3 catalyst were used to assess the reducibility of the
Considering the different acidic properties of alkane feeds as catalyst samples. They were conducted using a Micromeritics
well as of olefins products and consistent with our efforts to estab- AutoChem II 2920 Analyzer. For each experiment, an amount of
lish a catalyst for ODH, a new VOx/c-Al2O3 fludizable catalyst for 100–150 mg of catalyst sample was first pretreated for 2 h at
propane ODH was prepared in the present study. This catalyst 250 °C in a 5% O2/He gas mixture at a flow rate of 50 ml/min. This
was envisioned to have a controlled acid–base character on a gam- was followed by cooling down the sample to ambient temperature
ma-alumina (c-Al2O3) support. In line with this, the selected using the same flow rate. Then, the pretreated sample was flushed
c-Al2O3 type of support was thought to have a tailored low acidic with Argon (99.99%) at a 50 ml/min flow rate for 30 min. Hereafter,
character, resulting in VOx/c-Al2O3 catalysts with high Brönsted the argon flow was replaced by a flow of 10% H2/Ar gas mixture at a
acid site density. Moreover, the new c-Al2O3 support was envi- flow of 50 ml/min while heating the catalyst to 900 °C at a heating
sioned as having a low degree of dehydroxylation during the ther- rate of 10 °C/min. Hydrogen consumption was monitored at the
mal pre-treatment phases, with most of hydroxyl groups exit gas stream by using a thermal conductivity detector (TCD).
remaining on the support surface. The amount of hydrogen uptake by the sample can be calculated
Given the above described facts, the goal for this study is to re- through numerical integration of the area of the TPR profile.
port the performance of a new VOx/c-Al2O3 fludizable catalyst for
propane ODH in the absence of gas phase oxygen. This 2.2.4. NH3 Temperature Programed Desorption (TPD)
VOx/c-Al2O3 catalyst is prepared with carefully chosen VOx load- Temperature programed desorption of ammonia (NH3-TPD)
ings. This is done to ensure that surface sub-monolayer VOx cover- experiments was used to determine the acid property of the cata-
age is achieved, with a close control of acidity and Brønsted sites. lysts. The AutoChem II Analyzer from Micrometrics was used for
NH3-TPD experiments. For each experiment, an amount of 100–
2. Materials and methods 150 mg of catalyst sample was first pretreated for 1 h at 550 °C
in a 5% O2/He gas mixture at a flow rate of 50 ml/min. Then, the
2.1. Catalyst preparation samples were cooled to 100 °C and brought to saturation using a
4.55% NH3/He gas mixture at a flow rate of 50 ml/min for 1 h. After
The VOx/c-Al2O3 catalyst samples used in this investigation that, the ammonia flow was switched off, and replaced by a flow of
were prepared by the wet saturation–impregnation of a commer- a pure inert gas (He) at a rate of 50 ml/min for 1 h at the same tem-
cial porous c-Al2O3 support (SASOL, CataloxÒ SCCa 5/200), with perature to remove physically adsorbed ammonia. The tempera-
an aqueous solution of NH4VO3 (Sigma Aldrich, 99%) and oxalic ture was then raised up to 600 °C at a rate of 10 °C/min. As the
S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140 123

temperature was increased, the ammonia desorbed as it gained en- lished using the Chemical Reactor Engineering Center (CREC) Riser
ough energy to overcome the activation energy barrier. The NH3 Simulator. The CREC Riser Simulator is a bench scale mini-fluidized
concentration of the effluent gas from the sample was monitored bed reactor invented by de Lasa (1992) [80]. This mini-fluidized
by a thermal conductivity detector (TCD). reactor (capacity of 50 cm3) operates under batch conditions and
is designed for catalyst evaluation and kinetic studies under fluid-
2.2.5. Pyridine FTIR ized bed (riser/downer) reactor conditions. A detailed description
The nature of catalyst surface acid sites was determined using of the CREC Riser Simulator, the operational procedure and system
Fourier Transform Infrared Spectroscopy (FTIR) and by using pyri- set-up have already been described in a recent publication by
dine as probe molecule. Before the analysis, the catalyst samples Al-Ghamdi et al. [44]. A schematic diagram of the CREC Riser
were dried for 2 h under N2 flow at 550 °C and then cooled to Simulator’s components and the fluidization patterns inside the
100 °C. The samples were kept at 100 °C and saturated with pyri- Riser Simulator catalysts are provided in Fig. 1a.
dine using a N2 stream containing pyridine. Adsorption of pyridine Thermal and catalytic experiments of propane ODH were car-
was allowed for 1 h. Then, the samples were flushed with pure N2 ried out at four temperature levels (475, 500, 525 and 550 °C),
at 100 °C for 90 min, in order to remove weakly adsorbed pyridine. one catalyst loading (0.76 g) and four contact times (5, 10, 15
Following this, samples were placed in a diffuse reflectance infra- and 20 s). All thermal and catalytic runs were repeated 3 times
red spectroscopy (DRIFTS) cell. FTIR spectra were recorded at room to ensure reproducibility of the experimental results. The carbon
temperature using a Bruker IFS55 FTIR spectrometer operating at a mass balance closures, which considered CO, CO2, CH4, ethylene,
4 cm1 resolution and 100 scans. ethane, propylene, propane and carbon deposited over the catalyst
were in the ±2% range. Total carbon formed on the spent catalyst
2.2.6. O2 pulse chemisorption was measured by a TOC Analyzer with a spent catalyst sample that
Oxygen pulse chemisorption was used to determine the degree had been used in the most severe conditions in this study (i.e. after
of dispersion of the prepared VOx/c-Al2O3 catalysts. Experiments 10 consecutive propane injections, 20 s contact time and 550 °C).
were performed using the AutoChem II 2920 Analyzer from Total carbon formed was found to be in the range of 1–1.5 wt.%.
Micromeritics. For each experiment, an amount of 100–150 mg of The effluent from the reactor is analyzed online by a gas chro-
fresh catalyst was first pretreated for 1 h at 550 °C in a 5% O2/He matograph (Shimadzu GC-2010) equipped with three packed col-
gas mixture at a flow of 50 ml/min. This was followed by cooling umns: HayeSepD 100/120 mesh (30 ft  1/800 O.D. S.S., Supelco)
of the sample to ambient temperature under the same flow rate. column, Carboxen-1000 60/80 mesh (15 ft  1/800 O.D. S.S., Supe-
Then, the pretreated sample was flushed with Argon (99.99%) at lco) column and Carboxen-1004 80/100 mesh (6.5 ft  1/1600 O.D.
a flow rate of 50 ml/min for 30 min to remove any adsorbed oxy- S.S., Supelco) column. The Carboxen-1000 and Carboxen-1004 col-
gen. Hereafter, the samples were pre-reduced at 370 °C for 2 h in umns were used for separating H2, O2, N2, Ar, CO, and CO2, and they
a flow of 10% H2/Ar gas mixture at a flow rate of 50 ml/min. Sam- were connected in series to a thermal conductivity detector (TCD).
ples were then purged for 30 h at the same temperature using pure On the other hand, all hydrocarbons (methane, ethane, ethylene,
He gas (99.99%) at a 30 ml/min flow rate to remove physically ad- propane and propylene) were separated in the HayeSepD column
sorbed hydrogen. Subsequently, pulses of a known volume of oxy- and detected by a flame ionization detector (FID).
gen (1 ml loop volume, 5% O2/He gas mixture) were injected into The performance of the catalyst in the oxidative dehydrogena-
the carrier gas (He) until the saturation of the sample was attained. tion of propane was examined on the basis of propane conversion
A thermal conductivity detector (TCD) was employed to quantify and selectivities. Propane conversion, selectivity to a carbon con-
oxygen at the exit gas stream. Oxygen chemisorption uptake was taining product and propylene yield were defined as follows:
determined as the difference of two successive adsorption iso- P
i mi n i
therms measured at 370 °C. The dispersion was then expressed Propane Conversion; X C3 H8 ð%Þ ¼ P  100 ð1Þ
3npropane þ i mi ni
as the ratio between the oxygen uptake and the vanadium atoms
available at the catalyst surface (O/Vsurf), assuming an O to V
mi ni
chemisorption stoichiometry of 1:1. Selectivity to a product; Si ð%Þ ¼ P  100 ð2Þ
i mi ni

2.2.7. Laser Raman Spectroscopy (LRS)


XC3 H8 ð%Þ  SC3 H6 ð%Þ
Laser Raman Spectroscopy (LRS) was employed to gain funda- Propylene Yield ðYÞ ¼ ð3Þ
mental information about molecular structures of different vana- 100
dium oxide surface species (VOx) on a c-Al2O3 support at where ni is the moles of gaseous carbon containing product i, mi is
different V loadings. Laser Raman spectra were obtained using the number of carbon atoms in gaseous carbon containing product
dehydrated catalyst samples in a Renishaw Model 2000 Raman i and npropane is the moles of unconverted propane in the product
Spectrometer equipped with a thermoelectrically cooled CCD stream.
detector (73 °C). The samples were excited with a 633 nm Ar Fluidizability is an important catalyst property for its use in flu-
ion laser line. A laser power of approximately 2 mW irradiating idized-bed reactors. Thus, to confirm its adequacy for fluidized bed
the sample was used. Most of the samples were analyzed at full conditions, the fluidizability of the prepared catalysts was assessed
power; only the un-supported V2O5 sample was analyzed at 10% based on their particle size distribution and apparent density. The
power as the signal was too strong. The sample dehydration was particles size distribution (PSD) of the prepared VOx/c-Al2O3 cata-
carried out at 450 °C under dry air for 2 h and then cooled to room lysts was measured using a Mastersizer 2000 from Malvern Instru-
temperature. Following this, Raman spectra were obtained at room ments. The PSD of the various prepared VOx/c-Al2O3 catalysts
temperature at a spectral resolution of 1 cm1. Sample exposure displayed a narrow particle size distribution around 90 lm. Almost
times were typically 10 and 30 s, with this being a function on 50% of the amount of catalyst particles was between 40 and 80 lm
the vanadium loading. in size while the other 50% of catalyst particles were in the range
between 80 and 110 lm.
2.3. Propane ODH catalytic test The apparent particle density (AD) of the catalyst samples was
assessed using a method established at the CREC [81]. Apparent
The catalytic reaction runs of propane oxidative dehydrogena- particle density values for the various VOx/c-Al2O3 catalysts used
tion over supported vanadium oxide catalyst samples were estab- are reported in Table 1a.
124 S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140

Fig. 1a. Overview of the CREC riser simulator reactor body. Adapted from Al-Ghamdi et al. [44].

Table 1a vanadium loadings. The surface area of the fresh (activated) bare
Apparent particle densities of VOx/Al2O3 catalysts. c-Al2O3 support was found to be 203.52 m2/g. After calcination at
Sample Apparent particle density (g/cm3)
600 °C during 4 h, the metal-free calcined c-Al2O3 support showed
a slight decrease in the specific surface area to 190.84 m2/g. This
c-Al2O3 1.21
mild reduction in surface area was attributed to the collapse/block-
5% VOx/c-Al2O3 1.53
7% VOx/c-Al2O3 1.70 age of some of the support pores during calcination. Furthermore,
10% VOx/c-Al2O3 1.81 after impregnation with the vanadium metal and calcination at
600 °C for 4 h, the specific surface area of the V-loaded catalyst
samples displayed a further gradual decrease with vanadium load-
ing. This VOx/c-Al2O3 surface area reduction with the vanadium
addition can be ascribed to the blocking of some of the alumina
pores by VOx species.
Moreover, Fig. 2 shows the pore size distributions for the bare
c-Al2O3 support and the VOx/c-Al2O3 catalyst samples with differ-
ent vanadium loadings, as determined from the desorption branch
of the N2 isotherms. The pore size distribution displays a 7.2 nm
average for the samples. A very small number of pore counts have
pore size values below 2 nm and almost none above 50 nm. This
indicates that very few of the pores lie outside of the mesoporous
range of 2–50 nm.

3.1.2. VOx surface species and monolayer coverage


Table 1 reports the nominal vanadium oxide surface density on
Fig. 1b. Geldart diagram for classification of particles fluidization [82]. a c-Al2O3 support. Vanadium surface density is defined as the
number of vanadium atoms present per unit specific surface area
Using the average particle size, the particle apparent density
of the catalyst and is expressed in terms of Vanadium atoms per
obtained above, the fluidization regime of the VOx/c-Al2O3 parti-
nm2. In order to find the VOx surface density for the various cata-
cles were determined, as shown in Fig. 1b using Geldart’s powder
lyst samples, the vanadium weight fraction and BET surface area
classification chart [82]. These characteristics were further con-
of each sample are required as shown in the following equation:
firmed experimentally using a Plexiglas model of the CREC Riser
Simulator, specially manufactured for flow visualization [81]. Surface density ðVatoms =nm2 Þ ¼
Finally, the chemical species transport processes from the bulk
wt:% Vanadium loading  6:023  1023
of the fluid to the particle external surface and inside the particles ð4Þ
Molecular Weight of Vanadium  100  Surface Area
are neglected [81].
Fig. 3a and b describes the catalyst specific surface areas and the
calculated VOx surface densities for the VOx/c-Al2O3 catalysts as a
3. Results and discussion function of vanadium loading. It can be seen that the presence of
VOx species only had a modest effect on the specific surface area
3.1. Catalyst characterization of the support as shown in Fig. 3a. As a result and as shown in
Fig. 3b, VOx surface densities increased almost linearly with
3.1.1. Surface area increasing vanadium content.
Table 1b reports the BET surface areas of the bare c-A12O3 Regarding vanadium oxide coverage, the maximum amount of
supports and the VOx/c-Al2O3 catalyst samples with different amorphous two-dimensional vanadium oxide surface species
S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140 125

Table 1b
Characteristics of alumina support and supported vanadia catalyst.

Sample V contenta (wt.%) SBET (m2/g) Vpb (cm3/g) Dpc (nm) Surface density (V/nm2) % Monolayerd
c-Al2O3 (activated) 0 203.5 0.54 7.13 – –
c-Al2O3 (calcined) 0 190.8 0.52 7.27 – –
5% V/c-Al2O3 5 188.6 0.45 6.74 3.1 31.3
7% V/c-Al2O3 7 179.2 0.46 7.46 4.6 46.2
10% V/c-Al2O3 10 153.3 0.39 7.17 7.7 77.1
a
V content, in wt.% was determined by ICP.
b
Vp is a single-point pore volume calculated from the adsorption isotherm at P/Po = 0.99.
c
The average pore sizes were calculated by applying the BJH method on the desorption branches of the isotherms.
d
Calculated assuming surface area of 0.1 nm2 per VO34 species.

monolayer coverage. It is at these conditions at which crystalline


3.0
Al2O3 V2O5 surface species start forming. For such a surface density
5% VO x/Al2O3 range, VOx surface species exist simultaneously with different frac-
2.5 7% VO x/Al2O3
tions as isolated vanadate species (monovanadate) and polymeric
10% VO x/Al2O3
vanadate species (polyvanadates). However, the 5% VOx/c-Al2O3
2.0 sample shows a surface density value of 3.1 which is near the the-
dV/dlog (D)

oretical monolayer coverage of monovanadate surface species.


1.5
3.1.3. H2-TPR and degree of reduction
1.0 TPR experiments were used to investigate the reducibility of
vanadium oxide surface species deposited on the c-Al2O3 support
0.5
H2-TPR profiles of bulk V2O5 and VOx/c-Al2O3 samples with differ-
ent vanadium loadings are shown in Fig. 4. The reduction profile
for bulk V2O5 exhibits four major peaks at 650, 690, 740 and
0.0
1 10 850 °C. Similar observations on the reduction of bulk V2O5 were re-
Pore Diameter (nm) ported by Bosc et al. [84] and Koranne et al. [85]. These authors ob-
served similar multiple reduction peaks when bulk V2O5 was
Fig. 2. Pore size distribution for the c-Al2O3 support and the VOx/c-Al2O3 catalyst treated in a 5% H2 in argon up to 1000 °C. The presence of multiple
samples determined from nitrogen desorption isotherms.
peaks is attributed to the reduction of V2O5 to V2O3 through the
intermediate formation of oxide species with different oxidation
(VOx) in contact with the oxide support as a monomolecular layer states according to the following sequence:
is called monolayer coverage. The horizontal lines in Fig. 3b show
V2 O5 ! V6 O13 ! V2 O4 ! V2 O3
the VOx surface densities corresponding to theoretical monolayer
coverage of monovanadates (2.9 Vatoms/nm2) or polyvanadates On the other hand, H2-TPR profiles for the VOx/c-Al2O3 catalyst
(8.8 Vatoms/nm2). These values of theoretical monolayer coverage samples show only a single symmetric reduction peak extending
were found based on structural calculations given by Centi [83]. from 310 °C to 550 °C with a Tmax between 432 and 452 °C. How-
The values of surface density were found to be 3.1, 4.6 and ever, the 10% VOx/c-Al2O3 sample shows a second weak peak at
7.7 Vatoms/nm2 for the 5% VOx/c-Al2O3, 7% VOx/c-Al2O3 and 10% about 575 °C. The first low temperature reduction peak observed
VOx/c-Al2O3 catalyst samples, respectively. These values are be- in all VOx/c-Al2O3 prepared samples can be assigned to the reduc-
tween the 2.9 and 8.8 Vatoms/nm2 values for theoretical monolayer tion of amorphous monomeric and polymeric VOx surface species.
coverage of monovanadate and polyvanadates, respectively. This The second highest temperature peak displayed for the 10%
allows us to argue that the VOx/c-Al2O3 catalysts of the present VOx/c-Al2O3 sample can only be assigned to the reduction of bulk
study are likely to be in the monolayer coverage configuration V2O5-like surface species. It can also be noticed in Fig. 4 that the
which is below the upper limit for theoretical polymeric vanadia first reduction peak shifts to a lower temperature with increasing

300
(a) 10 (b)
Polyvanadate
VOx Surface density (Vatoms/nm2)

250
8
Surface Area (m2/g)

200

6
150

4
Monovanadate
100

2
50

0
0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
V content (wt.%) V content (wt.%)

Fig. 3. BET surface areas (a) and VOx surface densities (b) for supported vanadium oxide catalysts.
126 S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140

0.14 reported in Table 2. In a similar manner, the degree of reduction


was developed assuming that V2O5 was the sole initial species
0.12 present on the supported catalyst. With these data, the Average
Oxidation States (AOS) of reduced vanadium oxides were
Bulk V2O5
0.10 determined.
TCD signal (a.u.)

On this basis, the oxidation states of vanadium oxides found


0.08 were in the V2O3.02–V2O3.37 range. One should notice that the most
10% VO x/γ-Al2O3
likely vanadium oxide structure around this value is V2O3. This
0.06 vanadium species has been reported by several authors as a very
7% VO x/γ-Al2O3 stable structure among various vanadium oxides [86,87]. This sug-
0.04 gests that the vanadium phase after complete reduction of V2O5 in
5% VO x/γ-Al2O3 the TPR experiments is the V2O3 or mixed phases of V2O3, with
0.02 V3O3.4 and V3O3.7.

0.00
50 150 250 350 450 550 650 750 850 3.1.4. NH3-TPD
Temperature (οC) In addition to the redox properties, surface acidity is a very
important aspect of supported vanadia catalysts. Surface acidity
Fig. 4. H2-TPR profile of bulk V2O5 and calcined VOx/c-Al2O3 catalyst samples: can determine their catalytic activity; and it is governed by the
temperature up to 900 °C (operating conditions: heating rate: 10 °C/min; reducing
agent: 10 mole% H2; balance Ar at 50 cm3/min; oxidizing agent: 5 mole% O2;
nature of the support used and the surface structure of surface van-
balance He at 50 cm3/min). adia species (VOx). Thus, a correlation of surface acidity with the
structure of supported vanadia and hence with their catalytic
activity is very important.
vanadium loading. This suggests that a sub-monolayer coverage TPD of ammonia was used in this study to compare the total
with easily reducible species is formed at higher loading. Thus, it acidity and acidity strength of the bare c-Al2O3 support and the
is speculated that at higher vanadium loadings, there is a higher VOx/c-Al2O3 catalyst samples with different vanadium loadings.
abundance of polymeric VOx species which is considered more Fig. 5 shows the relationship between the thermal conductivity
reducible than the monomeric VOx that exists at lower vanadium of the desorbed species (NH3 pre-adsorbed at 100 °C) with a linear
loadings. This was further confirmed by determining the degree increase in temperature (15 °C/min).
of reduction (% reduction) of each catalyst sample. Table 3 reports ammonia uptake by various samples and their
The degree of reduction of a supported vanadium oxide catalyst desorption temperatures (Tdes.). The total acidity of each catalyst
can be defined as the ratio of the exposed reducible vanadium sample was established on the basis of integration of desorbed
oxide to the actual vanadium oxide amount on the catalyst. This NH3 as given by the TCD (temperature conductivity detector) pro-
degree of reduction can be calculated for each sample by knowing: files. One should note that the NH3-TPD profile for bare c-Al2O
(a) the amount of reducible metal in the catalyst sample, which can showed a broad initial desorption peak at 170 °C followed by a
be related to the amount of hydrogen consumed during the TPR high temperature desorption peak at 350 °C. These two peaks can
experiments, (b) the actual vanadium amount on the catalyst, (c) be attributed to NH3 desorption from weak and strong Lewis acid
the initial reducible V2O5 species present on the supported catalyst sites, respectively. It can also be seen in Fig. 5 that the addition
(Reaction 1). of vanadium on c-Al2O3 resulted in acidity changes: from
Table 2 reports the hydrogen consumption, the maximum c-Al2O3 related Lewis sites to vanadium oxide promoted Brönsted
reduction temperature and the reduction degree for each catalyst. sites. As a result, this leads to a decreased acid strength, yielding a
It can be observed in Table 2 that the degree of catalyst reduction lower density of weaker acid sites. Moreover, for the vanadium-
increases with vanadium loading. These results support the containing samples, it was observed that raising the vanadium
assumption that there is increased catalyst reducibility under loading resulted in increased acidity of catalyst samples. This
hydrogen atmosphere for catalyst samples with higher vanadium was confirmed by the increased NH3 uptake at higher vanadium
loading. species loading.
The acidity of supported-vanadia catalysts has been the subject
V2 O5 þ 2H2 ! V2 O3 þ 2H2 O ð5Þ
of several studies [88–93]. Unsupported bulk V2O5 powder was
Furthermore, in an attempt to determine the nature of vana- found by Busca et al. [91] to show both surface Brönsted and Lewis
dium oxide surface species, the average oxidation state of vana- acid sites. Moreover, it was also observed by Datka et al. [94] and
dium was calculated using the amount of hydrogen consumed Miyata et al. [95] that the formation of VOx surface species on oxide
via TPR. This allowed us to qualitatively establish the nature of supports such as Al2O3 and TiO2 is accompanied by a decrease in
existing phases by comparing their average oxidation states. These the density of surface Lewis acid sites and an increase in the
average oxidation states were compared with those expected as density of surface Brönsted acid sites.

Table 2
TPR data comparing hydrogen consumption using VOx/c-Al2O3 catalysts with varying amounts of vanadium (wt.%).

Sample Tmax (°C) H2 uptake (cm3 STP/g) % Reductiona AOSb VOx surface species
T1 T2
5% V/c-Al2O3 451.8 – 16.42 81.28 3.37 V2O3.37
7% V/c-Al2O3 445.7 – 22.74 83.06 3.34 V2O3.34
10% V/c-Al2O3 432.6 575 37.03 99.18 3.02 V2O3.02
Average oxidation state of V V2 O5 ! V6 O13 ! VO2 ! V4 O7 ! V2 O3
þ5 þ4:33 þ4 þ3:5 þ3

a
Based on V2O5 + 2H2 ? V2O3 + 2H2O.
b
Average Oxidation State (AOS).
S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140 127

0.09 1490 Py-Bronsted


Py-Lewis
10%VO x/Al2O3 1540 1610 1640
1448
0.08 Old 10% V/c-Al 2O3

0.07
7%VO x/Al2O3
0.06 10% V/γ-Al2O3

Absorbance
TCD (a.u.)

0.05
5%VO x/Al2O3 Fresh γ-Al2O3

0.04
7% V/γ-Al2O3

0.03

0.02 5% V/γ-Al2O3

0.01
γ-Al2O3
0.00
100 150 200 250 300 350 400 450 500 550 600 1350 1400 1450 1500 1550 1600 1650 1700
Temperature (οC) Wavenumber (cm-1)

Fig. 5. NH3-temperature programed desorption profiles as measured with a TCD for Fig. 6. FTIR spectra of pyridine adsorbed on c-Al2O3 and VOx/c-Al2O3 samples.
fresh c-Al2O3 and various VOx/c-Al2O3 catalyst samples. (Sample weight: 0.1 g;
heating rate: 15 °C/min, He flow rate: 30 ml/min, NH3 pre-adsorbed at 100 °C.)

on surface Brønsted acid sites showing that the VOx/c-Al2O3 sam-


Table 3 ples possess both Lewis and Brønsted acid properties. It can also
Temperature programed desorption of NH3 for VOx/c-Al2O3 catalyst samples and bare be observed that the relative intensity of the bands at 1540 cm1
c-Al2O3 support. augments with increasing vanadium loading. This is in agreement
Catalyst sample Tdes. (°C) NH3 uptake (cm3 STP/g) with the results obtained by NH3-TPD where the total acidity of the
VOx/c-Al2O3 samples increases with vanadium loading.
c-Al2O3 200, 350 8.26
5% V/c-Al2O3 194 10.94
Based on these findings, it can be concluded that the total acid-
7% V/c-Al2O3 180 11.50 ity and the number of surface Brønsted acid sites augment with
10% V/c-Al2O3 196 13.02 increasing surface VOx coverage. These results are consistent with
findings of other studies [32,57,73]. In addition, the multi-bands in
the spectral region between 1580 and 1660 cm1 are generally as-
Zou et al. [88] studied the surface acidity of SiO2, c-Al2O3 and signed to hydrogen-bonded pyridine while the bands at 1490 cm1
TiO2 supported vanadia catalysts. This was done using both micro- relate to both Lewis and Brønsted acid species [96,97].
calorimetry and infrared spectroscopy as well as using ammonia as Furthermore, the 1540 cm1 and 1448 cm1 peaks were used to
a probe molecule. The acidity in terms of nature, number and quantify the Brønsted and Lewis acid sites concentrations, respec-
strength was correlated to surface structures of vanadia species tively. Table 4 reports the Brønsted/Lewis acid sites ratios for all
on the catalyst surface. It was found that surface acidity of sup- VOx/c-Al2O3 samples. It can be noticed that Brønsted acid sites
ported vanadia catalysts is strongly affected by the vanadia species are dominant over Lewis acid sites. This was, in fact, true for all
dispersion, surface structure, nature of the support and vanadium samples, with the Brønsted/Lewis ratio increasing steadily with
loading. In addition, microcalorimetric studies of NH3 adsorption vanadium loading. Such an increase in Brønsted acidic sites of
showed a gradual decrease in the heat of ammonia adsorption VOx-based catalysts with vanadium loading has been reported in
and ammonia coverage with the addition of VOx on the c-Al2O3 several studies. This is an important factor to improve VOx catalytic
support. The heat of ammonia adsorption appears to decrease pro- activity for ODH of light alkanes [1,32,49,57,71,73].
gressively with the increase of the vanadium loading up to the Fig. 6 and Table 4, both report data for the 10%VOx/c-Al2O3 cat-
point when the monolayer coverage is complete. This means that alyst sample prepared in our previous study for ethane ODH [44]. It
VOx species may cover some of the strong Lewis acid sites of the can be noticed that at the same V-loading (10%), the less-acidic
c-Al2O3 support. Thus, as vanadium loading increases, this leads c-Al2O3 support resulted in a higher Brønsted/Lewis ratio and a
to a higher abundance of Brönsted sites. larger number of Brønsted acid sites. This acidity difference is par-
ticularly noticeable if one compares the properties of the catalyst
of this study with the ones of other catalysts developed by our
3.1.5. Pyridine FTIR research group for ethane ODH [44].
To gain additional information about the acidity of the prepared
VOx/c-Al2O3 catalysts samples, Fourier Transform Infrared Spec- 3.1.6. Metal dispersion
troscopy (FTIR) using pyridine as a probe molecule was used. By In supported metal oxide catalysts, metal dispersion plays an
using this characterization technique, one is able to assess the cat- important role in determining their catalytic activity. There are
alyst’s acidity based on Brønsted or Lewis sites concentrations. several factors which can control the dispersion of active metal
Fig. 6 shows the FTIR spectra of the c-Al2O3 support and all the on the support materials such method of preparation, type of
VOx/c-Al2O3 samples used in the present study. It can be noted in
Fig. 6 that the FTIR spectra of adsorbed pyridine on the bare
Table 4
c-Al2O3 support do not exhibit any bands other than that at Acid properties of VOx/c-Al2O3 catalyst samples using pyridine FTIR.
1448 cm1 which arise from the adsorption of pyridine on surface
Catalyst sample Brønsted/Lewis sites ratio
Lewis acid sites. This shows that the c-Al2O3 support does not con-
tain any Brønsted acid sites strong enough to react with pyridine. 5% V/c-Al2O3 1.8
On the other hand, in addition to the band at 1448 cm1, the FTIR 7% V/c-Al2O3 4.2
10% V/c-Al2O3 5.1
spectra of adsorbed pyridine on the VOx/c-Al2O3 samples showed a )old) 10% V/c-Al2O3 3.2
1540 cm1 band. This band arises from the adsorption of pyridine
128 S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140

support and metal precursor used. The most commonly used 20000
180
method for the determination of metal oxide dispersion of a cata-
lyst is the selective chemisorption of probe gases like oxygen on a 325 993
700
pre-reduced catalyst sample. 15000 448 520 567
235 345 944 1033
O2 chemisorption experiments were conducted by dosing small (e)
amounts (17.2 lmol) of O2 over a pre-reduced catalyst sample.

Intensity (a.u.)
Oxygen doses were repeated on the sample every 60 s until two (d)
10000
consecutive peaks with constant area were recorded in the TCD.
Oxygen uptake after each pulse was calculated from the area dif- (c)

ference between two consecutive peaks. The pulses were termi-


5000 (b)
nated when the areas of two consecutive peaks were the same,
indicating that no more O2 uptake was taking place. O2 uptake
curves obtained at 370 °C, on various VOx/c-Al2O3 catalyst samples, (a)

as a function of O2 pulses, are shown in Fig. 7. 0


Table 5 reports the cumulative oxygen uptake values obtained 200 300 400 500 600 700 800 900 1000 1100

at 370 °C on various VOx/c-Al2O3 catalyst samples as a function Raman shift (cm-1)


of V loading. It is shown that oxygen uptake increases with vana-
Fig. 8. Dehydrated Raman spectra for (a) calcined c-Al2O3, (b) 5% VOx/c-Al2O3, (c)
dium content up to 10 wt.% VOx loading. Therefore, using a stoichi-
7% VOx/c-Al2O3, (d) 10% VOx/c-Al2O3 and (e) bulk V2O5.
ometry of O/V = 1/1, the metal dispersion, defined as the ratio of
oxygen uptake to V content, was estimated for each sample. This
was achieved by utilizing the total number of V atoms present in 2 h. Raman spectra results are shown for bulk V2O5, activated
the sample as well as the number of chemisorbed oxygen atoms. c-Al2O3 support and various VOx/c-Al2O3 catalysts with different
Metal dispersion is found to decrease with the increase in vana- V loadings. Table 6 also provides the positions of the bands associ-
dium loading. Moreover, the metal dispersions as reported in Ta- ated with various VOx species. The narrow 1030–1035 cm1 band
ble 5 also give an active equivalent particle diameter of vanadia is assigned to the stretching mode of the V@O bond in isolated
particles smaller than 4 nm. This vanadia particle diameter was monovanadate surface species. On the other hand, the broad bands
found to be consistent with the X-ray diffraction patterns where in the 700–945 cm1 region are ascribed to the stretching modes of
no diffraction lines above the background were observed. This V@O (945–1030 cm1) and VAOAV (670–945 cm1) in polyvana-
was true for all catalyst samples prepared. date surface species [62,72,98]. The remaining bands, appearing
at 180, 235, 325, 345, 448, 520, 567 and 993 cm1 are assigned
to bulk V2O5 crystals [59,99].
3.1.7. Laser Raman Spectroscopy
Fig. 8 also shows that the c-Al2O3 support does not exhibit any
Fig. 8 reports the Raman spectra obtained at ambient tempera-
Raman bands in the 100–1100 cm1 region due to the ionic charac-
ture after sample thermal treatment at 450 °C under air flow for
ter of the AlAO bonds [32]. Moreover, it can be seen that the struc-
tures of the VOx surface species as detected by Raman
16 spectroscopy are dependent on vanadium loading and hence VOx
5% VO x/g-Al2O3 surface density. Thus, as the vanadium loading increases from
14 7% VO x/g-Al2O3 5 wt.% to 10 wt.%, the proportion of V2O5 or polyvanadate to
Cumulative O2 Uptake (cm3 STP/g)

10% VO x/g-Al2O3 monovanadate species increases. For example, in the catalyst sam-
12
ple with 5% V loading, the only band detected is that for mono-
vanadate species at 1033 cm1. The absence of polyvanadate
10
bands in Raman spectra suggests that VOx species exist as isolated
8 monovanadate structures. On the other hand, Raman bands attrib-
uted to polyvanadate species are detected in the 750–1000 cm1
6 range for the catalyst sample with 7% V loading and higher which
indicates the coexistence of both polyvanadate and monovanadate
4
species on c-Al2O3. At higher loadings of 10% V, Raman bands cor-
responding to crystalline V2O5 are also detected.
2
In summary, the Raman spectra results presented in Fig. 8 and
0 Table 6 suggest that isolated tetrahedral monovanadate structures
1 3 5 7 9 11 13 15 17 19 21 23 25 anchored on the support with VAOAA1 bonds are prevalent at low
O2 Pulse No. vanadium loading. As V loading increases, two-dimensional poly-
vanadates form, leading to the formation of VAOAV bonds which
Fig. 7. Cumulative oxygen uptake versus number of oxygen pulses on VOx/c-Al2O3 connect to neighboring VOx species. At higher V loadings, three-
catalyst samples (Tads. = Tred. = 370 °C). dimensional structures form via the reaction of VOx species
exceeding monolayer coverage with polyvanadates structures,
which ultimately crystallize into bulk V2O5 form.
Table 5
Oxygen uptake, metal dispersion and active particle diameter on various VOx/c-Al2O3
catalysts.
3.1.8. X-ray diffraction
Sample O2 uptake (cm3/g) % Dispersiona (O/V) Active particle Fig. 9 describes the X-ray diffraction patterns of bulk V2O5, the
diameter (nm)
c-Al2O3 support and various VOx/c-Al2O3 catalysts of various VOx
5% V/c-Al2O3 6.96 63.26 1.97 loadings calcined at 600 °C. No diffraction lines corresponding
7% V/c-Al2O3 9.98 43.25 1.99
to the vanadium oxide compounds were detected in the
10% V/c-Al2O3 13.16 59.81 2.09
VOx/c-Al2O3 samples except in those for the bare c-Al2O3 support
a
Dispersion = fraction of vanadium atoms at the surface, assuming Oads/Vsurf = 1. at 2h positions of 48° and 67°. This may be an indication to either
S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140 129

Table 6
Raman band assignments for vanadium oxide species (Sameer: table was expanded to see all numbers clearly).

Sample Monovanadate (V@O) Polyvanadates (V@O, VAOAV and VAOAAl) V2O5 Support
Bulk V2O5 None None 180, 235, 325, 345, 448, 520, 567 and 993 –
c-Al2O3 (activated) – – – None
5% V/c-Al2O3 1033 None None –
7% V/c-Al2O3 1033 650–945 None –
10% V/c-Al2O3 1033 650–945 180, 235, 325, 448, 520, 567 and 993 –

that VOx species exist as a dispersed amorphous phase on the alu-


300
mina surface or that V2O5 if present, exist as small XRD undetect-
able crystalline nanoparticles (<4 nm). Based on Laser Raman
250
Spectroscopy (LRS) and H2-TPR results reported previously in Sec-

+
tion 3.1.7 of this manuscript, the samples with 5% and 7% V loading
showed only an amorphous VOx phase. However, the sample with 200

+
Intensity (a.u.)
10% V loading showed some crystalline V2O5 nanoparticles coexis- Spent 10%VO x/ γ−Al2O3
ting with the amorphous VOx phase. All these phases were in the 150
sub-monolayer coverage as indicated by the O2-chemisorption dis-
persion results. These findings are in agreement with the published
100 Spent 7%VO x/ γ−Al2O3
data with similar VOx loadings on c-Al2O3, where VOx primarily
present as vanadate or polyvanadate which is known to be X-ray
amorphous. Some crystalline V2O5 nanoparticles could exist at 50 Spent 5%VO x/ γ−Al2O3
lower V loadings which might form due to catalyst preparation
procedures [32,65,85]. Moreover, no evidence of AlV3O9 was de- 0
tected in any of the samples, suggesting that the reaction between 10 20 30 40 50 60 70 80 90
VOx and c-Al2O3 is negligible during the treatment at 600 °C. Intensity (a.u.)
Furthermore, in order to establish the structural stability of the
Fig. 10. XRD patterns of Spent VOx/c-Al2O3 catalysts with different vanadium
VOx/c-Al2O3 catalysts, XRD were considered for spent catalysts
loadings after 10 consecutive propane injections. (Operation conditions: T = 550 °C,
after 10 consecutive propane injections for propane ODH reactions contact time = 20 s, C3H8 injected = 10 ml, catalyst loaded = 0.76 g.).
at the most severe reaction conditions considered in this study (i.e.
20 s contact time and 550 °C). Fig. 10 reports the XRD patterns of Catalytic propane ODH runs were studied under an oxygen-free
the spent VOx/c-Al2O3 catalysts. One can notice the absence of atmosphere employing lattice oxygen of vanadium oxide in the
any diffraction patterns due to structural changes other than that VOx/c-Al2O3 catalysts. To achieve this, two sets of experiments
for the bare c-Al2O3 support at 2h positions of 48° and 67°. This were considered: single-injection and multi-injections experi-
can be viewed as an indication of high catalyst stability under ments. Single-injection experiments are used to study the interac-
the selected reaction conditions. tion of propane with the fully oxidized catalyst in a reaction/
regeneration cycles. In contrast, multi-injection experiments are
used to change the catalyst state from completely oxidized to a
3.2. Propane ODH in the CREC Riser Simulator partially reduced one and to study the influence of the reduction
degree of the catalyst on its performance in propane ODH in con-
Propane ODH experiments were developed in a CREC Riser secutive catalyst reduction cycles. In each set of experiments, the
Simulator operated as a batch reactor mode under fluidized bed only identifiable carbon-containing products other than propylene
conditions. The fluidizability of the various VOx/c-Al2O3 catalyst were CO, CO2, CH4, C2H6 and C2H4.
samples used in this study was ensured in the riser basket by con-
ducting the experiments at high impeller speeds of 5500 rpm. 3.2.1. Thermal runs
Blank thermal runs (without catalyst) were performed using the
empty reactor prior to the catalytic studies to determine the possi-
350 ble contribution of homogenous gas-phase reactions. This allowed
one to clearly distinguish between the catalytic oxidative dehydro-
*

*
*

300 genation effects versus the thermal conversion effect. The thermal
*
*
*

Bulk V 2 O 5 runs were tested using the same reactant (propane) at a tempera-
*
*
*

*
*

250
ture range of 475–550 °C and 20 s contact time which were consid-
Intensity (a.u.)

200 10% VO x / γ− Al 2 O 3 ered as the most severe conditions in the experimental plan. Fig. 11
shows propane conversion in the reactor without any catalyst at
150 the temperature range used in the catalytic runs. It can be seen
7% VO x / γ− Al 2 O 3
in Fig. 11 that propane conversion due to thermal effect displayed
100 low values between 0.21% and 3.12%. Based on these results, ther-
5% VO x / γ− Al 2 O 3
mal cracking was neglected under the conditions studied, and the
+

50
+

γ− Al 2 O 3
conversion observed during the catalytic runs truly represents the
0 catalytic activity of VOx/c-Al2O3 catalysts.
10 20 30 40 50 60 70 80 90

3.2.2. Single-injection propane ODH experiments
Fig. 9. XRD patterns for Al2O3 support and VOx/Al2O3 catalysts with different Single-injection ODH experiments with cycles of reaction
vanadium loadings (*: V2O5, +: Al2O3). followed by regeneration were developed. Thus, in these
130 S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140

3.5 of the interaction with lattice oxygen of the VOx surface species.
Furthermore, low propylene and high COx selectivities at any given
3.0 propane conversion were obtained during single injection ODH
experiments as shown in Fig. 12. These low propylene and high
Propane Conversion (%)

2.5 COx selectivities can be related to both the amount and type of sur-
face oxygen species available on the fresh/regenerated catalyst sur-
2.0 face. An excessive amount of non-stoichiometric, available lattice
oxygen as well as the existence of non-selective oxygen species
1.5 on the catalyst surface is believed to favor the conversion of pro-
pane and the produced propylene into combustion products.
1.0 Fig. 12 shows the effect of an excess amount (non-stoichiome-
tric) of lattice oxygen where propane conversion augments with
0.5 VOx loading on the catalyst. This is accompanied with a parallel in-
crease in COx selectivity and a decrease in propylene selectivity.
0.0 This finding leads us to conclude that an optimum amount of
450 475 500 525 550 575
VOx loading is needed to have the stoichiometric amount of lattice
Temperature (oC) oxygen necessary to selectively convert propane to propylene dur-
Fig. 11. Propane conversion as a function of temperature for the blank runs. ing ODH. On the other hand, there exist non-selective oxygen spe-
(Operating conditions: reaction time = 20 s, C3H8 injected = 10 ml). Error bars cies on the catalyst surfaces that could be either loosely bound
correspond to standard deviation of three repeats. lattice oxygen from the surface and bulk of the catalyst or weakly
adsorbed oxygen species produced from gas-phase O2 during the
experiments, the catalyst was repeatedly reduced by reaction with pre-treatment/regeneration steps. This loosely bound lattice oxy-
propane and then re-oxidized with air at various reaction temper- gen is considered more reactive, and thus, likely to cause car-
atures and contact times. Between the reduction and oxidation, the bon–carbon bond breakage and promote total oxidation. As a
catalyst was flushed with pure Argon flow for 15 min. For each result, both types of oxygen species may participate in the total
reaction cycle, the conversion and selectivity for the main products oxidation of propane and contribute to low propylene selectivity.
were obtained. At the beginning of each experiment, the fresh cat- Fig. 13 describes the selectivity of ODH products over various
alyst was first heated to the reaction temperature under inert catalysts as a function of propane conversion at different reaction
(Argon) gas flow and then pretreated with air flow for 20 min to temperatures. One can notice that there is a consistent decrease
ensure that the catalyst was in a fully oxidized state. in propylene selectivity with a corresponding increase in COx selec-
Fig. 12 reports propane ODH over repeated reaction–regenera- tivity at high conversions. This shows that propylene is the primary
tion cycles at 550 °C using the various VOx/c-Al2O3 catalysts of reaction product of the propane ODH while COx (CO and CO2) are
the present study. Results show that propane conversions and the secondary products of the consecutive oxidation of propane
product selectivity remain stable during the different runs with and propylene. Based on these results, a classical ‘‘triangular’’
the following trends being observed: reaction network with ‘‘direct alkane combustion’’ together with
alkene formation and later ‘‘alkene combustion’’ could be envi-
(a) Propane conversion stays at the 60%, 73.2% and 80.1% levels sioned for propane ODH. Moreover, non-zero selectivities of COx
over the 5%VOx/c-Al2O3, 7%VOx/c-Al2O3 and 10%VOx/c-Al2O3 and C3H6 can be expected upon extrapolation to the zero degree
catalysts samples, respectively. of propane conversion. The non-zero selectivities are an indication
(b) Propylene selectivities are in the 2.1%, 0.93% and 0.57% levels that all reaction products are formed from propane via a competi-
over the 5%VOx/c-Al2O3, 7%VOx/c-Al2O3 and 10%VOx/c-Al2O3 tive (parallel) reaction network.
catalysts samples, respectively. The interaction of propane feed with the oxygen species on the
(c) COx remaining stable at 88%, 90% and 93% levels over the catalyst during ODH reactions in the absence of gas phase oxygen
5%VOx/c-Al2O3, 7%VOx/c-Al2O3 and 10%VOx/c-Al2O3 cata- have been the subject of several studies in the literature [28,100–
lysts samples, respectively. 104]. It has often been suggested for oxidative dehydrogenation
reactions that the selectivity is dependent on the binding strength
As gas phase oxygen was not present in the propane injection, it of the oxygen. Moreover, it has also been proposed that there may
can be concluded that all of the reaction products were the result be different types of oxygen species on the catalyst surface,

90 5 100

80
4 90
COx Selectivity (%)
C3H8 Conversion (%)

70
(a) (b) COx
C3H6 Selectivity (%)

3 80

60 C3H6
2 70
50
5% VOx /Al 2O3
7% VOx /Al 2O3 1 60
40
10% VOx /Al 2O3

30 0 50
0 1 2 3 4 5 6 7 8 9 10 11 0 1 2 3 4 5 6 7 8 9 10 11
Cycle # Cycle #

Fig. 12. (a) Propane conversion and (b) propylene and COx selectivities during consecutive ODH cycles over the various VOx/c-Al2O3 catalysts. (Operating conditions: catalyst
regenerated after every injection, T = 550 °C, C3H8 injected = 10 ml, reaction time = 20 s, catalyst loaded = 0.76 g). Error bars correspond to standard deviation of three repeats.
S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140 131

20 100 10 100
(a) (b)
90

Propylene Selectivity (%)


8
Propylene Selectivity (%)

COx Selectivity (%)


15

COx Selectivity (%)


80 95
6
10 70

4
60 90
5
2
50

0 40 0 85
0 10 20 30 40 50 60 0 10 20 30 40 50 60 70 80
Propane Conversion (%) Propane Conversion (%)

6 100

5
(c)
Propylene Selectivity (%)

COx Selectivity (%)


4 475 oC
500 oC
525 oC
3 95 550 oC

0 90
0 10 20 30 40 50 60 70 80
Propane Conversion (%)

Fig. 13. C3H6 selectivity as a function of C3H8 conversion at various temperatures for single-injection propane ODH over (a) 5% VOx/c-Al2O3 (b) 7% VOx/c-Al2O3 and (c) 10%
VOx/c-Al2O3 catalyst samples. (Operating conditions: C3H8 injected = 10 ml, catalyst loading = 0.76 g). Error bars correspond to standard deviation of three repeats.

including some that are reactive enough to cause the direct and to- In view of the above, it can be concluded that fully oxidized
tal oxidation of propane. Haber and Witko [100] have proposed (fresh) VOx/c-Al2O3 catalysts are active but not selective for pro-
two type of active oxygen surface species: electrophilic oxygen pane ODH reactions. This could be attributed to the availability
forms (e.g. O2, O 
2 , O ) which have been claimed to perform total of an excess amount (non-stoichiometric) of lattice oxygen as well
oxidation, and nucleophilic lattice oxygen ions O2 which have as the existence of non-selective oxygen species on the catalyst
been claimed to be the more selective toward propylene. More- surfaces. Both types of oxygen species participate in the total oxi-
over, Che et al. and Libre et al. [105–107] have found that oxygen dation of propane and propylene product. Thus, an optimized and
species with different bond strength and electronic character could controlled character (nucleophilic or electrophilic) and an appro-
exist on the catalyst surface. As a result, different types of oxygen priate number of oxygen species on the catalyst surface are impor-
species could form on the catalyst surface resulting in different tant factors to be considered for more selective catalysts.
selectivities toward the ODH products. To elucidate this matter, a second set of experiments was con-
In addition to the nucleophilic and electrophilic natures of dif- sidered in the context of the present study. These experiments
ferent surface oxygen species, Kung and Andersen [108] studied were performed with partially reduced catalysts using a series of
the oxidation of butane over two V/c-Al2O3 catalysts with loadings consecutive propane injections. These experiments were devel-
of 2.9 and 8.2 V/nm2 using both pulses and steady butane feeds. oped for propane ODH reaction with no catalyst regeneration in
These authors found that the number of the oxygen atoms on the between injections. In this manner, the effect of the degree of
catalyst surface and the place where these atoms are incorporated reduction (i.e. the number of available active oxygen species) on
in the reactant molecule are two important factors for determining the activity and selectivity of the VOx/c-Al2O3 catalysts in ODH
the selectivity of these catalysts for butane ODH. Thus, it is impor- reaction was examined very effectively. More details about consec-
tant to control the number of oxygen atoms that are incorporated utive-injections propane ODH experiments are given in the follow-
into the reactant molecule. This number is normally controlled via ing section of the present study.
the residence time of the molecule on the surface, the number of
oxygen atoms available at the active sites and the reactivity of oxy- 3.2.3. Consecutive-injections propane ODH experiments
gen on the surface (i.e. its electrophilicity or nucleophilicity). A As stated above, as an alternative, catalyst samples were subject
similar conclusion on the effect of the number of active oxygen to successive propane injections in the CREC Riser Simulator. In
species was made by Dinse et al. [109] in his study to investigate these ODH experiments, the VOx/c-Al2O3 catalysts are progres-
the oxidative dehydrogenation (ODH) of ethane on alumina-sup- sively reduced via the consecutive propane injections without cat-
ported vanadia. In this study, transient-response experiments were alyst regeneration between the injections. With these data and for
carried out with both a fully oxidized sample of 10 wt.% VOx/c- each of the injections, the instantaneous conversion and selectivi-
Al2O3 and a sample that had been partially reduced in H2 to remove ties for the main products were calculated.
half of the reducible oxygen from the vanadia. The rate of ODH was Every experimental set consisted of 10 successive propane
found to depend on the amount of reactive oxygen available on the injections, contact times between 5 and 20 s and temperatures
catalyst surface. between 475 and 550 °C. In order to minimize the effect of the
132 S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140

catalyst to feed ratio, all the experiments were conducted using the decreases as the degree of reduction of the catalyst augments with
same amount of catalyst (0.76 g) and injecting the same amount of the number of consecutive propane injections. This significant
propane (10 ml). The degree of reduction of the catalyst and its change in selectivity indicates that a certain degree of catalyst
selectivities for ODH products was determined by analyzing the reduction is needed in order to obtain good propylene selectivity.
various products resulting from each propane injection. For example, high propylene selectivities of 92.1%, 84.9% and
Fig. 14 reports propane conversion and product selectivities as 77.6% were obtained using the 5%VOx/c-Al2O3, 7%VOx/c-Al2O3
functions of the number of propane injections (pulses) over differ- and 10%VOx/c-Al2O3 catalysts, at conditions where these catalysts
ent VOx/c-Al2O3 catalysts at 550 °C. It can be observed in Fig. 14a were reduced up to 83.2%, 67.4% and 55.06%, respectively.
that the conversion of propane is highest in the first injection for Thus, the above reported results support the view that propyl-
all catalysts and then decreases progressively with increasing ene selectivity in propane ODH reaction over VOx-based catalysts
number of propane injections. However, a small but noticeable is strongly influenced by the binding energy of lattice oxygen.
activity level was maintained even after 10 pulses of propane over When the catalyst is in a more oxidized state, lattice oxygen is
the different VOx/c-Al2O3 catalysts. This could be ascribed to the loosely bound and it is more likely to promote deep oxidation of
availability of an extended supply of lattice oxygen as will be dis- hydrocarbons to carbon oxides. Moreover, there is also the possi-
cussed in more details later in this section. As no molecular oxygen bility that in the first injection, adsorbed oxygen from the regener-
is fed with the injection, the catalyst is believed to supply oxygen ation step (which has a mainly non-selective contribution) could
for the ODH reaction. Following the first propane injection, pro- remain on the VOx/c-Al2O3 catalyst surface. In this case, a selective
pane conversion declined drastically from 58.8%, 76.5% and catalyst surface would be obtained only after the adsorbed oxygen
81.98% to 12.5%, 14.4% and 16.4% over the 5%VOx/c-Al2O3, had been consumed via the first propane injection. This is quite
7%VOx/c-Al2O3 and 10%VOx/c-Al2O3 catalysts samples, respec- apparent for the subsequent propane injections with the propylene
tively. There was a further gradual decrease to 11.5%, 12.4% and selectivity augmenting with the number of propane injections. It is
13.6% with the subsequent propane injections. Thus, it is plausible under these conditions that combustion products (COx) decrease.
that the decrease in propane conversion over VOx/Al2O3 catalysts is Thus, these findings show that in propane ODH reaction network
mainly the result of progressive consumption of the reactive lattice with the catalyst not being regenerated in between propane injec-
oxygen species. tions (oxygen-free environment), propane conversion could be
Fig. 14b also shows the relative variation of propylene and COx achieved with high propylene selectively. This high selectivity for
selectivities. It can be seen that for all catalysts the first injection propylene is the result of the lattice oxygen being the main driver
has very low propylene selectivity and very high COx selectivity. of the ODH reaction. One can also notice that following the first
However, the selectivity for propylene increases and that for COx ethane injection (first pulse), the catalyst tested displays a stable

100 100

90 (a) (b)
80 80
70
C3H8 Conversion (%)

Selectivity (%)

60 60 C3H6
50

40

15

20 COx

10 0
0 1 2 3 4 5 6 7 8 9 10 11 0 1 2 3 4 5 6 7 8 9 10 11
C3H8 Injection # C3H8 Injection #

0.9

0.8
(c)
% Reduction of Catalyst

0.7

0.6

0.5 5% VO x /Al2 O3
7% VO x /Al2 O3
0.4 10% VO x /Al2 O3

0.3

0.2

0.1

0.0
0 1 2 3 4 5 6 7 8 9 10 11
C3H8 Injection #

Fig. 14. ODH propane conversion with successive propane injections (a) propylene and COx selectivity and (b) catalyst degree of reduction as a function of the successive
propane injections ODH over various VOx/c-Al2O3 catalyst samples. (Operating conditions: T = 550 °C, reaction time = 20 s, C3H8 injected = 10 ml, catalyst loaded = 0.76 g).
Error bars correspond to standard deviation of three repeats.
S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140 133

performance in terms of propane conversion and propylene selec- 1.0


5% VO x/Al2O3 Catalyst
tivity even after ten successive reduction runs. 0.9
The effect of the catalyst oxidation state on oxidative dehydro-
0.8
genation selectivity over vanadium based catalysts has been the

Reduction degree of Catalyst


subject of several studies in the literature [43,44,49,104,110– 0.7
112]. In these studies, the ODH reaction was carried out in the ab-
0.6
sence of gas-phase O2 and the catalyst was reduced via successive
pulses of paraffins (e.g. ethane, propane and butane). It was shown 0.5
that olefin selectivities of ODH reactions over various vanadium 0.4
based catalysts augments when increasing the reduction degree
0.3
of the catalyst. Yoon et al. [113], in a study considering propane
475 oC
oxidative dehydrogenation over a MgAMoAO catalyst attributed 0.2 500 oC
this to the progressive catalyst reduction. In this respect, the bind- 525 oC
0.1
ing strength of the remaining lattice oxygen augments and as a re- 550 oC

sult the catalyst becomes more selective. On the other hand, López- 0.0
0 1 2 3 4 5 6 7 8 9 10 11
Nieto et al. [49] and Balcaen et al. [104] found in their studies on
C3H8 Injection #
propane and butane ODH over VOx/c-Al2O3 catalysts that the
reducibility of the V-atoms can influence the selectivity to ODH Fig. 15. Variation of catalyst degree of reduction with reaction temperature as a
products. This effect was assigned to the selective redox rate of function of successive-injection propane ODH over a 5% VOx/c-Al2O3 catalyst.
VOx surface species. In this respect, a higher reaction temperature (Operating conditions: reaction time = 20 s, C3H8 injected = 10 ml, catalyst loa-
ded = 0.76 g). Error bars correspond to standard deviation of three repeats.
was found to accelerate the redox exchanges on the catalyst sur-
face and hence increases the selectivity. Thus, the rate at which
the redox processes take place on the catalyst surface was found as a consequence in a smaller lattice oxygen supply. In a study by
to be another relevant factor in determining the catalytic perfor- Balcaen et al. [114,115], the activation of propane both in the
mance of V-based oxidative dehydrogenation catalysts. In another absence and presence of gas-phase O2 or CO2 over a CuOACeO2/
recent study from our research group, promising results were re- c-Al2O3 catalyst was investigated. These researchers used a TAP
ported by Al-Ghamdi et al. [43,44] on ethane ODH reaction over reactor in the 350–600 °C range. It was found by performing isoto-
VOx/c-Al2O3 catalysts in CREC Riser Simulator in the absence of pic multi-pulse experiments with C18O2 over the 16O2-pretreated
molecular oxygen. It was shown that in the absence of oxygen, eth- catalyst at 600 °C that at higher temperatures there is an increased
ylene could be produced via ODH with 57.684.5% ethylene selec- mobility of lattice oxygen.
tivity at ethane conversions between 6.5% and 27.6% and at To further investigate the extent of lattice oxygen supply by the
temperatures between 550 and 600 °C. It was concluded that the VOx/c-Al2O3 catalysts of the present study, a set of experiments
absence of gas-phase oxygen is critical for the selective conversion were carried out where the catalyst sample with the lowest V load-
of ethane into ethylene with a major role of lattice oxygen in sus- ing (5%) was considered for successive propane injections. This was
taining the ODH over several reaction cycles. done at the most severe experimental conditions (i.e. 20 s contact
The degree of catalyst reduction reported in Fig. 14c is defined time and 550 °C reaction temperature). This experiment was con-
as the ratio of the number of oxygen molecules consumed from tinued until all the lattice oxygen supply was depleted, as shown
the catalyst during the ODH reaction to the total amount of avail- by the full reduction (degree of reduction of 1) reported in
able exchangeable lattice O atoms. The later was determined by Fig. 16(a).
first reducing the VOx/c-Al2O3 catalyst sample in a flow of H2. This Fig. 16(a) reports the propane conversion for 23 consecutive
was then followed by successive pulses of molecular O2 (O2 titra- injections. One can notice a progressive decrease in propane con-
tion) of the reduced sample until no further O2 uptake was moni- version with consecutive injections. In addition, one can observe
tored. Thus, the oxygen consumed during the catalyst re-oxidation that propylene selectivity, as illustrated in Fig. 16(b), showed an
is believed to re-oxidize the surface and replenish the lattice oxy- initial increase from a very low value of 2% to around 91%, remain-
gen. On the other hand, the number of oxygen molecules con- ing stable up to the 14th injection. Following this, propylene selec-
sumed from the catalyst was calculated after each propane tivity showed a gradual decrease to 67.4%. Moreover, this decrease
injection by analyzing the amounts of oxygen-containing products in propylene selectivity after the 14th injection is accompanied by
(i.e. CO, CO2 and H2O). As the ODH reaction was conducted in the a gradual decrease in both COx selectivity and propylene yield as
absence of a gas-phase oxygen, removable lattice oxygen was as- shown in Fig. 16(c). This is an indicator of a decreased supply of lat-
sumed to be the only source of O-atoms in the ODH products. tice oxygen for the ODH reaction. Hence, lower amount of products
It can be also observed in Fig. 14c that different VOx/c-Al2O3 are formed. By the 23rd injection, the catalysts showed a 100% de-
catalysts show different degrees of reduction. The catalyst sample gree of reduction with no further COx formation (0% COx selectiv-
with the highest V loading (10% V) showed a degree of reduction of ity). It is interesting to see that the values of propane conversion
55.1% after 10 propane injections while that with the lowest V and propylene selectivity, at this point, are similar to those ob-
loading (5% V) showed degree of reduction of 80.4% by the 10th tained in thermal runs. At this stage, this is also a strong indicator
injection. This was attributed to the differences in VOx content of of the absence of any lattice oxygen available for the ODH reaction.
the VOx/c-Al2O3 catalysts. In fact higher levels of lattice oxygen Based on these findings, it can be concluded that the lattice oxy-
supply can extend the ODH reaction significantly. gen reservoir on the VOx/c-Al2O3 catalysts makes their use possible
Besides the VOx loading, the extent of the lattice oxygen supply in continuous twin redox fluidized bed reactors. In this type of unit,
can be affected by the reaction temperature. Fig. 15 reports the ex- only a small fraction of the catalyst or about 1/23 would be contin-
tent of reduction for the 5% VOx/c-Al2O3 catalyst sample as a func- uously transferred from the reduction ODH unit (where lattice oxy-
tion of reaction temperature and propane injections. It can be gen is consumed in the ODH reaction) to the oxidizing unit (where
observed that an extended supply of lattice oxygen could be main- lattice oxygen is replenished). Once the catalyst is oxygen replen-
tained by the catalysts at a lower temperature. Thus, as tempera- ished, it should return to the ODH unit.
ture increases, the degree of catalyst reduction per injection Table 7 reports propane conversion and product selectivities for
increases and hence more lattice oxygen is consumed. This results successive-injections propane ODH experiments with pure C3H8
134 S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140

1.0 13.4%, and 15.1% at 550 °C, for the same catalysts samples,
0.9 respectively.
(a) On the other hand, in contrast to the ODH experiments over
0.8
Reduction degree of Catalyst

fully oxidized catalysts (i.e. single-injection ODH experiments),


0.7
propylene selectivity showed a gradual increase with increasing
0.6 reaction temperature, as shown in Fig. 17(b). For example, the cat-
0.5 alyst sample with the 5% vanadium loading gives the highest C3H6
selectivities between 84.4% and 85.9% over the temperature range
0.4
from 475 to 550 °C. For the sample with 7% vanadium loading, the
0.3 C3H6 selectivity increases from 71.2% at 475 °C to 75.3% at 550 °C.
0.2 The C3H6 selectivity of the 10% V catalyst sample is the lowest of
0.1 any of the catalysts examined in this work, with C3H6 selectivity
of 64.1% at 475 °C. Furthermore, it increases to 67.8% at 550 °C. This
0.0
increase in propylene selectivity with reaction temperature is
100
attributed to the variation of catalyst degree of reduction with both
90 (b) reaction temperatures and the number of propane injections.
80 As discussed earlier in this section, the catalyst sample is pro-
Conversion/Slectivity (%)

70 gressively reduced with the successive propane injections. It has


been also shown that the reduction degree of the catalyst is more
60
pronounced at higher temperatures due to higher mobility of the
C 3H 8 conversion
50 lattice oxygen. As a result of this, as the number of propane injec-
C 3H 6 selectivity
40 CO x selectivity tions is increased at higher reactions temperatures, a higher degree
30
of reduction of the catalyst is attained. At such higher degrees of
reduction of the catalysts, the selective pathway toward ODH is
20
preferred over that for combustion and COx formation due to lower
10 availability of lattice oxygen. This is apparent from the observed
0 catalyst performance reported in Fig. 17(c): The increase in propyl-
20
ene selectivity with reaction temperature is accompanied by a
18 decrease in the COx selectivity (CO and CO2).
(c)
16
3.2.3.2. Effect of reaction contact time. The effect of the reaction con-
14
Propylne Yield (%)

tact time was examined by conducting the experiments at four dif-


12 ferent contact times: 5, 10, 15 and 20 s. Fig. 18(a, c and e) reports
10 propane conversions and selectivities to propylene and COx during
8
the ODH reaction runs at various reaction times and 550 °C. It is
observed that propane conversion for all catalysts increases as
6
the reaction time increases giving 4.1%, 5.2% and 6.2% propane con-
4 version at 5 s contact time and increase to 11.7%, 13.4%, and 15.1%
2
at 20 s reaction time for the 5%V, 7%V and 10%V catalysts samples,
respectively.
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 One can notice that propylene selectivity augments slightly
C3H8 Injection # with increasing reaction time as shown in Fig. 18(b, d and f). The
catalyst sample with the 5% vanadium loading shows C3H6 selec-
Fig. 16. Changes of degree of catalyst reduction (a) propane conversion, COx tivities between 70.9% and 76.4% at 5 s contact time and increase
selectivity and propylene selectivity and (b) propylene yield (c) with the number of to 84.4–85.9% at 20 s contact time. For the samples with 7% and
propane injections propane ODH over a 5% VOx/c-Al2O3 catalyst. Operating 10% vanadium loading, the C3H6 selectivity increases from 60.7%
conditions: T = 550 °C, reaction time = 20 s, C3H8 injected = 10 ml, catalyst
to 66.2% and 55.1% to 60.7% at 5 s contact time to 71.2–75.3%
loaded = 0.76 g). Error bars correspond to standard deviation of three repeats.
and 64.13–67.8% at 20 s contact time, respectively. This increase
in propylene selectivity with reaction contact time can be assigned
over various VOx/c-Al2O3 catalysts at various contact times and dif- to the variation of catalyst degree of reduction with contact time is
ferent temperature levels. The only identifiable carbon-containing analogous to the effect of reaction temperature discussed earlier.
products other than propylene were CO, CO2, CH4, C2H6 and C2H4. Longer contact times between the catalyst and reactant at any
given propane injection yield higher consumption of catalyst lat-
tice oxygen, resulting in higher degree of catalyst reduction for
3.2.3.1. Effect of reaction temperature. The effect of the reaction the subsequent injections.
temperature was examined by conducting successive-injection On the other hand, and as a complementary evidence of this
propane ODH experiments over the various VOx/c-Al2O3 catalysts finding, one can observe that the product combustion selectivity
at reaction temperatures between 475 and 550 °C. Fig. 17 reports (COx), shown in Fig. 18(b, d and e), displays a gradual decrease with
propane conversions and selectivities to propylene and COx during increasing reaction time. This demonstrates that limited lattice
the ODH reaction runs at various reaction temperatures. oxygen is available at higher extents of reduction and longer con-
It can be observed that catalyst samples with different vana- tact times. Both conditions favor high propane conversions with
dium loadings showed different reactivity during the successive- high propylene selectivities.
injection propane ODH experiments, resulting in 5.4%, 6.4% and Thus, one can conclude that the performance of the various
7.2% propane conversion at 475 °C for the 5%V, 7%V and 10%V cat- VOx/c-A12O3 catalysts of the present study is affected considerably
alysts samples, respectively. As shown in Fig. 17(a), as the reaction by both ODH operation conditions and catalyst regeneration. It ap-
temperature increases, propane conversion increases to 11.7%, pears that consecutive reaction injections (pulses) are a preferred
S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140 135

Table 7
Conversion and product distribution results for Successive-Injections propane ODH over various VOx/c-Al2O3 catalysts. Reported values are the average of 10 consecutive C3H8
injections.

Catalyst T (°C) Time (s) XC3 H8 (%) Selectivity (%) C3H6 yield (%)
C3H6 COx CH4 C2H6 C2H4
5% VOx/c-Al2O3 475 5 2.35 70.89 86.49 2.76 0.24 0.70 1.63
10 3.29 76.86 89.88 2.97 0.27 0.67 2.50
15 4.19 80.35 91.53 2.97 0.29 0.67 3.35
20 5.38 84.42 94.45 2.94 0.31 0.61 4.53
500 5 3.08 73.37 86.26 4.73 0.60 1.30 2.18
10 4.59 77.53 89.96 4.20 0.69 1.06 3.46
15 5.77 81.89 91.89 4.98 0.94 1.12 4.71
20 7.32 84.70 95.22 4.15 0.89 0.96 6.21
525 5 3.83 75.44 88.01 6.14 1.51 2.26 2.85
10 6.32 79.92 91.09 5.67 2.10 2.03 5.03
15 7.99 83.28 92.90 5.58 2.60 1.99 6.67
20 8.88 85.79 95.52 4.84 2.30 1.75 7.64
550 5 4.08 76.38 89.66 6.30 1.54 3.78 3.13
10 6.08 81.24 91.86 5.49 2.20 2.76 4.98
15 9.12 84.52 93.88 5.05 3.30 2.44 7.73
20 11.73 85.94 96.90 5.41 3.62 1.55 10.11
7% VOx/c-Al2O3 475 5 3.24 60.73 35.81 2.57 0.25 0.64 1.62
10 4.18 67.20 29.30 2.61 0.30 0.59 2.59
15 5.14 67.68 28.76 2.64 0.32 0.61 3.29
20 6.42 71.20 25.19 2.73 0.34 0.55 4.36
500 5 3.74 64.75 29.00 4.45 0.65 1.16 2.14
10 5.60 68.94 24.78 4.29 0.89 1.11 3.53
15 6.89 71.58 22.00 4.37 1.01 1.03 4.69
20 8.11 73.09 21.03 4.02 0.99 0.87 5.82
525 5 4.60 65.29 24.36 6.31 1.76 2.28 2.63
10 6.55 70.60 19.46 5.79 2.20 1.94 4.34
15 9.77 71.98 19.11 5.00 2.31 1.61 6.78
20 11.88 73.79 17.51 4.78 2.40 1.51 8.45
550 5 5.06 66.20 20.01 7.12 2.43 4.24 3.18
10 8.41 71.49 15.64 6.60 2.98 3.30 5.90
15 11.21 74.46 11.67 6.73 3.92 3.22 8.36
20 13.36 75.34 10.43 7.21 4.44 2.58 10.06
10% VOx/c-Al2O3 475 5 3.73 55.12 41.52 2.48 0.27 0.61 1.48
10 4.67 61.99 34.59 2.49 0.33 0.59 2.57
15 5.90 61.67 34.82 2.60 0.34 0.57 3.32
20 7.16 64.13 32.45 2.53 0.36 0.53 4.24
500 5 4.22 57.24 36.74 4.27 0.71 1.04 2.01
10 6.27 63.43 30.05 4.44 1.03 1.06 3.42
15 7.68 64.64 29.27 4.13 1.03 0.92 4.52
20 9.14 65.73 28.66 3.72 1.07 0.82 5.70
525 5 5.26 58.79 30.29 6.60 2.06 2.26 2.47
10 7.24 64.75 25.05 6.04 2.28 1.88 4.10
15 10.79 65.78 25.89 4.85 2.07 1.41 6.56
20 13.41 67.29 23.58 5.21 2.48 1.43 8.32
550 5 6.16 60.64 23.79 7.83 3.18 4.56 3.49
10 10.50 66.18 18.79 7.51 3.68 3.84 6.64
15 12.93 66.81 16.13 8.53 4.75 3.78 8.53
20 15.05 67.77 15.01 8.72 4.99 3.51 10.16

mode of operation for ODH. As a result, regeneration should only medium loading. This occurs until the monolayer coverage is
be allowed once this series of consecutive reaction cycles is com- achieved with 100% polymerized species at monolayer surface
pleted. One can, for instance, envision that this is equivalent to coverage. Thus, the observed activities and selectivities toward dif-
operating a twin circulating fluidized reactor process (reactor– ferent products obtained during propane ODH over the various
regenerator), where the following occurs: (a) most of the catalyst VOx/c-Al2O3 catalysts can be further related to their surface
stream leaving the reactor is returned to the ODH unit directly; density. This would allow the determination of the relative contri-
(b) only a small fraction of the catalyst stream is directed toward bution of isolated and polymeric surface VOx species for this cata-
the regenerator. This operational strategy can allow the function- lytic reaction. In order to do this, the initial propane ODH rate
ing of industrial scale ODH processes with high propylene selectiv- (normalized per gram of catalyst) was compared.
ities, under conditions close to the ones reported in Fig. 14(a and In this respect, one can calculate the initial average reaction rate
b). (per gram of catalyst) of propane ODH using the following
equation:
3.2.3.3. Effect of vanadium loading. VOx/c-Al2O3 catalysts for pro-  
pane ODH have been reported in the technical literature as being Nopropane X propane mol
r0ODH ¼ ð6Þ
affected by both vanadium loading and hence vanadium surface W cat t gs
density [1,32,50,62,73,116–118]. At low vanadium loading (lower
VOx surface density), a highly dispersed amorphous vanadate where N opropane is the number of moles of propane injected in each
phase is formed. Its structure changes from isolated VO4 species run into the CREC Riser Simulator, Xpropane is the conversion of pro-
(monovanadates) to polymeric VO4 species (polyvanadates) at pane, Wcat is the weight of catalyst used and t is the reaction time.
136 S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140

16 90

14
(a) (b)
85
12
C3H8 Conversion (%)

C3H6 Selectivity (%)


80
10

8 75

6
70
4
65
2

0 60
460 470 480 490 500 510 520 530 540 550 560 470 480 490 500 510 520 530 540 550 560
Temperature ( C) ο Temperature (οC)

35

30 (c)
25
COx Selectivity (%)

5% VOx /Al2O3
7% VOx /Al2O3
20 10% VOx /Al2O3

15

10

0
470 480 490 500 510 520 530 540 550 560
Temperature (οC)

Fig. 17. Temperature effect on ODH propane conversion and product distribution: (a) C3H8 conversion, (b) C3H6 selectivity and (c) COx selectivity for successive-injections
propane ODH over various VOx/c-Al2O3 catalyst samples. (Operating conditions reaction time = 20 s, C3H8 injected = 10 ml, catalyst loaded = 0.76 g). Data points are the
averages values of 10 successive injections excluding the first injection on fully oxidized (fresh) catalyst). Error bars correspond to standard deviation of three repeats.

Fig. 19(a) reports initial propane ODH rates as a function of surface density while the initial rate of formation (per gram of cat-
vanadium surface density at different reaction temperatures. As alyst) of COx increased with VOx surface density. Thus, the propyl-
shown in Fig. 19a, for all temperatures, the ODH rate increases ene selectivity changes as shown in Fig. 20 can be traced to the
with increasing VOx surface density. This suggests that polyvana- formation of extra COx at higher VOx surface species concentration.
date surface species, which are more dominant at higher VOx sur- This increase in COx formation with VOx surface density can be
face densities, are more active than monovanadate species. This attributed to the excess surface oxygen species available at higher
type of dependence is expected as the catalysts used in this work loading. These findings provide further evidence of the role as-
are all in the sub-monolayer VOx surface coverage as shown in signed to the polyvanadates species as being more active but less
Fig. 4. It is expected that the initial propane ODH rate will increase selective than monovanadate. With the increased activity of poly-
with VOx surface density. This will occur until monolayer surface vanadate species, more oxygen species are available for the ODH
coverage (8.8 VOx/nm2) is reached. After this, a decrease in the reaction resulting in reduced propylene and increased COx selectiv-
initial propane ODH rate will start to develop with further in- ities as shown in Fig. 20.
creases in VOx surface density. This decrease in the propane ODH The reactivity variation of different types of VOx surface species
rate with VOx surface density above monolayer coverage has been (monovanadate versus Polyvanadates) could also be explained
attributed to the formation of crystalline V2O5 nanoparticles. These from their ease-of-reduction. Early in this section, H2-TPR was used
nanoparticles decrease the number of exposed catalytic active sites to determine the extent of reduction of each catalyst used in this
[48,57]. study. The obtained% reductions were 81.3%, 83.1% and 99.2% for
Furthermore, Fig. 20 describes the influence of VOx surface den- the 5% VOx/c-Al2O3, 7% VOx/c-Al2O3 and 10% VOx/c-Al2O3 catalysts,
sity on product selectivities during propane ODH over VOx/c-Al2O3 respectively. It appears that the V5+ species, and the (VAOAAl)
catalysts. It can be seen that at all reaction temperatures, the sam- bond in the monomeric VOx surface species are more difficult to
ple with the lowest vanadia surface density (3.1 VOx/nm2) and con- reduce than the (VAOAV) or (V@O) bonds in the polymeric VOx
taining predominately monovanadates surface species shows the surface species. This can also be justified considering that propane
highest propylene selectivities. However, increasing vanadia sur- ODH reaction initial reduction rates over the VOx/c-Al2O3 catalyst
face density yields a decrease in propylene and increase in COx involve formation of easy-to-reduce VOx surface species. Previous
selectivities products. studies [49,59,71] have suggested that the higher propane ODH
These results are in well agreement with the initial rates of for- rates observed on supported metal oxide catalysts depend on their
mation of propylene and COx (per gram of catalyst) reported in reducible nature. It was found that for a given metal oxide (e.g.
Fig. 19(b). It is shown in Fig. 19(b) that the initial rate of formation VOx, Mo W, or Nb), propane turnover rates increased as the reduc-
of propylene (per gram of catalyst) was almost independent of VOx tion rate of the oxide catalyst using H2 increased.
S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140 137

90 50

(a) o
475 C
500 oC
80
(b)
40
10 525 oC 70
o
550 C C3H6
C3H8 Conversion (%)

C3H6 Selectivity (%)

COx Selectivity (%)


60
30
50

40 COx
5 20
30

20 10

10

0 0 0
0 5 10 15 20 0 5 10 15 20
Reaction Time (s) Reaction Time (s)

15 80 60

(c) 70 (d) 50
60
C3H6
C3H8 Conversion (%)

C3H6 Selectivity (%)

COx Selectivity (%)


10 40
50

COx
40 30

30
5 20
20
10
10

0 0 0
0 5 10 15 20 0 5 10 15 20
Reaction Time (s) Reaction Time (s)

70
70

15 (e) (f) 60
60
C3H6
C3H8 Conversion (%)

50
C3H6 Selectivity (%)

COx Selectivity (%)


50

10 40
40 COx

30
30

5 20 20

10 10

0 0 0
0 5 10 15 20 0 5 10 15 20
Reaction Time (s) Reaction Time (s)

Fig. 18. Influence of contact time on propane conversion, propylene and COx selectivities over: (a and b) 5%V, (c and d) 7% V and (e and f) 10%V catalyst samples. (Operating
conditions: T = 550 °C, C3H8 injected = 10 ml, catalyst loaded = 0.76 g). Data points are the average values of 10 successive injections excluding the first injection on a fully
oxidized (fresh) catalyst). Error bars correspond to standard deviation of three repeats.

The results obtained in the present study with regard to the VOx species, in the terminal (V@O) bond in the isolated (mono-
reactivity-structure of the VOx/c-Al2O3 catalysts for propane ODH meric) VO4 surface species or in the bridging (VAO-support) bond
are in good agreement with several previously reported studies available at a given VOx surface density on a specific type of
for supported vanadium oxide catalysts in propane ODH support.
[48,119,120]. In all these studies, it was found that the specific pro- In summary, it is proven in the present study that in the
pane reaction rate (expressed in turnover frequency TOF) in ODH absence of gas phase molecular oxygen, propane could be con-
augments with increasing VOx surface density. This indicated that verted with high selectivity toward propylene via the contribution
the polyvanadate phase is more active than the monovanadate of lattice oxygen. Such high propylene selectivity is achieved via
phase. This variation in ODH catalytic activity with VOx surface the design of a catalyst with carefully selected properties such as
density was attributed to the different contributions of lattice oxy- acidity, reducibility and balance of active phases (i.e. polyvanadate
gen atoms in the bridging (VAOAV) bonds in the polymeric surface and monovanadate phases). With this being accomplished and
138 S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140

4.5x10-3 (a) A new VOx/c-Al2O3 catalyst for propane ODH with carefully
475 oC
selected properties such as acidity, reducibility and balance
4.0x10-3 o
500 C of active phases (i.e. polyvanadate and monovanadate
o
525 C
phases) was developed. This catalyst was characterized
Reaction Rate (mol/gcat.s)

o
550 C
3.5x10-3
using BET surface area analysis and showed a moderate
decrease in the total surface area of the catalyst after vana-
3.0x10-3
dium loading on the calcined c-Al2O3 support.
(b) It was shown that the VOx surface structure, reducibility and
2.5x10-3
acidity of the developed catalyst depend on vanadium load-
2.0x10-3
ing. Monomeric VOx surface species were a dominant species
at low vanadium loadings while polymeric VOx species were
1.5x10-3 formed at higher loadings. In addition, it was noticed that
(a) the polymeric surface VOx species were more reducible
1.0x10-3 and displayed a higher abundance of Brønsted acidity than
2 3 4 5 6 7 8 9 the isolated VOx species.
VOx Surface Density (V/nm2) (c) It was proven that the prepared catalyst VOx/c-Al2O3
displayed good propane conversions (11.73–15.11%)
0.6 with promising propylene selectivity (67.65–85.89%) at
475–550 °C. This was demonstrated in a CREC fluidized-bed
0.5
Riser Simulator under consecutive propane injections. It
was found that recently regenerated VOx/c-Al2O3 catalysts
Formation Rate (mol/gcat.s)

were active, however, not selective for propane ODH conver-


0.4
sion. On the other hand, it was proven that the VOx/c-Al2O3
catalysts following the second and subsequent propane injec-
0.3
tions yielded steady high propylene selectivities in ODH. This
promising propylene selectivity was attained under an oxy-
0.2 C3H6 gen free atmosphere with a controlled degree of catalyst
reduction via the contribution of VOx/c-Al2O3 lattice oxygen.
COx
0.1

(b)
0.0 Acknowledgements
2 3 4 5 6 7 8
VOx Surface Density (V/nm2) The authors wish to acknowledge the National Sciences and
Engineering Research Council of Canada (NSERC) for their financial
Fig. 19. Effect of VOx Surface density on: (a) propane ODH (per gram of catalyst) support of this research and the Saudi Arabian Oil Company (Saudi
and (b) propylene and COx formation rates (per g-catalyst) over VOx/c-Al2O3
ARAMCO), Saudi Arabia, for the scholarship awarded for the devel-
catalysts. (Operating conditions: reaction time = 20 s, C3H8 injected = 10 ml, cata-
lyst loaded = 0.76 g). Error bars correspond to standard deviation of three repeats. opment of this Ph.D. research at the University of Western Ontario.

References
90 100
[1] Blasco T, López-Nieto JM. Oxidative dehydrogenation of short chain alkanes
90
85 475 oC on supported vanadium oxide catalysts. Appl Catal A Gen 1997;157:117–42.
500 oC 80 [2] Mamedov EA, Corberan VC. Oxidative dehydrogenation of lower alkanes on
80 525 oC vanadium oxide-based catalysts. The present state of the art and outlooks.
550 oC
C3H6 Selectivity (%)

COx Selectivity (%)

70 Appl Catal A Gen 1995;127:1–40.


75 [3] Baerns M, Buyevskaya O, Mulla SAR. A comparative study on non-catalytic
60
and catalytic oxidative dehydrogenation of ethane to ethylene. Appl Catal A
C3H6
70 50 Gen 2002;226:73–8.
[4] Shen Z, Liu J, Xu H, Yue Y, Hua W, Shen W. Dehydrogenation of ethane to
40 ethylene over a highly efficient Ga2O3/HZSM-5 catalyst in the presence of
65
COx
30 CO2. Appl Catal A Gen 2009;356:148–53.
60 [5] Čapek L, Bulánek R, Adam J, Smoláková L, Sheng-Yang H, Čičmanec P.
20 Oxidative dehydrogenation of ethane over vanadium-based hexagonal
55 mesoporous silica catalysts. Catal Today 2009;141:282–7.
10
[6] Shi X, Ji S, Wang K. Oxidative dehydrogenation of ethane to ethylene with
50 0 carbon dioxide over Cr–Ce/SBA-15 catalysts. Catal Lett 2008;125:331–9.
2 3 4 5 6 7 8 [7] Čapek L, Adam J, Grygar T, Bulánek R, Vradman L, Košová-Kučerová G, et al.
Oxidative dehydrogenation of ethane over vanadium supported on
VOx Surface Density (V/nm2)
mesoporous materials of M41S family. Appl Catal A Gen 2008;342:99–106.
[8] Klose F, Wolff T, Lorenz H, Seidelmorgenstern A, Suchorski Y, Piorkowska M,
Fig. 20. Dependence of propylene and COx selectivities on VOx surface density for et al. Active species on c-alumina-supported vanadia catalysts: nature and
VOx/c-Al2O3 catalysts. reducibility. J Catal 2007;247:176–93.
[9] Danica B, Desislava A, Mirko P, Stefan H. An experimental study of the partial
based on the promising propane ODH results obtained in the pres- oxidation of ethane to ethylene in a shallow fluidized bed reactor. J Serbian
Chem Soc 2007;72:183–92.
ent study in the CREC Fluidized-Bed Riser Simulator, industrial
[10] Haddad N, Bordes Richard E, Hilaire L, Barama A. Oxidative dehydrogenation
scale propane ODH application and scale up can be envisioned. of ethane to ethene on alumina-supported molybdenum-based catalysts
modified by vanadium and phosphorus. Catal Today 2007;126:256–63.
[11] Karamullaoglu G, Dogu T. Oxidative dehydrogenation of ethane over
4. Conclusions chromium–vanadium mixed oxide and chromium oxide catalysts. Ind Eng
Chem Res 2007;46:7079–86.
[12] Tope B, Zhu Y, Lercher JA. Oxidative dehydrogenation of ethane over Dy2O3/
On the basis of the data and results obtained in the present MgO supported LiCl containing eutectic chloride catalysts. Catal Today
study, the following key conclusions can be reported: 2007;123:113–21.
S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140 139

[13] Cavani F, Ballarini N, Cericola A. Oxidative dehydrogenation of ethane and [44] Al-Ghamdi SA, Volpe M, Hossain MM, de Lasa HI. VOx/c-Al2O3 catalyst for
propane: how far from commercial implementation? Catal Today oxidative dehydrogenation of ethane to ethylene: desorption kinetics and
2007;127:113–31. catalytic activity. Appl Catal A Gen 2013;450:120–30.
[14] Lemonidou AA, Heracleous E. Reaction pathways of ethane oxidative and [45] Bañares MA, Khatib SJ. Structure-activity relationships in alumina-supported
non-oxidative dehydrogenation on c-Al2O3 studied by temperature- molybdena–vanadia catalysts for propane oxidative dehydrogenation. Catal
programmed reaction (TP-reaction). Catal Today 2006;112:23–7. Today 2004;96:251–7.
[15] Solsona B, Dejoz A, Garcia T, Concepcion P, López-Nieto JM, Vazquez M, et al. [46] Sugiyama S, Sugimoto N, Hirata Y, Nakagawa K, Sotowa K. Oxidative
Molybdenum–vanadium supported on mesoporous alumina catalysts for the dehydrogenation of propane on vanadate catalysts supported on various
oxidative dehydrogenation of ethane. Catal Today 2006;117:228–33. metal hydroxyaptites. Phosphorus Res Bull 2008;22:13–6.
[16] Heracleous E, Lemonidou AA. Ni–Nb–O mixed oxides as highly active and [47] Ballarini N, Cavani F, Cericola A, Cortelli C, Ferrari ML, Capannelli G, et al.
selective catalysts for ethene production via ethane oxidative Supported vanadium oxide-based catalysts for the oxidehydrogenation of
dehydrogenation. Part II: Mechanistic aspects and kinetic modeling. J Catal propane under cyclic conditions. Catal Today 2004;92:99–104.
2006;237:175–89. [48] Bell AT, Iglesia E, Argyle MD, Chen K. Effect of catalyst structure on oxidative
[17] Heracleous E, Lemonidou AA. Ni–Nb–O mixed oxides as highly active and dehydrogenation of ethane and propane on alumina-supported vanadia. J
selective catalysts for ethene production via ethane oxidative Catal 2002;208:139–49.
dehydrogenation. Part I: Characterization and catalytic performance. J Catal [49] López-Nieto JM, Soler J, Concepcion P, Herguido J, Menendez M, Santamaria J.
2006;237:162–74. Oxidative dehydrogenation of alkanes over V-based catalysts: influence of
[18] Botella P, Dejoz A, López-Nieto JM, Concepcion P, Vazquez M. Selective redox properties on catalytic performance. J Catal 1999;332:324–32.
oxidative dehydrogenation of ethane over MoVSbO mixed oxide catalysts. [50] Lemonidou AA, Nalbandian L, Vasalos IA. Oxidative dehydrogenation of
Appl Catal A Gen 2006;298:16–23. propane over vanadium oxide based catalysts effect of support and alkali
[19] Nakagawa K, Miyake T, Konishi T, Suzuki T. Oxidative dehydrogenation of promoter. Catal Today 2000;61:333–41.
ethane to ethylene over NiO loaded on high surface area MgO. J Mol Catal A [51] Grzybowska B, Klisinska A, Samson K, Gressel I. Effect of additives on
Chem 2006;260:144–51. properties of V2O5/SiO2 and V2O5/MgO catalysts: I. Oxidative
[20] López-Nieto JM, Botella P, García-González E, Dejoz A, Vazquez M, González- dehydrogenation of propane and ethane. Appl Catal A Gen 2006;309:10–6.
Calbet J. Selective oxidative dehydrogenation of ethane on MoVTeNbO mixed [52] Solsona B, López-Nieto JM, Pena ML, Rey F, Vidal-Moya A, Blasco T. Vanadium
metal oxide catalysts. J Catal 2004;225:428–38. oxide supported on mesoporous MCM-41 as selective catalysts in the
[21] Anon R. New propylene production technologies hold great promise. Oil Gas J oxidative dehydrogenation of alkanes. J Catal 2001;203:443–52.
2004;102:50–2. [53] Dinse A, Frank B, Hess C, Habel D, Schomäcker R. Oxidative dehydrogenation
[22] Bhasin M. Is true ethane oxydehydrogenation feasible? Top Catal of propane over low-loaded vanadia catalysts: Impact of the support material
2003;23:145–9. on kinetics and selectivity. J Mol Catal A Chem 2008;289:28–37.
[23] Resasco D, Haller G. Catalytic dehydrogenation of lower alkanes. Catal R Soc [54] Taylor MN, Carley AF, Davies TE, Taylor SH. The oxidative dehydrogenation of
Chem 1994:11. propane using vanadium oxide supported on nanocrystalline ceria. Top Catal
[24] Chan KYG, Inal F, Senkan S. Suppression of coke formation in the steam 2009;52:1660–8.
cracking of alkanes: ethane and propane. Ind Eng Chem Res 1998;37: [55] Routray K, Reddy KRS, Deo G. Oxidative dehydrogenation of propane on V2O5/
901–7. Al2O3 and V2O5/TiO2 catalysts: understanding the effect of support by
[25] Dharia D, Letzsch W, Kim H, McCue D, Chapin L. Increase light olefins parameter estimation. Appl Catal A Gen 2004;265:103–13.
production. Hydrocarb Process 2004;83:61–6. [56] Bell AT, Dinse A, Schomäcker R. The role of lattice oxygen in the oxidative
[26] Ren T, Patel M, Blok K. Olefins from conventional and heavy feedstocks: dehydrogenation of ethane on alumina-supported vanadium oxide. Phys
energy use in steam cracking and alternative processes. Energy Chem Chem Phys 2009;11:6119–24.
2006;31:425–51. [57] Martinez-Huerta MV, Gao X, Tian H, Wachs IE, Fierro JLG, Bañares MA.
[27] Farrauto RJ, Bartholomew CH. Fundamentals of industrial catalytic processes. Oxidative dehydrogenation of ethane to ethylene over alumina-supported
1st ed. London: Chapman & Hall; 1997. vanadium oxide catalysts: relationship between molecular structures and
[28] Kondratenko EV, Steinfeldt N, Baerns M. Transient and steady state chemical reactivity. Catal Today 2006;118:279–87.
investigation of selective and non-selective reaction pathways in the [58] Argyle MD. Oxidative dehydrogenation of light alkanes on metal oxide
oxidative dehydrogenation of propane over supported vanadia catalysts. catalysts. University of California, Berkeley; 2003.
Phys Chem Chem Phys 2006;8:1624–33. [59] Blasco T, Galli A, López-Nieto JM, Trifiro F. Oxidative dehydrogenation of
[29] Chen K, Xie S, Bell AT, Iglesia E. Alkali effects on molybdenum oxide catalysts ethane and n-butane on VOx/Al2O3 catalysts. J Catal 1997;169:203–11.
for the oxidative dehydrogenation of propane. J Catal 2000;195:244–52. [60] Fierro JLG, López-Nieto JM, Kremenic G. Selective oxidation of propene over
[30] Kung H. Oxidative dehydrogenation of light (C2 to C4) alkanes. Adv Catal supported vanadium oxide catalysts. Appl Catal 1990;61:235–51.
1994;40:1–38. [61] Bañares MA. Supported metal oxide and other catalysts for ethane
[31] Chen K, Khodakov A, Yang J, Bell AT, Iglesia E. Isotopic tracer and kinetic conversion: a review. Catal Today 1999;51:319–48.
studies of oxidative dehydrogenation pathways on vanadium oxide catalysts. [62] Wachs IE, Weckhuysen BM. Structure and reactivity of surface vanadium
J Catal 1999;333:325–33. oxide species on oxide supports. Appl Catal A Gen 1997;157:67–90.
[32] Bell AT, Khodakov A, Olthof B, Iglesia E. Structure and catalytic properties of [63] Bozon-Verduraz F, Enache DI, Bordes Richard E, Ensuque A. Vanadium oxide
supported vanadium oxides: support effects on oxidative dehydrogenation catalysts supported on zirconia and titania I. Preparation and
reactions. J Catal 1999;181:205–16. characterization. Appl Catal A Gen 2004;278:93–102.
[33] Chen K, Xie S, Bell AT, Iglesia E. Structure and properties of oxidative [64] Wu Z, Kim H, Stair PC, Rugmini S, Jackson SD. On the structure of vanadium
dehydrogenation catalysts based on MoO3/Al2O3. J Catal 2001;198:232–42. oxide supported on aluminas: UV and visible Raman spectroscopy, UV–
[34] Fukudome K, Ikenaga N, Miyake T, Suzuki T. Oxidative dehydrogenation of visible diffuse reflectance spectroscopy, and temperature-programmed
propane using lattice oxygen of vanadium oxides on silica. Catal Sci Technol reduction studies. J Phys Chem B 2005:2793–800.
2011:987–98. [65] Reddy EP, Varma RS. Preparation, characterization, and activity of Al2O3-
[35] Al-Zahrani SM, Abasaeed AE, Putra M. Kinetics of oxidehydrogenation of supported V2O5 catalysts. J Catal 2004;221:93–101.
propane over alumina-supported Sr–V–Mo catalysts. Catal Commun [66] Sham E, Murgia V, Gottifredi JC, Torres EMF. Oxidative dehydrogenation of
2012;26:98–102. propane and n butane over alumina supported vanadium catalysts. Lat Am
[36] Lin X, Poeppelmeier KR, Weitz E. Oxidative dehydrogenation of ethane with Appl Res 2004;34:75–82.
oxygen catalyzed by K–Y zeolite supported first-row transition metals. Appl [67] Arena F, Frusteri F, Parmaliana A. Structure and dispersion of supported-
Catal A Gen 2012;381:114–20. vanadia catalysts. Influence of the oxide carrier. Appl Catal A Gen
[37] Piumetti M, Bonell B, Massiani P, Millot Y, Dzwigaj S, Gaberova L, et al. Novel 1999;176:189.
vanadium-containing mesocellular foams (V-MCF) obtained by direct [68] Auroux A, Le Bars J, Forissier M, Vedrine JC. Active sites of V2O5-c-Al2O3
synthesis. Microporous Mesoporous Mater 2011;142:45–54. catalysts in the oxidative dehydrogenation of ethane. J Catal 1996;162:250–9.
[38] Piumetti M, Bonell B, Massiani P, Dzwigaj S, Rossetti I, Casale S, et al. Effect of [69] Yoshida S, Tanaka T, Nishima Y, Mizutani H. The local structures of vanadium
vanadium dispersion and of support properties on the catalytic activity of V- oxide on silica and c-alumina studied by X-ray absorption (XANES-EXAFS)
containing silicas. Catal Today 2012;179:140–8. spectroscopy – the effect of hydration. 9th Int congr catal; 1988. p. 1473.
[39] Rubio O, Herguido J, Menendez M. Oxidative dehydrogenation of n-butane on [70] Corma A, López-Nieto JM, Paredes N, Perez M, Shen Y, Cao H, et al. Oxidative
V/MgO catalysts-kinetic study in anaerobic conditions. Chem Eng Sci dehydrogenation of propane over supported-vanadium oxide catalysts. Stud
2003;58:4619–27. Surf Sci Catal 1992:213–20.
[40] Soler J, López-Nieto JM, Herguido J, Menendez M, Santamaria J. Oxidative [71] Chen K, Bell AT, Iglesia E. The relationship between the electronic and redox
dehydrogenation of n-butane in a two-zone fluidized-bed reactor. Ind Eng properties of dispersed metal oxides and their turnover rates in oxidative
Chem Res 1999;38:90–7. dehydrogenation reactions. J Catal 2002;209:35–42.
[41] Rubio O, Herguido J, Menendez M. Oxidative dehydrogenation of butane in an [72] Roozeboom F, Mittelmeijer-Hazeleger MC, Moulijn JA, Medema J, Beer VHJ
interconnected fluidized-bed reactor. AIChE J 2004;50:1510–22. De, Gellings PJ. Vanadium oxide monolayer catalysts. 3. A Raman
[42] Lemonidou AA. Oxidative dehydrogenation of C4 hydrocarbons over VMgO spectroscopic and temperature-programmed reduction study of monolayer
catalyst—kinetic investigations. Appl Catal A Gen 2001;216:277–84. and crystal type vanadia on various supports. J Phys Chem 1980;84:2783–91.
[43] Al-Ghamdi SA, Hossain MM, de Lasa HI. Kinetic modeling of ethane oxidative [73] López-Nieto JM. The selective oxidative activation of light alkanes. From
dehydrogenation over VOx/Al2O3 catalyst in a fluidized-bed riser simulator. supported vanadia to multicomponent bulk V-containing catalysts. Top Catal
Ind Eng Chem Res 2013;52:5235–44. 2006;41:3–15.
140 S.A. Al-Ghamdi, H.I. de Lasa / Fuel 128 (2014) 120–140

[74] Wachs IE. Recent conceptual advances in the catalysis science of mixed metal [98] Wachs IE. Raman and IR studies of surface metal oxide species on oxide
oxide catalytic materials. Catal Today 2005;100:79–94. supports: supported metal oxide catalysts. Catal Today 1996;27:437–55.
[75] Ishida S, Imamura S, Fujimura Y. Acid–base properties of supported vanadium [99] Zhao B, Xu X, Ma H, Sun D, Gao J. Monolayer dispersion of oxides and salts on
oxide catalyst. React Kinet Catal Lett 1991;43:453–9. surface of ZrO2 and its application in preparation of ZrO2-supported catalysts
[76] Blasco T, López-Nieto JM, Dejoz A, Vazquez M. Influence of the acid–base with high surface areas. Catal Lett 1997;45:237–44.
character of supported vanadium catalysts on their catalytic properties for [100] Haber J, Witko M. Quantum-chemical modelling of hydrocarbon oxidation on
the oxidative dehydrogenation of n-butane. J Catal 1995;157:27–282. vanadium-based catalysts. Catal Today 1995;23:311–6.
[77] Shen J, Zou H, Li M. Surface acidity of supported vanadia catalysts. J Therm [101] Arnold EW, Sundaresan S. The role of lattice oxygen in the dynamic behaviour
Anal Calorim 2003;72:209–21. of oxide catalysts. Chem Eng Commun 2007:37–41.
[78] Deo G, Wachs IE. Reactivity of supported vanadium oxide catalysts: the [102] Genser F, Pietrzyk S. Oxidative dehydrogenation of propane on V2O5/TiO2
partial oxidation of methanol. J Catal 1994;146:323–34. catalysts under transient conditions. Chem Eng Sci 1999;54:4315–25.
[79] Kung H, Michalakos PM, Kung MC, Jahan I. Selectivity patterns in alkane [103] Crapanzano S, Babich I, Lefferts L. Selection of mixed conducting oxides for
oxidation over Mg3(VO4)2–MgO, Mg2V2O7, and (VO)2P2O7. J Catal oxidative dehydrogenation of propane with pulse experiments. Appl Catal A
1993;140:226–42. Gen 2011;391:70–7.
[80] de Lasa HI. Canadian patent 1,284,017 (1991). USA patent 5,102,628 (1992); [104] Balcaen V, Sack I, Olea M, Marin GB. Transient kinetic modeling of the
1992. oxidative dehydrogenation of propane over a vanadia-based catalyst in the
[81] Al-Ghamdi SA. Thesis: oxygen-free propane oxidative dehydrogenation over absence of O2. Appl Catal A Gen 2009;371:31–42.
vanadium oxide catalysts: reactivity and kinetic modelling. University of [105] Che M, Tench AJ. Characterization and reactivity of molecular oxygen species
Western Ontario; 2013. on oxide surfaces. Adv Catal 1982:1–148.
[82] Geldart D. Types of gas fluidization. Powder Technol 1973;7:285–92. [106] Che M, Tench AJ. Characterization and reactivity of mononuclear oxygen
[83] Centi G. Nature of active layer in vanadium oxide supported on titanium species on oxide surfaces. Adv Catal 1982;31:77–133.
oxide and control of its reactivity in the selective oxidation and [107] Libre JM, Barbaux Y, Grzybowska B, Bonnelle JP. A surface potential study of
ammoxidation of alkylaromatics. Appl Catal A Gen 1996;147:267–98. adsorbed oxygen species on a Bi2Mo3O12 catalyst. React Kinet Catal Lett
[84] Bosc H, Bert JK, Van Ommen JG, Gellings PJ. Factors influencing the 1982;20:249–54.
temperature-programmed reduction profiles of vanadium pentoxide. J [108] Kung H, Andersen PJ. The effect of oxygen binding energy on the selective
Chem Soc Faraday Trans 1984;80:2479–88. oxidation of butane over V/c-Al2O3. Stud Surf Sci Catal 1993;75:205–17.
[85] Koranne MM, Goodwin JG, Marcelin G. Characterization of silica- and [109] Dinse A, Schomäcker R, Bell AT. The role of lattice oxygen in the oxidative
alumina-supported vanadia catalysts using temperature programmed dehydrogenation of ethane on alumina-supported vanadium oxide. Phys
reduction. J Catal 1994;148:369–77. Chem Chem Phys 2009;11:6119–24.
[86] Muhler M, Besselmann S, Freitag C, Hinrichsen O. Temperature-programmed [110] Owen OS, Kung MC, Kung H. The effect of oxide structure and cation
reduction and oxidation experiments with V2O5/TiO2 catalyst. Phys Chem reduction potential of vanadates on the selective oxidative dehydrogenation
Chem Phys 2001;3:4633–8. of butane and propane. Catal Lett 1992;12:45–50.
[87] Van Vuuren CP, Stander F. The reduction of V2O5. Part II. Reduction by [111] Creaser D, Andersson B, Hudgins RR, Silverston PL. Transient kinetic analysis
hydrogen. Thermochim Acta 1990;165:85–91. of the oxidative dehydrogenation of propane. J Catal 1999;182:264–9.
[88] Zou H, Li M, Shen J, Auroux A. Surface acidity of supported vanadium [112] Creaser D, Andersson B, Hudgins RR, Silverston PL. Transient study of
catalysts. J Therm Anal Calorim 2003;72:209–21. oxidative dehydrogenation of propane. Appl Catal A Gen 1999;187:147–60.
[89] Kantcheva MM, Hadjiivanov KI, Klissurski DG. An IR spectroscopy study of the [113] Yoon YS, Ueda W, Moro-oka Y. Oxidative dehydrogenation of propane over
state and localization of vanadium-oxo species adsorbed on TiO2 (anatase). J magnesium molybdate catalysts. Catal Lett 1995;35:57–64.
Catal 1992;134:299–310. [114] Balcaen V, Roelant R, Poelman H, Poelman D, Marin GB. TAP study on the
[90] Turek AM, Wachs IE. Acidic properties of alumina-supported metal active oxygen species in the total oxidation of propane over a CuO–CeO2/
oxide catalysts: an infrared spectroscopy study. J Phys Chem 1992;96: c-Al2O3 catalyst. Catal Today 2010;157:49–54.
5000–7. [115] Balcaen V, Poelman H, Poelman D, Marin GB. Kinetic modeling of the total
[91] Busca G, Ramis G, Lorenzelli V. FT-IR study of the surface properties of oxidation of propane over Cu- and Ce-based catalysts. J Catal
polycrystalline vanadia. J Mol Catal 1989;50:231–40. 2011;283:75–88.
[92] Khader MM. Surface acidity of V2O5/Al2O3 catalysts: IR and TPD studies. J Mol [116] Mattos ARJ, da Silva San Gil RA, Rocco MLM, Eon JG. Zinc-modified, alumina-
Catal A Chem 1995;104:87–94. supported vanadium oxides as catalysts for propane oxidative
[93] Le Bars J, Vedrine JC, Auroux A. Role of surface acidity on vanadia/silica dehydrogenation. J Mol Catal A Chem 2002;178:229–37.
catalysts used in the oxidative dehydrogenation of ethane. Appl Catal A Gen [117] Heracleous E, Machli M, Lemonidou AA, Vasalos IA. Oxidative
1992;88:179–95. dehydrogenation of ethane and propane over vanadia and molybdena
[94] Datka J, Turek AM, Jheng M, Wachs IE. Acidic properties of supported niobium supported catalysts. J Mol Catal A Chem 2005;232:29–39.
oxide catalysts: an infrared spectroscopy investigation. J Catal 1992;135: [118] Schwarz O, Frank B, Schomäcker R. Characterisation and catalytic testing of
186–99. VOx/Al2O3 catalysts for microstructured reactors. Catal Commun
[95] Miyata H, Fujii K, Ono T. Acidic properties of vanadium oxide on titania. J 2008;9:229–33.
Chem Soc Faraday Trans 1 1988;84:3121–8. [119] Khodakov A, Olthof B, Bell AT, Iglesia E. Structure and catalytic properties of
[96] Tonetto G, Atias J, de Lasa HI. FCC catalysts with different zeolite crystallite supported vanadium oxides: support effects on oxidative dehydrogenation
sizes: acidity, structural properties and reactivity. Appl Catal A Gen reactions. J Catal 1999;181:205–16.
2004;270:9–25. [120] Tian H, Ross EI, Wachs IE. Quantitative determination of the speciation of
[97] Chester AW, Derouane EG. Zeolite characterization and catalysis. 5th ed. New surface vanadium oxides and their catalytic activity. J Phys Chem B
York: Springer; 2009. 2006;110:9593–600.

You might also like