Вы находитесь на странице: 1из 13

RSC Advances

View Article Online


PAPER View Journal | View Issue

Study on the reaction mechanism and kinetics of


CO hydrogenation on a fused Fe–Mn catalyst
Published on 16 October 2015. Downloaded by Tubitak - Ulakbim on 4/2/2019 12:26:32 PM.

Cite this: RSC Adv., 2015, 5, 95287


A. A. Mirzaei,*a E. Rezazadeh,a M. Arsalanfar,*b M. Abdouss,*b M. Fatemia
and M. Sahebia

The kinetics of the CO hydrogenation reaction over a Fe–Mn fused catalyst was investigated in a fixed-bed
micro-reactor under the following conditions: temperatures of 573–603 K, pressures of 1–15 bar, H2/CO
feed ratios of 0.7–3.4 and a space velocity of 4500 h1. A reaction rate equation for the supported Fe–
Mn catalyst was derived on the basis of the Langmuir–Hinshelwood–Hougen–Watson and Eley–Rideal
models. An activation energy of 105  3.7 kJ mol1 was obtained for the best fitted model. In addition,
Received 9th July 2015
Accepted 16th October 2015
the power-law equation model was also evaluated for the experimental data. According to the power-
law model, the activation energy was obtained as 95.5  2.5 kJ mol1. Furthermore, the effect of
DOI: 10.1039/c5ra13427g
temperature on the reaction partial order was investigated with respect to the reactants, using four
www.rsc.org/advances simple power law equations. Characterization of the catalyst was carried out using BET and XRD techniques.

may block the pores and sites of the catalyst, resulting in diffusion
1. Introduction limitations and decreasing the activity of the catalyst.10 Tempera-
Fischer–Tropsch synthesis (FTS) has great potential for the ture plays an important role in the amount of carbon deposition
production of ultraclean transportation fuels, like diesel and jet on the catalyst active sites during exothermic FTS, and this reac-
fuel, from synthesized gas produced from more abundant tion should be performed in a way that maintains the near-
resources such as coal, natural gas and biomass. It has been isothermal conditions inside the catalyst beds.11 Numerous
found that several metals, such as nickel (Ni), cobalt (Co), studies have reported about the FTS kinetics over iron-based
ruthenium (Ru) and iron (Fe), can be activated for the FT reac- catalysts. Most kinetic expressions have been developed empiri-
tion.1 In the high-temperature Fischer–Tropsch (HTFT) process, cally, by tting the data to a simple power-law relationship; it was
the company Sasol used a catalyst prepared from fused iron generally found that the reaction order of hydrogen was positive,
oxides together with chemical and structural promoters.2 The whereas that of carbon monoxide was negative.12 Some researchers
catalyst prepared from fused iron oxides was non-porous, so it derived rate expressions of the reactant consumption based on
obviously had a lower surface area compared to other preparation Langmuir–Hinshelwood–Hougen–Watson- (LHHW) or Eley–
methods.3 However, structural promoters, such as the oxides of Rideal-type mechanisms.13,14 One of the most popular mechanisms
aluminum, magnesium, lanthanum or titanium, were added to for hydrocarbon formation on iron catalysts is the surface carbide
increase the active surface area of the catalyst.4 The most mechanism using CH2 insertion.15–19 In particular, iron-based
important catalysts prepared by this method are the promoted catalysts form stable carbides under the FTS reaction.20,21 The
iron for high-temperature FTS and catalysts for ammonia differences in the rate expressions proposed for the consumption
synthesis.3 For iron-based catalysts, the water-gas shi (WGS) of synthesis gas are mainly because of the effect of adsorbed CO,
reaction can affect the FTS reaction rate by changing the H2, and their products (H2O and CO2) on the catalyst surface.
hydrogen and carbon monoxide partial pressures. The addition Carbon monoxide and water adsorb more strongly on the catalyst
of small amounts of manganese to the catalyst enhanced the surface than H2 and CO2 do.22,23 The most evident consequence is
formation of olenic products.5 Fe–Mn catalysts have attracted the usual assumption that water has a strong inhibiting inuence
much attention due to their high olen selectivity, lower methane on the reaction rate.14 The perceived negative inuence of water on
selectivity and excellent stability.5–9 Under a high reaction the reaction rate was ascribed to the competitive adsorption
temperature, deactivation of the catalyst may occur due to the between water and CO on the catalyst surface. With increasing
deposition of carbon on the catalyst surface. The carbon deposits water concentration in the surface of the catalyst, the fraction of
CO converted to hydrocarbons decreases due to an increase in the
a
Department of Chemistry, Faculty of Sciences, University of Sistan and Baluchestan, WGS reaction. Therefore, water has essentially an indirect effect on
P. O. Box 98135-674, Zahedan, Iran. Fax: +98 541 2447231; Tel: +98 541 2447231. the FTS reaction rate, and increasing the water partial pressure will
E-mail: mirzaei@hamoon.usb.ac.ir reduce the amount of surface carbon, which leads to a decrease in
b
Department of Chemistry, Amirkabir University of Technology, Hafez Ave, Tehran, the rate of hydrocarbon formation.24 The following simple
Iran. E-mail: maryam_galavy@yahoo.com; phdabdouss44@aut.ac.ir

This journal is © The Royal Society of Chemistry 2015 RSC Adv., 2015, 5, 95287–95299 | 95287
View Article Online

RSC Advances Paper

relationships exist between the rates of the FTS reaction and the limitations, the obtained catalyst was crushed and screened to
WGS reaction:25 collect the catalyst particles of 30–70 mesh (210–590 mm).

rWGS ¼ rCO2 (1)


2.2. Catalysts characterization
rFT ¼ rCO  rCO2, (2) 2.2.1. X-ray diffraction (XRD). Powder XRD measurements
were performed using a D5000 X-ray diffractometer (Siemens,
where rCO2 is the rate of CO2 formation and rCO is the rate of CO Germany).
consumption. Dry reported that the CO2 inhibition is not as Scans were taken with a 2q step size of 0.02 and a counting
strong as the water inhibition, due to the large difference in time of 1.0 s, using a CuKa radiation source (l ¼ 0.15406 nm)
adsorption coefficients.23 The negative effect of CO was ascribed
Published on 16 October 2015. Downloaded by Tubitak - Ulakbim on 4/2/2019 12:26:32 PM.

generated at 35 kV and 20 mA. Specimens for XRD were


to an extensive coverage of the catalyst surface by adsorbed prepared by compaction into a glass-backed aluminum sample
carbon monoxide, inhibiting the adsorption of hydrogen.24,26 holder. Data were collected over a 2q range from 5 to 70 . The
The main objectives of the present work are to investigate the line broadening of the Fe2O3 and MnO2 diffraction peaks
kinetics of the CO hydrogenation reaction over a Fe–Mn fused localized at 33.2 and 39 2q values was used to estimate the
catalyst, and also to investigate the power-law equation model average particle sizes, according to Scherrer’s equation.
and obtain the kinetic parameters using these models. 2.2.2. BET measurements. BET surface areas, pore volumes
Furthermore, we also attempt to investigate the effect of reac- and average pore sizes of the catalyst precursor and calcined
tion temperature on the reactant partial orders. samples (before and aer the test) were measured by N2 phys-
isorption using a Quantachrome Nova 2000 automated system
(USA). Each catalyst sample was degassed under a nitrogen
2. Experimental atmosphere at 300  C for 3 h. In order to obtain the BET surface
2.1. Catalyst preparation areas, pore volumes and average pore sizes, different samples
were evacuated at 196  C for 66 minutes.
The Fe–Mn catalyst used in the present work was prepared
using a fusion procedure. In order to prepare the fused iron
catalyst, the required amounts of Fe(NO3)3$9H2O (99% Merck), 2.3. Catalyst testing
Mn(NO3)2$4H2O (99% Merck), La2O3 (10 wt% based on the total The experiments were carried out in a xed-bed tubular stain-
catalyst weight) and Cs2O (1 wt% based on the total catalyst less steel micro-reactor. A schematic representation of the
weight), with a nominal composition of 50% Fe–50% Mn–10 experimental set-up is shown in Fig. 1. All gas lines to the
wt% La2O3–1 wt% Cs2O, were premixed in a crucible. The ob- reactor bed were made from 1/400 stainless steel tubing. Three
tained mixture was heated and dried at 120  C for 14 h in an mass ow controllers (Brooks, Model 5850E), equipped with
oven to give a material denoted as the catalyst precursor. The a four-channel read-out and control equipment (Brooks 0154),
obtained precursor was fused in an electrical furnace at 1500  C were used to automatically adjust the ow rate of the inlet gases
for 2 h and then cooled slowly. To prevent mass transfer (CO, H2 and N2, with purities of 99.999%). The mixed gases in

Fig. 1 Schematic representation in a flow diagram of the reactor used: 1 ¼ gas cylinders, 2 ¼ pressure regulators, 3 ¼ needle valves, 4 ¼ ball
valves, 5 ¼ mass flow controllers (MFC), 6 ¼ digital pressure controllers, 7 ¼ pressure gauges, 8 ¼ non-return valves, 9 ¼ mixing chamber, 10 ¼
valves, 11 ¼ tubular furnace, 12 ¼ tubular reactor and catalyst bed, 13 ¼ temperature indicators (digital program controller), 14 ¼ resistance
temperature detector (RTD), 15 ¼ condenser, 16 ¼ trap air, 17 ¼ back pressure regulator (BPR), 18 ¼ flow meter, 19 ¼ silica gel column, 20 ¼ gas
chromatograph (GC) and 21 ¼ hydrogen generator.

95288 | RSC Adv., 2015, 5, 95287–95299 This journal is © The Royal Society of Chemistry 2015
View Article Online

Paper RSC Advances

the mixing chamber passed into the reactor tube, which was preheating zone, the catalyst bed and the underneath zone of
placed inside a tubular furnace (Atbin, model ATU 150-15) the reactor, were checked using three separate thermocouples
capable of producing temperatures up to 1500  C and placed in different parts of the reactor. The temperature of the
controlled by a digital programmable controller (DPC). The catalyst bed was monitored with a thermocouple located exactly
reactor tube was constructed from stainless steel tubing; its in the middle of the catalyst bed. The inlet feed gas arrived from
internal diameter was 20 mm, with the catalyst bed situated in the top of the reactor, and the outlet products exited from the
the middle of the reactor. This single tubular micro-reactor was lower part of the reactor. The meshed catalyst (2.0 g) was held in
surrounded by an alumina jacket to achieve a uniform wall the middle of the reactor using quartz wool. An electronic back
temperature along the length of the reactor. A preheating zone pressure regulator was used, which can control the total pres-
ahead of the catalyst packing was lled with inert quartz glass sure of the desired process using a remote control via integra-
Published on 16 October 2015. Downloaded by Tubitak - Ulakbim on 4/2/2019 12:26:32 PM.

beads. External heating was provided by an electrical element tion with the TESCOM soware package, which can improve or
wrapped around the alumina jacket and placed in through the modify its efficiency; it is capable of working on pressures
rebrick part. The required amount of the catalyst was diluted ranging from atmospheric pressure to 100 bar. The catalyst was
using inert silica sand with the same particle size range as the pre-reduced in situ at atmospheric pressure under H2–N2 (ow
catalyst sample, and placed among the inert quartz glass beads. rate of each gas ¼ 50 ml min1) at 350  C for 16 h before
The temperatures of all of the different zones, including the synthesis gas exposure. The FTS was carried out under reaction

Table 1 Experimental conditions for kinetic evaluations at Ptot ¼ 1–15 bar, T ¼ 573–603 K, H2/CO ¼ 1–3 and GHSV ¼ 4500 h1 in a fixed-bed
reactor (FBR)

No. T (K) Ptotal (bar) H2/CO PCO (bar) PH2 (bar) PH2O (bar) R (mol min1 gcat1)

1a 573.15 1 0.97 0.38 0.37 0.03 8.00  106


2 573.15 1 1.95 0.22 0.43 0.04 1.55  105
3 573.15 3 1.97 0.64 1.26 0.11 2.22  105
4a 573.15 6 1.02 2.27 2.32 0.17 2.03  105
5 573.15 6 2.02 1.25 2.53 0.25 2.94  105
6 573.15 9 3.15 1.27 4.00 0.46 3.32  105
7a 573.15 12 3.19 1.66 5.30 0.70 3.35  105
8 573.15 15 1.04 5.54 5.76 0.36 3.12  105
9 573.15 15 2.03 3.00 6.10 0.80 3.50  105
10 573.15 15 3.33 2.00 6.65 0.90 3.49  105
11a 583.15 1 0.95 0.38 0.36 0.03 1.26  105
12 583.15 1 1.95 0.22 0.43 0.03 1.76  105
13 583.15 6 0.99 2.25 2.22 0.19 2.48  105
14 583.15 9 1.03 3.25 3.36 0.21 3.54  105
15a 583.15 9 1.97 1.78 3.51 0.45 3.88  105
16 583.15 12 1.97 2.39 4.70 0.60 3.89  105
17 583.15 12 3.25 1.54 5.00 0.75 4.04  105
18 583.15 15 1.03 5.39 5.55 0.42 3.46  105
19a 583.15 15 2.03 2.93 5.95 0.80 3.69  105
20 583.15 15 3.34 1.86 6.22 0.93 4.00  105
21 593.15 3 2.00 0.62 1.24 0.14 3.16  105
22 593.15 3 3.07 0.43 1.32 0.15 3.24  105
23a 593.15 6 0.91 2.20 2.00 0.23 3.10  105
24 593.15 6 3.05 0.80 2.44 0.43 3.60  105
25 593.15 9 1.99 1.76 3.50 0.50 3.89  105
26 593.15 9 3.28 1.15 3.77 0.60 3.97  105
27a 593.15 12 2.05 2.29 4.70 0.70 3.95  105
28 593.15 12 3.36 1.49 5.00 0.86 4.10  105
29 593.15 15 1.96 2.84 5.56 1.00 3.97  105
30 593.15 15 3.37 1.78 6.00 1.15 3.97  105
31a 603.15 1 0.89 0.37 0.33 0.03 1.98  105
32 603.15 1 1.85 0.20 0.37 0.03 2.10  105
33 603.15 1 2.87 0.15 0.43 0.04 2.08  105
34a 603.15 3 0.91 1.10 1.00 0.10 3.08  105
35 603.15 6 0.88 2.15 1.90 0.27 3.70  105
36 603.15 9 0.94 3.20 3.00 0.40 3.67  105
37a 603.15 12 0.74 5.00 3.70 0.80 3.85  105
38 603.15 12 1.88 2.23 4.20 0.90 4.00  105
39 603.15 15 1.82 2.74 5.00 1.00 3.95  105
40a 603.15 15 3.33 1.80 6.00 1.20 4.05  105
a
Experiments repeated three times.

This journal is © The Royal Society of Chemistry 2015 RSC Adv., 2015, 5, 95287–95299 | 95289
View Article Online

RSC Advances Paper

conditions of T ¼ 573–603 K, pressures of 1–15 bar, H2/CO feed


ratios of 0.7–3.4 and a space velocity of 4500 h1. In each test,
2.0 g of catalyst was loaded and the reactor operated for about
24 h to ensure that steady-state operations were attained. Some
experiments (reported in Table 1) were repeated three times,
and a comparison of the obtained results showed that a steady
state was achieved aer 24 h. Reactant and product streams
were analyzed on-line using a gas chromatograph (Thermo
ONIX UNICAM PROGC+) equipped with a sample loop, two
thermal conductivity detectors (TCDs) and one ame ionization
Published on 16 October 2015. Downloaded by Tubitak - Ulakbim on 4/2/2019 12:26:32 PM.

detector (FID) able to perform the analysis of a wide variety of


gaseous hydrocarbon mixtures; one TCD was used for the
Fig. 3 Variation of reaction rate as a function of GHSV value; condi-
analysis of hydrogen, and the other one was used for all the
tions: T ¼ 603 K, P ¼ 1 bar.
permanent gases such as N2, O2 and CO. The FID was used for
the analysis of hydrocarbons. The system is applicable to the
analysis of non-condensable gases, i.e. methane through to C8
effective heat removal should be accomplished from catalytic
hydrocarbons. The contents of the sample loop were injected
active sites during exothermic FTS.29 Before the kinetic experi-
automatically into an alumina capillary column (30 m  0.550
ments, mass transfer limitations in the xed-bed reactor were
mm). Helium was employed as a carrier gas for optimum
investigated by changing gas space velocity and catalyst particle
sensitivity (ow rate ¼ 30 ml min1). The GC calibration was
size. These conditions require the elimination of both pore
carried out using various calibration mixtures and pure
diffusion and intra-particle (lm resistance) mass transfer resis-
compounds obtained from the gas company Matheson (USA).
tances. Preliminary experiments were performed to test the pore
diffusion via decreasing the catalyst particle size. The fresh
2.4. Heat and mass transfer limitations catalyst was crushed and sieved to particles with diameters of
Heat and mass transfer limitations are two important factors 210–590 mm (30–70 ASTM mesh), then isometric catalysts were
which inuence the reaction rate when heterogeneous catalysts loaded under the same operating conditions. As shown in Fig. 2,
are employed, especially at high temperatures. Measuring the no pore diffusion limitation was observed for particles with sizes
reaction rate should not be inuenced by deactivation of the lower than 250 mm. As the particle size became smaller, the
catalyst or by heat and mass transfer limitations. When the mass reaction rate remained constant. In the second set of experi-
transfer rate is smaller than the reaction rate, it results in ments, intra-particle mass transfer limitation (lm resistance) for
a signicant effect of mass transfer on the total observed rate, or the reaction was investigated by variation of the space velocity
it even controls and limits the reactant transfer from the gas (resident time) with the upstream addition of N2 to the ow. As
phase to the catalyst surface.26,27 In the presence of mass transfer shown in Fig. 3, over the range of space velocities between 1500
limitations, the apparent activation energy for the reaction will be h1 and 6500 h1, as the space velocity increased, the reaction
approximately one-half the true activation energy for the surface rate remained constant and lm resistance was negligible. In
reaction.28 Gas space velocity, catalyst particle size and catalyst addition, by using a small amount of high-density catalyst (low
amount are three important factors inuencing heat and mass volume) prepared by the fusion method, the heat and mass
transfer in heterogeneous catalytic systems. With using a xed- transfer limitations are minimized.
bed reactor, the heat and mass transfer limitation problem
could be overcome by increasing the reaction temperature, and
3. Kinetic experiments
The FTS kinetic experiments were carried out with a mixture of
H2, CO and N2 at the temperature range of 573–603 K, P ¼ 1–15
bar, H2/CO ¼ 0.7–3.4 and GHSV ¼ 4500 h1. The required
amount of the catalyst (2.0 g) was diluted using 10 g of inert
silica sand with the same particle size as the catalyst sample and
placed among the inert quartz glass beads. For the kinetic
measurement tests, the reactor was operated for 24 hours until
the measurements became stable. The experimental conditions
and obtained data are presented in Table 1. To avoid the effect
of deactivation, fresh catalysts were loaded in each experiment
series. To achieve isothermal conditions in a catalytic bed, the
catalyst was diluted with inert materials (quartz and asbestos);
axial temperature distribution was ensured using Mear’s crite-
Fig. 2 Variation of reaction rate as a function of particle size; condi- rion.30,31 In order to avoid channelization phenomena, the
tions: T ¼ 603 K, P ¼ 1 bar and GHSV ¼ 3600 h1. following simplied relation between catalyst bed length (Lb)

95290 | RSC Adv., 2015, 5, 95287–95299 This journal is © The Royal Society of Chemistry 2015
View Article Online

Paper RSC Advances

and mean catalyst particle diameter (dp) was fullled: Lb/dp > 50. CO + 2s # Cs + Os; (9)
We have a differential ow reactor when we choose to consider
the rate to be constant at all points within the reactor. Since k1PCOqs2 ¼ k1qCqO; (10)
rates are concentration-dependent, this assumption is usually k1 qC qO
reasonable only for small conversions or for shallow small K1 ¼ ¼ ; (11)
k1 PCO qs 2
reactors. For each run in a differential reactor, the plug ow
performance equation becomes as follows: k1
ð XCO;out ð XCO;out qC qO ¼ aPCO qs 2 ; a ¼ ; (12)
Wcat dXCO 1 XCO;out  XCO;in k1
 ¼ ¼ dXCO ¼ ;
FCO XCO;in rCO ðrCO Þavg XCO;in ðrCO Þavg
qC ¼ qO ¼ (aPCO)0.5qs, (13)
Published on 16 October 2015. Downloaded by Tubitak - Ulakbim on 4/2/2019 12:26:32 PM.

(3)
 where a is the equilibrium constant of the CO adsorption step.
FCO ðXCO;out  XCO;in Þ FCO;in  FCO;out
ðrCO Þavg ¼ ¼ : (4) qH is obtained from the below steps:
Wcat Wcat
Therefore, in brief: H2 + 2s # 2Hs; (14)
Wcat XCO
 ¼ ; (5) k2PH2qs2 ¼ k2qH2; (15)
FCO rCO
k2 qH 2
and hence: K2 ¼ ¼ ; (16)
k2 PH2 qs 2

XCO FCO
rCO ¼ : (6)  
Wcat k2 k2
qH 2 ¼ PH2 qs 2 ; b ¼ ; (17)
k2 k2
The molar ow rate of carbon monoxide in the feed is calculated
as follows:
qH ¼ (bPCO)0.5qs. (18)
 n PCO
FCO ¼ n CCO ¼ : (7)
RT
The free sites fraction (qs) is calculated from the site balance:
X
n

4. Results and discussion qs þ qis ¼ 1; (19)


i¼1
4.1. Kinetic models and rate equations
where qis refers to the surface fraction occupied with adsorbed
In order to obtain the rate equation, rstly a reaction mecha-
species (C, H and O here). The site balance thus becomes:
nism should be considered. For determination of kinetic
models, four different mechanisms are presented, on the basis qs + qC + qO + qH ¼ 1. (20)
of different monomer formation and carbon chain repartition
rate. The elementary reactions of these four offered mecha-
Table 2 Reaction schemes of CO hydrogenation
nisms are summarized in Table 2. For the derivation of each
kinetic model, rstly one of the elementary reaction steps was Model Number Elementary reaction
considered as the rate-determining step (RDS) and all the other
steps were assumed to be at equilibrium. With consideration of FT-I 1 CO + 2s 4 Cs + Os
different RDSs for the proposed models, 14 different rate 2 H2 + 2s 4 2Hs
3 Cs + Hs 4 HCs + s
expressions were obtained (presented in Table 3). Finally, all of
4 HCs + Hs 4 H2Cs + s
the resulting rate expressions were tted separately against 5 Os + Hs / HOs + s
experimental data. According to the obtained results, FT-I-1 was 6 HOs + Hs / H2O + 2s
chosen as the best tted model (the rst step of the elementary FT-II 1 CO + s 4 COs
reaction of the FT-I model was considered as the RDS). The 2 H2 + 2s 4 2Hs
3 COs + Hs 4 HCOs + s
elementary reaction steps of the FT-I model are presented in
4 HCOs + Hs 4 Cs + H2O + s
Table 2. The reaction rate of the RDS 1 for FT-I is as follows: 5 Cs + Hs 4 CHs + s
6 CHs + Hs 4 CH2s + s
rCO ¼ k1PCOqs2 (molCO gcat1 min1), (8) 7 Os + Hs 4 HOs + s
8 HOs + Hs / H2O + 2s
where qs refers to the fraction of free sites. As can be observed in FT-III 1 CO + s 4 COs
Table 2, the adsorbed species in the FT-I model are C, H and O. 2 COs + H2 4 H2COs
3 H2COs + H2 4 CH2s + H2O
The terms qC, qH and qO, respectively, refer to the surface frac-
FT-IV 1 CO + s 4 COs
tions occupied by carbon, hydrogen and oxygen, which can be 2 H2 + s 4 H2s
calculated via the site balance, with the preceding reaction steps 3 COs + H2s 4 H2COs + s
being at quasi-equilibrium: 4 H2COs + H2s 4 H2O + s

This journal is © The Royal Society of Chemistry 2015 RSC Adv., 2015, 5, 95287–95299 | 95291
View Article Online

RSC Advances Paper

Table 3 Reaction rate expressions for CO hydrogenation

Number of proposed
model Rate equation Parameters MARR (%)

FT-I-1 KPCO K ¼ k1 5.50


 RCO ¼ h i2 k1
1þ 2ðaPCO Þ0:5 þ ðbPH2 Þ0:5 a¼
k1
k2

k2
FT-I-2 KPH2 K ¼ k2 10.62
 RCO ¼ h i2
0:5 0:5 k1
1þ 2ðaPCO Þ þ ðbPH2 Þ a¼
Published on 16 October 2015. Downloaded by Tubitak - Ulakbim on 4/2/2019 12:26:32 PM.

k1
k2

k2
FT-I-3 KPCO 0:5 PH2 0:5 K ¼ k3K10.5K20.5 8.15
 RCO ¼ h i2 k1
1þ 2ðaPCO Þ 0:5
þ ðbPH2 Þ0:5 a¼
k1
k2

k2
FT-I-4 KPCO 0:5 PH2 K ¼ k4K3K2K10.5 15.54
 RCO ¼ h i2 k1
1þ 2ðaPCO Þ 0:5
þ ðbPH2 Þ0:5 a¼
k1
k2

k2
FT-II-1 KPCO K ¼ k1 12.14
 RCO ¼  0:5
1þ aPCO þ bPH2 k1

k1
k2

k2
FT-II-2 KPH2 K ¼ k2 17.67
 RCO ¼  0:5
1þ aPCO þ bPH2 k1

k1
k2

k2
FT-II-3 KPCO PH2 0:5 K ¼ k3K1K2 22.43
 RCO ¼ h  0:5 i2 k1
1þ aPCO þ bPH2 a¼
k1
k2

k2
FT-II-4 KPH2 PCO K ¼ k4K2K3K4 21.32
 RCO ¼ h  0:5 i2 k1
1þ aPCO þ bPH2 a¼
k1
k2

k2
FT-III-1 KPCO K ¼ k1 29.36
 RCO ¼
1þ aPCO k1

k1
FT-III-2 KPCO PH2 K ¼ k2K1 15.32
 RCO ¼
1þ aPCO k1

k1
FT-III-3 KPCO PH2 2 K ¼ k3K1K2 23.32
 RCO ¼ k1
1þ aPCO a¼
k1
FT-IV-1 KPCO K ¼ k1 18.30
 RCO ¼
1þ aPCO þ bPH2 k1

k1
k2

k2
FT-IV-2 KPH2 K ¼ k2 28.65
 RCO ¼
1þ aPCO þ bPH2 k1

k1
k2

k2
FT-IV-3 KPCO PH2 K ¼ k3K1K2 32.12
 RCO ¼  2
1þ aPCO þ bPH2 k1

k1
k2

k2

95292 | RSC Adv., 2015, 5, 95287–95299 This journal is © The Royal Society of Chemistry 2015
View Article Online

Paper RSC Advances

Substituting eqn (13) and (18) into eqn (20), the free active sites by the soware Polymath 6.0, were used to assess the quality of
fraction is obtained as follows: the regression models and to compare the various models. The
parameters that were used in the soware Polymath 6.0 consist of
qs + 2(aPCO)0.5qs + (bPH2)0.5qs ¼ 1; (21) the following: graph, residual plot, condence interval, R2, Radj2,
variance and Rmsd. These are dened as follows:
qs(1 + 2(aPCO)0.5 + (bPH2)0.5) ¼ 1; (22) 4.2.1. Graph. Graph is a plot on the basis of the calculated
1 and measured values of RCO for each proposed model. An
qs ¼   : (23)
1þ 2ðaPCO Þ 0:5
þ ðbPH2 Þ0:5 inappropriate model shows differing trends. Fig. 5 compares
the experimental RCO with that calculated from the expression
which was obtained for FT-I-1 (Table 3), with the assumption
Published on 16 October 2015. Downloaded by Tubitak - Ulakbim on 4/2/2019 12:26:32 PM.

By substituting eqn (23) into eqn (8), the following rate that step 1 is the rate-controlling step.
expression is obtained: 4.2.2. Residual plot. The residual plot is a plot that shows
the difference between the calculated and measured values of
KPCO
rCO ¼  2 ; (24) the dependent variable as a function of the measured values.
1þ 2ðaPCO Þ0:5 þ ðbPH2 Þ0:5 The residuals between the proposed model and the experi-
mental values should be normally distributed, with a zero
where K ¼ k1.
average line. A comparison between calculated and experi-
A schematic representation of the CO hydrogenation reaction
mental CO conversion is presented in Fig. 6. This gure shows
over the Fe–Mn fused catalyst, for production of different hydro-
that the residual relative errors (RRs) between model and
carbons according to the best tted model, is illustrated in Fig. 4.
experiment are mostly distributed around the zero line.
4.2.3. Condence interval. If the condence interval is
4.2. Kinetic parameters estimation smaller than (or at least equal to) the respective parameter
Model parameters were calculated from the experimental data values (in absolute values), then the regression model is stable
and optimized with statistical indicators. Various plots, provided and statistically valid.

Fig. 4 Schematic description of the FT mechanism according to the FT-I-1 model.

This journal is © The Royal Society of Chemistry 2015 RSC Adv., 2015, 5, 95287–95299 | 95293
View Article Online

RSC Advances Paper


Published on 16 October 2015. Downloaded by Tubitak - Ulakbim on 4/2/2019 12:26:32 PM.

Fig. 5 Comparison between experimental and calculated reaction rates using eqn (8). Reaction conditions: T ¼ 573–603 K, P ¼ 1–15 bar, H2/
CO ¼ 1/1–3/1 and GHSV ¼ 4500 h1.

4.2.4. R2 and Radj2. The correlation coefficients were used 4.2.5. Variance and Rmsd. Variance and Rmsd are dened as
to judge whether the model correctly represents the data. These shown in eqn (28) and (29):
parameters are dened as shown in eqn (25)–(27): X
n
! ð yi  yÞ2
1 X n
i¼1
y¼ yi ; (25) S2 ¼ ; (28)
n i¼1 exp ðn 1Þ

n  2 " #
X n  2 2
1 X
yiexp  yicalc Rmsd ¼ yiexp  yicalc : (29)
i¼1 n i¼
R2 ¼ 1 n 
X  ; (26)
yiexp  y
i¼1
Some statistical indicators that were used to assess the
  quality of the proposed model (expression FT-I, RDS 1) in Table
1  R ðn 1Þ2
Radj 2 ¼ 1 : (27) 3 are summarized in Table 4.
np
For estimation of the best kinetic model, we made the
following assumptions: (1) the mass transport limitations and
In the above formulas, the notation n, yi, exp and calc pressure drop are negligible, and (2) the suitable values for all
denotes the number of experimental data points, specic parameters must be positive and all offered models with nega-
observations, observed data and calculated data, respectively. tive values of parameters will be refused.

Fig. 6 The relative residuals for the CO consumption rate ( Rexp  Rcal).

95294 | RSC Adv., 2015, 5, 95287–95299 This journal is © The Royal Society of Chemistry 2015
View Article Online

Paper RSC Advances

Table 4 Obtained values of kinetic parameters for the fitted model


(FT-I-1) and statistical criteria

Parameter Value Dimension

Ea 105.00  3.70 kJ mol1


K0 5.93  106 mol gcat1 min1
bar1
k(573) 1.30  104 mol gcat1 min1
bar1
k(583) 1.96  104 mol gcat1 min1
bar1
Published on 16 October 2015. Downloaded by Tubitak - Ulakbim on 4/2/2019 12:26:32 PM.

k(593) 2.95  104 mol gcat1 min1


bar1
k(603) 4.35  104 mol gcat1 min1
bar1
Fig. 7 Arrhenius plot of the rate constant (k) according to the FT-I-1
a0 1.28  108 bar1
model results.
DHCO 68.00  4.50 kJ mol1
b0 2.98  107 bar1
DHH2 48.00  3.20 kJ mol1
R2 0.94 — The relation between the temperature and the reaction rate
Radj2 0.91 — constant, according to the kinetic parameters obtained from the
Rmsd 3.918  107 —
FT-I-1 model, is shown in Fig. 7. According to the Arrhenius-type
Variance 7.23  1012 —
MARR (%) 5.50 — equation (eqn (30)), a plot of ln(k) versus 1/T should give
a straight line with a negative slope of –Ei/R. The relation
between the temperature and the adsorption enthalpy (van’t
Estimation of parameters and model distinction have been Hoff plot) for the FT-I-1 model is shown in Fig. 8.
accomplished using a nonlinear regression model and the The mean absolute relative residual (MARR%) between the
soware Polymath 6.0. Fig. 6 displays the residuals between experimental and the calculated consumption rates of CO is
the offered model and the experimental data, distributed dened as follows:
randomly around the zero line. This gure conrms that the Nexp 
X 
rexp  rcal 
offered model is in good agreement with the experimental MARRð%Þ ¼    1 100; (33)
 r  N
i¼1 exp exp
data.
According to the obtained results, the best expression that where Nexp is the number of experimental points. Eqn (24)
describes the experimental results for the FT reaction over the shows the best t to the experimental data. A comparison of the
Fe–Mn fused catalyst is as follows (eqn (24)): calculated and experimental consumption rates of CO for the
KPCO FT-I-1 model is shown in Fig. 9; the MARR% of this model was
RCO ¼  2 : obtained as 5.50%. This value is reasonable and shows that the
1þ 2ðaPCO Þ0:5 þ ðbPH2 Þ0:5
predicted values are 5.50% different from the observed values.
The MARR% values of the other obtained kinetic models are
Kinetic and adsorption parameters which depend on presented in Table 3; as can be seen in this table, the FT-I-1
temperature were described and calculated according to the model has the lowest MARR% value. The obtained activation
Arrhenius and van’t Hoff equations, respectively (eqn (30) and energy for FT-I-1 was found to be 105  3.7 kJ mol1.
(31)):

ki(T) ¼ ki,0 exp(Ei)/RT; (30) 4.3. Kinetic investigation using a power law model
a(T) ¼ a0 exp(DHads)/RT. (31) The effect of reactants on the reaction rate has been investi-
gated by many researchers.12,22,24–26,32–34 In the present work, we
In these equations, E and DH refer to the activation energy attempted to investigate the relationship between the partial
and heat of adsorption, respectively; by substituting eqn (30) pressures of the reactants and temperature changes. By using
and (31) in the best tted model (FT-I-1), we have: the power law equation (eqn (34)), the order of reaction was
obtained at four temperatures for the FTS. To investigate the


ðKi;0 expðEi Þ RTÞPCO
RCO ¼  2 : (32)
ð1 þ 2ða0 expðDHCO Þ=RTÞPCO Þ0:5 þ ððb0 expðDHH2 Þ RTÞPH2 Þ0:5

This journal is © The Royal Society of Chemistry 2015 RSC Adv., 2015, 5, 95287–95299 | 95295
View Article Online

RSC Advances Paper

effect of water on the reaction rate, water pressure was entered 573 to 603 K, the partial orders of CO and H2O increased and the
in the power law equation: partial order of hydrogen decreased. The effect of temperature
on the reaction order is plotted in Fig. 10. Previous research
rFT ¼ kFTPCOaPH2bPH2Oc. (34) showed that the reaction order of H2 [b] was positive but the
reaction orders of CO [a] and H2O [c] were negative.23,24,35 It
The obtained results are summarized in Table 5. The ob- could be argued that the higher CO partial pressure leads to
tained results showed that, with increasing temperature from higher coverage of the catalyst surface by adsorbed CO. As is
shown in Fig. 10, upon increasing the temperature, the reaction
order was changed for all reactants. With increasing tempera-
ture, adsorbed CO molecules are consumed more rapidly, and
Published on 16 October 2015. Downloaded by Tubitak - Ulakbim on 4/2/2019 12:26:32 PM.

so the number of free active sites for adsorption of H2 molecules


is increased. An increase in the number of free active sites leads
to easier adsorption of hydrogen molecules on the catalyst
surface, so that the positive effect of higher H2 partial pressure
on the reaction rate is decreased. The calculated activation
energy for the CO hydrogenation reaction according to the
power law model was found to be 95.54  2.5 kJ mol1; the high
activation energy for hydrocarbon formation suggests that the
diffusion interference is not signicant in experiments.36,37 As
with intra-particle diffusion limitations, the presence of
external mass transfer limitations could be detected via
measuring the apparent activation energy. An external mass
transfer control regime could lead to an apparent activation
energy of just a few kJ mol1.38
According to the power law equation (eqn (34)), four rate
constants (k) were obtained for the four different temperatures
(Table 6). In order to show the relationship between the inverse

Table 5 The obtained results for the reaction order, using the power
law equation, at different temperatures (573–603 K)

No. T (K) Power law equation

1 573 R573 ¼ 9.13  106PCO0.27PH20.84PH2O0.23


Fig. 8 Van’t Hoff plots of the adsorption coefficients of (A) a CO 2 583 R583 ¼ 1.44  105PCO0.18PH20.65PH2O0.15
molecule and (B) a H2 molecule, according to the results obtained 3 593 R593 ¼ 2.23  105PCO0.09PH20.39PH2O0.11
from the FT-I-1 model. 4 603 R603 ¼ 2.38  105PCO0.08PH20.28PH2O0.08

Fig. 9 The calculated CO consumption rate versus the experimental CO consumption rate for the FT-I-1 model.

95296 | RSC Adv., 2015, 5, 95287–95299 This journal is © The Royal Society of Chemistry 2015
View Article Online

Paper RSC Advances


Published on 16 October 2015. Downloaded by Tubitak - Ulakbim on 4/2/2019 12:26:32 PM.

Fig. 11 Arrhenius plot according to the power law equation results.


Fig. 10 Effect of reaction temperature on the reaction partial orders.

using powder X-ray diffraction, and the obtained patterns are


of the temperature and the logarithm of the rate constant ob- illustrated in Fig. 14. The actual phases identied in the fresh
tained from this model, the Arrhenius equation (eqn (30)) was catalyst were Fe2O3 (hexagonal), MnO2 (tetragonal), Fe3O4
used. The obtained plot, displayed in Fig. 11, illustrates this (cubic) and Mn2O3 (orthorhombic). In order to identify the
relationship, showing a straight line with the negative slope changes in the catalyst during the reaction and to detect the
(Ea/R). The value of E/R was found to be 11 492, which phases formed, this catalyst was characterized by XRD aer the
yields an activation energy of 95.54 kJ mol1. test and its phases were found to be MnO (cubic), FeO (cubic),
The kinetic parameters and activation energy of the power Fe (cubic) and Fe2C (monoclinic). As can be seen, in the tested
law model (eqn (34)) were calculated, and the obtained results catalyst there are oxidic and iron carbide phases, which are both
are presented in Table 6. A comparison of the calculated and active for FTS. Zhang and Schrader39 concluded that two active
experimental consumption rates of CO (for eqn (34)) is shown in sites operate simultaneously on the surface of iron catalysts:
Fig. 12; the MARR% of this model was obtained as 8.86%. This Fe0/Fe-carbides and magnetite (Fe3O4). The carbide phase is
value is reasonable and shows that the predicted values are active towards the dissociation of CO and formation of hydro-
8.86% different from the observed values. Fig. 13 shows carbons, while the oxide phase adsorbs CO associatively and
a comparison between the experimental and calculated produces predominantly oxygenated products. The crystallite
intrinsic reaction rates using the power law equation. size of the fresh catalyst was calculated by the Debye–Scherrer
equation from the XRD data. The average crystallite sizes for
Fe2O3 and MnO2 were calculated to be 91 and 83 nm,
4.4. Catalysts characterization respectively.
Characterization of the fresh catalyst (sample before the test)
and the used sample (catalyst aer the test) was carried out

Table 6 Values of kinetic parameters for the power law model (eqn
(34))

Parameter Value Dimension

Ea 95.54  2.5 kJ mol1


K0 5.01  103 mol gcat1 min1
bar1
k(573) 9.13  106 mol gcat1 min1
bar1
k(583) 1.44  105 mol gcat1 min1
bar1
k(593) 2.23  105 mol gcat1 min1
bar1
k(603) 2.38  105 mol gcat1 min1
bar1
R2 0.90 — Fig. 12 The calculated CO consumption rate versus the experimental
MARR (%) 8.86 — CO consumption rate, using the power law model.

This journal is © The Royal Society of Chemistry 2015 RSC Adv., 2015, 5, 95287–95299 | 95297
View Article Online

RSC Advances Paper


Published on 16 October 2015. Downloaded by Tubitak - Ulakbim on 4/2/2019 12:26:32 PM.

Fig. 13 Comparison between experimental and calculated reaction rates for the power law equation (eqn (34)).

Fig. 14 XRD patterns of calcined catalysts (before and after the test).

Characterization of the fresh catalyst was also carried out a range of operating conditions. Four different mechanisms,
using BET measurements; the obtained results showed that the according to the carbide mechanism using the Langmuir–Hin-
surface area of this sample was 14 m2 g1. The surface area of shelwood–Hougen–Watson and Eley–Rideal mechanisms, were
the catalyst could be partially improved during the reduction derived for CO hydrogenation. The unknown kinetic parameters
process.3,5 The addition of Mn appeared to increase the BET were estimated from experimental data using a nonlinear regres-
surface area of the Fe catalysts.4 Dry has suggested that an alkali sion (Levenberg–Marquardt) method. The reaction rate of CO
promoter can decrease the surface area of the Fe catalyst by hydrogenation is determined by the formation of the methylene
increasing the Fe crystallite size.40 monomer. In the best tted model (FT-I-1), both reactants (CO and
H2) were dissociated and adsorbed on the catalyst surface.
5. Conclusion Furthermore, the power law model was also proposed and evalu-
ated. The results in the present work also showed that upon
The kinetics of the CO hydrogenation reaction were investigated increasing the reaction temperature, the reaction partial orders of
over a fused Fe–Mn catalyst in a xed-bed micro-reactor over all reactants were changed.

95298 | RSC Adv., 2015, 5, 95287–95299 This journal is © The Royal Society of Chemistry 2015
View Article Online

Paper RSC Advances

13 G. P. V. d. Laan and A. A. C. M. Beenackers, Catal. Rev.: Sci.


Nomenclature
Eng., 1999, 41, 255–318.
14 F. G. Botes, Catal. Rev.: Sci. Eng., 2008, 50, 471–491.
rFT Rate of reaction 15 A. Nakhaei Pour, M. R. Housaindokht, S. F. Tayyari,
rWGS Rate of water gas shi reaction J. Zarkesh and M. R. Alaei, J. Nat. Gas Sci. Eng., 2010, 2,
rCO Rate of CO consumption 61–68.
rCO2 Rate of CO2 production 16 M. Sarkari, F. Fazlollahi, H. Ajamein, H. Atashi, W. C. Hecker

FCO Inlet molar ow of CO and L. L. Baxter, Fuel Process. Technol., 2014, 127, 163–170.
n Volumetric ow rate of input gas 17 G. P. V. d. Laan and A. A. C. M. Beenackers, Appl. Catal., A,
CCO Concentration of CO 2000, 193, 39–53.
Published on 16 October 2015. Downloaded by Tubitak - Ulakbim on 4/2/2019 12:26:32 PM.

PCO Partial pressure of CO 18 Y. N. Wang, W. P. Ma, Y. J. Lu, J. Yang, Y. Y. Xu, H. W. Xiang,


PH2 Partial pressure of H2
Y. W. Li, Y. L. Zhao and B. J. Zhang, Fuel, 2003, 82, 195–213.
PH2O Partial pressure of H2O
19 B. T. Teng, J. Chang, C. H. Zhang, D. B. Cao, J. Yang, Y. Liu,
T Gas temperature
X. H. Guo, H. W. Xiang and Y. W. Li, Appl. Catal., A, 2006,
R Universal gas constant
301, 39–50.
k Rate constant of reaction
20 W. Erley, P. H. Mcbreen and H. Ibach, J. Catal., 1983, 84,
Ea Activation energy
a Adsorption coefficient of CO 229–234.
b Adsorption coefficient of H2 21 B. H. Davis, Catal. Today, 2009, 141, 25–33.
DHH2 Adsorption enthalpy of H2 22 F. G. Botes and B. B. Breman, Ind. Eng. Chem. Res., 2006, 45,
DHCO Adsorption enthalpy of CO 7415–7426.
XCO The conversion of CO 23 M. E. Dry, T. Shingles, L. J. Boshoff and G. J. Oosthuizen, J.
W The catalyst weight Catal., 1969, 15, 190–199.
24 E. v. Steen and H. Schulz, Appl. Catal., A, 1999, 186, 309–320.
25 V. R. R. Pendyala, G. Jacobs, J. C. Mohandas, M. Luo, W. Ma,
M. K. Gnanamani and B. H. Davis, Appl. Catal., A, 2010, 389,
References
131–139.
1 N. Moazami, M. L. Wyszynski, H. Mahmoudi, A. Tsolakis, 26 F. G. Botes, Ind. Eng. Chem. Res., 2009, 48, 185–1865.
Z. Zou, P. Panahifar and K. Rahbar, Fuel, 2015, 154, 140–151. 27 X. Zhan and B. H. Davis, Appl. Catal., A, 2002, 236, 149–161.
2 A. P. Steynberg and M. E. Dry, Fischer–Tropsch technology, 28 P. Schneider and P. Mitschka, Chem. Eng. Sci., 1969, 24,
Elsevier Science & Technology Books, 2004, p. 533. 1725–1731.
3 M. S. Spencer, in Catalyst Handbook, ed. V. Twigg, Wolfe 29 H. J. Kim, J. H. Ryu, H. Joo, J. Yoon, H. Jung and J. I. Yang,
Publishing Ltd, London, 1989, p. 35. Res. Chem. Intermed., 2008, 34, 811–816.
4 M. E. Dry, J. A. K. Plessis and G. M. Leuteritz, J. Catal., 1966, 30 D. E. Mears, ACS Monographs, 1974, 133, 218–228.
6, 194–199. 31 M. Mollavali, F. Yaripour, H. Atashi and S. Sahebdelfar, Ind.
5 G. C. Maiti, R. Malessa, U. Lochner, H. Papp and M. Baerns, Eng. Chem. Res., 2008, 47, 3265–3273.
Appl. Catal., A, 1985, 16, 215–225. 32 D. Tristantini, S. Lögdberg, B. Gevert, O. Borg and
6 B. Liu, J. Zhang, W. Cao and X. Yang, Fuel Process. Technol., A. Holmen, Fuel Process. Technol., 2007, 88, 643–649.
2009, 90, 1480–1485. 33 S. Storsaeter, O. Borg, E. A. Blekkan and A. Holmen, J. Catal.,
7 C. H. Zhang, Y. Yang, B. T. Teng, T. Z. Li, H. Y. Zheng, 2005, 231, 405–419.
H. W. Xiang and Y. W. Li, J. Catal., 2006, 237, 405–415. 34 R. Zennaro, M. Tagliabue and C. H. Bartholomew, Catal.
8 F. Fazlollahi, M. Sarkari, H. Gharebaghi, H. Atashi, Today, 2000, 58, 309–319.
M. M. Zarei, A. A. Mirzaei and W. C. Hecker, Chin. J. Chem. 35 M. Zhuo, A. Borgna and M. Saeys, J. Catal., 2013, 297, 217–
Eng., 2013, 21, 507–519. 226.
9 X. Zhang, Y. Liu, G. Liu, K. Tao, Q. Jin, F. Meng, D. Wang and 36 A. Sari, Y. Zamani and S. A. Taheri, Fuel Process. Technol.,
N. Tsubaki, Fuel, 2012, 92, 122–129. 2009, 90, 1305–1313.
10 D. J. Moodley, J. v. d. Loosdrecht, A. M. Saib and 37 J. Chang, L. Bai, B. T. Teng, R. Zhang, J. Yang, Y. Y. Xu,
H. J. W. Niemantsverdriet, in Advances in Fischer–Tropsch H. W. Xiang and Y. W. Li, Chem. Eng. Sci., 2007, 62, 4983–
Synthesis, Catalysts, and Catalysis, ed. B. H. Davis and M. L. 4991.
Occelli, CRC Press, Boca Raton, 2010, p. 53. 38 A. Y. Khodakov, W. Chu and P. Fongarland, Chem. Rev.,
11 C. H. Bartholomew, Mechanisms of catalyst deactivation, 2007, 107, 1692–1744.
Appl. Catal., A, 2001, 212, 17–60. 39 H. B. Zhang and G. L. Schrader, J. Catal., 1985, 95, 325–332.
12 T. K. Das, X. Zhan, J. Li, G. Jacobs, M. E. Dry and B. H. Davis, 40 M. E. Dry and G. J. Oosthuizen, J. Catal., 1968, 11, 18–24.
in Fischer–Tropsch Synthesis, Catalysts and Catalysis, ed. B. H.
Davis and M. L. Occelli, Elsevier B.V., 2007, p. 298.

This journal is © The Royal Society of Chemistry 2015 RSC Adv., 2015, 5, 95287–95299 | 95299

Вам также может понравиться