You are on page 1of 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/322073216

Catalytic Steam Gasification of Mengdong Coal in the Presence of Iron Ore for
Hydrogen-rich Gas Production

Article  in  Journal- Energy Institute · December 2017


DOI: 10.1016/j.joei.2017.12.005

CITATIONS READS

4 85

7 authors, including:

Zixu Yang Junhao Hu


Oklahoma State University - Stillwater Huazhong University of Science and Technology
24 PUBLICATIONS   376 CITATIONS    17 PUBLICATIONS   138 CITATIONS   

SEE PROFILE SEE PROFILE

Chen Yingquan Kezhen Qian


Huazhong University of Science and Technology Oklahoma State University - Stillwater
95 PUBLICATIONS   1,395 CITATIONS    11 PUBLICATIONS   501 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Heavy metals investigations and wastewater treatment View project

Carbon-based Catalyst System for CFP Under Microwave Energy View project

All content following this page was uploaded by Junhao Hu on 22 July 2019.

The user has requested enhancement of the downloaded file.


Journal of the Energy Institute 92 (2019) 391e402

Contents lists available at ScienceDirect

Journal of the Energy Institute


journal homepage: http://www.journals.elsevier.com/journal-of-the-energy-
institute

Catalytic steam gasification of Mengdong coal in the presence of iron


ore for hydrogen-rich gas production
Zixu Yang a, b, Junhao Hu a, Yang Li c, Yingquan Chen a, Kezhen Qian a, b, Haiping Yang a, *,
Hanping Chen a
a
State Key Laboratory of Coal Combustion, School of Energy and Power Engineering, Huazhong University of Science and Technology, Wuhan, 430074,
China
b
Bioproducts, Sciences and Engineering Laboratory, Department of Biological Systems Engineering, Washington State University, Richland, WA,
99354-1671, United States
c
CNNC Key Laboratory on Nuclear Reactor Thermal Hydraulics Technology, Nuclear Power Institute of China, Chengdu, 610041, China

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

Article history: Hydrogen-rich gas production from catalytic steam gasification of coal was investigated in the presence
Received 30 October 2017 of iron ore in a vertical fixed bed reactor. The addition of iron ore significantly promoted the H2 yields.
Received in revised form The effects of operation parameters (upper zone temperature, lower zone temperature, steam concen-
15 December 2017
tration, and iron to coal ratio) on the yield of selected gaseous products (H2 and CO) during catalytic
Accepted 18 December 2017
steam gasification, were studied using Taguchi method. The results of signal-to-noise ratio indicated that
Available online 26 December 2017
steam concentration and iron to coal ratio were the most important parameters in determining the yield
of H2 and CO, respectively. Semi-quantification X-ray diffraction analysis of iron ores indicated that in-
Keywords:
Coal crease in steam concentration intensified the oxidization of low valence iron compounds to Fe3O4. In
Catalytic steam gasification addition, formation of Fe3O4 was also favored with increasing reaction temperatures (600  Ce900  C).
Iron ore However, the formation of Fe3O4 was inhibited at higher reaction temperature (1000  C) due to the
Hydrogen-rich gas destruction of porous structures of the iron ore.
Taguchi method © 2018 Energy Institute. Published by Elsevier Ltd. All rights reserved.

1. Introduction

Clean coal utilization technologies have been paid close attention in past decade due to severe environmental issues, such as global
warming and air pollution caused by inefficient and unrestrained utilization of coal [1]. Coal gasification, the conversion of coal into gaseous
products which can be used for power generation and/or as syngas, is considered one of the most efficient and environmentally benign ways
of coal utilization. H2 is the targeted gaseous product from coal gasification due to its widespread applications in advanced fuel cells as a
clean energy carrier with zero net emissions.
The production of H2 from coal gasification can be realized using steam as the gasifying agent and promoted with catalysts addition.
Conventional steam gasification typically occurs at high temperatures (>900  C) due to its endothermic nature, making the process highly
energy intensive and economically uncompetitive. Moreover, the process is susceptible to the issues of tar formation, leading to serious
operation problems such as fouling and blocking of downstream pipelines. Recently, more efforts have been focused on catalytic steam
gasification, which converts solid carbonaceous materials, such as coal, biomass and solid wastes into H2-rich syngas at lower temperatures
(~700  C) [2e5]. The use of catalysts in steam gasification can significantly increase the gasification reaction rate and the overall carbon
conversion by effectively enhancing tar reforming. In addition, catalysts also offer improved product selectivity, allowing more efficient
process control in order to selectively produce gaseous products (e.g. H2) to meet specific applications [6].
The performance of various in-bed catalysts such as alkali and alkali earth metals, supported transition metals, and naturally occurring
magnesium-containing minerals (e.g. dolomite and olivine) have been investigated [7e11]. Alkali and alkali earth metals, especially in the
form of carbonates, are commercially attractive due to their superior catalytic activities in steam gasification and their vast availability [11].

* Corresponding author. 1037 Luoyu Road, Wuhan, Hubei, 430074, China.


E-mail address: yhping2002@163.com (H. Yang).

https://doi.org/10.1016/j.joei.2017.12.005
1743-9671/© 2018 Energy Institute. Published by Elsevier Ltd. All rights reserved.
392 Z. Yang et al. / Journal of the Energy Institute 92 (2019) 391e402

The transition metal, Ni, has been widely used for syngas or coal gas production due to its high catalytic activities in tar cracking and
hydrogenation of tar fragments [12,13]. Several studies reported nearly 100% tar removal in the presence of Ni-based catalysts [14e16].
Another transition metal, Fe, is especially suited for synthesis of hydrocarbon fuels and chemicals and purifying and conditioning of syngas
due to its high water-gas shift capability, low cost and insusceptibility to poisoning [17e21]. Lu et al. [18] used a FeC/Fe nanoparticles catalyst
for olefin-rich hydrocarbons synthesis from biomass-derived syngas via Fischer-Tropsch reaction. The CO and H2 conversion during catalytic
synthesis achieved about 87.5% and 85% at mild reaction conditions (310  C, 1000 psig). Uddin et al. [17] decomposed biomass tar over Fe2O3
catalyst in a double bed micro-reactor, and the results showed a tar removal efficiency over 90%. The activity of Fe2O3 catalysts for tar
decomposition seemed to be stable with cyclic use. Naturally occurring magnesium-containing minerals (e.g. dolomite and olivine) are used
directly as in-bed additives to enhance gasification efficiency due to their high catalytic reactivity, low cost and availability [22,23]. Turson
et al. [22] used olivine as a catalytic bed material in a moving-bed coal gasifier and achieved high H2 concentration of 60.4 vol.% from steam
gasification of Dayan lignite at 800  C and S/C ratio of 1.0. Othman et al. [24]observed that addition of dolomite significantly enhanced water
gas shift and methanation reactions during gasification of Adaro coal.
Recently, the naturally occurring mineral, iron ore, has attracted significant interest due to its catalytic activity for tar reforming [9,25].
Zou et al. [26] studied catalytic reforming of toluene over hematite derived from natural limonite, and found that a toluene conversion of
90% was obtained when gas hourly space velocity was lower than 5662 h1 at 700  C. Lind et al. [27] reported a novel chemical loop
reforming system for catalytic tar cleaning using ilmenite as catalytic bed material. They showed that the ilmenite reduced the total amount
of tar by 35% at a gas residence time of 0.4e0.5 s and significantly promoted water-gas shift reaction.
Iron ore offers advantages of wide availability, low cost, and non-toxicity. In addition, the catalyst regeneration step is not needed for iron
ore because the deactivated iron ore can be recycled as feedstock for ironmaking [9,25,28e30]. Tar produced from pyrolysis of biomass
diffuses into the channels of iron ore and results in carbon deposition (carbonized tar), providing a promising reducing agent in ironmaking
ðFeO þ C/Fe þ COÞ [30].
Coal or biomass tar can act as a cheap reducing agent of iron ore, and in return, iron ore can catalyze tar cracking [25,30]. However, iron
ore has rarely been used as a catalyst for conditioning of gas derived from coal gasification. The objective of this study was to investigate the
catalytic effects of iron ores on conditioning of coal gas for H2-rich gas production. Taguchi method was used in this study to identify the
influence of operation parameters (upper zone temperature, lower zone temperature, steam concentration, and iron to coal ratio) on
targeted gaseous products (H2 and CO). The variation of compositions and morphologies of iron ore after conditioning was also investigated
using X-ray diffraction (XRD) and Scanning Electron Microscopy (SEM).

2. Materials and method

2.1. Iron ore and coal samples

The iron ore was obtained from Dushan County at Hubei Province without any pretreatment. A bituminous coal, coming from Mengdong
coal mine was used in this study [31]. The samples were pulverized to 0.15e0.25 mm and dried at 105  C prior to use. Proximate, elemental
compositions and mineral contents of coal are summarized in Table 1.

2.2. Experimental setup and procedure

The coal gasification experiments were carried out in a vertical tubular reactor system (Fig. 1.), as described in our previous study [32].
Gasification occurs in a quartz tube reactor (I.D. 40 mm, total length: 500 mm) with two temperature ranges (Upper zone: gasification zone,
length: 250 mm; Lower zone: tar reforming zone, length: 250 mm). The fuel feeding system consists of a hopper and an auger feeder
powered by a motor. Prior to each trial, about 4.0 g of iron ore was loaded on the catalyst bed located 150 mm from the bottom of the reactor
and then heated in a N2 (high-purity, 99.99 vol.%) flow of 400 ml/min to the desired reaction temperature (600  Ce1000  C), which was held
constant for 10 min. Deionized water was introduced to the reactor through a syringe pump and vaporized by a home-made steam
generator, which consisted of a flask and an electric heater. When the desired reactor temperature was reached, N2 purge was switched to a
steam/N2 mixture of different concentrations (0 vol.%, 10 vol.%, 25 vol.%, 40 vol.% and 55 vol.%). The coal sample was fed continuously at a
rate of 0.2e0.3 g/min from the top of the reactor and thermally decomposed at the upper zone of the reactor. The volatiles were entrained
downward to the catalyst bed and exited the reactor from the bottom. The experiment was allowed to run for 40 min and then the reactor
was cooled to room temperature. The condensable volatiles were sampled using two impinger bottles chilled by an ice-water trap. The non-
condensable gases were sampled with a gas bag and analyzed using an Agilent 3000 Micro GC equipped with a thermal conductivity
detector (TCD). Gas analysis for each test was normalized, and the N2 concentrations in the reactor inlet and outlet were used as the internal

Table 1
Proximate, elemental analyses and mineral content of coal.

Proximate analysis (wt.%, ad.) Elemental analysis (wt.%, ad.)


M V A FC N C H S Oa
2.29 30.65 28.07 36.84 0.78 50.08 3.93 0.93 9.19
Mineral content (wt, ad/%)
Al2O3 SiO2 K2O CaO MgO Fe2O3 TiO2 P2O5 SO3
23.11 61.09 2.84 2.21 1.67 4.29 0.79 0.14 3.78

ad: air dried basis.


M: moisture; V: volatile; A: ash; FC: fixed carbon.
a
calculated by difference.
Z. Yang et al. / Journal of the Energy Institute 92 (2019) 391e402 393

Fig. 1. Reactor system for iron ore and coal gasification (1. Gas valve, 2. Flow meter, 3. Syringe pump controller, 4. Syringe pump, 5. Quartz bottle, 6. Heater, 7. Furnace, 8. Motor of
feeder, 9. Coal Storage, 10. Tubular reactor, 11. Ice bath, 12 and 13. Impinger bottle, 14 and 15. Gas filter, 16. Gas pump, 17. Gas analyzer, 18. Coal, 19. Iron ore).

standard to calculate the flow rate of the gas produced in the gasification. The yield of major gaseous products (H2, CO, CH4, CO2) is
quantified by Equation (1):

moles of the selected species in produced gas


Gas yield ðmmol=gÞ ¼ (1)
mass of biomass feed
The coal residue and spent iron ore were collected and sampled separately for further analysis.
Taguchi optimization of operating parameters in catalytic steam gasification was also performed in the aforementioned experimental
setup in a batch mode. The reason for selecting the batch mode for Taguchi optimization instead of continuous feeding is that feeding rate
itself is considered as a noise factor that needs to be avoided. The experimental procedure was similar to the abovementioned procedure,
except for the coal feeding process. Initially, a quartz basket with approximately 4 g coal sample was held on the top of the reactor. After the
temperatures at both zones reached the setting values and stabilized for 10 min, the quartz basket was quickly dropped into the upper zone
of the reactor. The coal sample was exposed to rapid decomposition and the volatile products were reformed by the iron ore in the lower
zone. The product sampling and analysis procedures were the same as stated above.

2.3. Experimental design

Taguchi design, also called robust design, applies fractional factorial design to determine the optimal operation conditions. Orthogonal
arrays and statistical analysis are combined in order to reduce the time and effort required to determine the important factors in operations.
In this study, Taguchi method was used to study the effects of four operational parameters (upper zone temperature, lower zone tem-
perature, iron ore to coal ratio, and steam concentration) on four responses (H2 yield, CO yield, H2þCO yield, and H2/CO ratio). The four
factors and their levels are summarized in Table 2. The complete design matrix of the experiments and results are shown in Table 6. The
collected data was analyzed using a statistical package, Minitab® 17 (Minitab Inc., State College, PA, USA).

Table 2
The control factors and their levels.

Factors Control factors Level

1 2 3 4 5
A Upper zone temperature ( C) 600 700 800 900 1000
B Lower zone temperature ( C) 600 700 800 900 1000
C Iron ore to coal ratio 0 1:4 1:2 1:1 2:1
D Steam concentration (%) 0 10 25 40 55
394 Z. Yang et al. / Journal of the Energy Institute 92 (2019) 391e402

Table 3
Diffraction peak information of iron species.

Species Card no. Diffraction peak position


Fe2O3 01-089-8103 33.2 (100%) 104, 35.7 (72.3%) 110, 49.6 (34.7%) 024, 54.2 (41.5%), 62.6 (26.4%), 64.2 (25.1%)
Fe3O4 01-075-0033 30.1 (28.7%), 35.5 (100%), 43.1 (19.9%), 57.0 (24.1%), 62.6 (33.6%)
FeO 01-075-1150 36.1 (67.4%), 42.0 (100%), 60.8 (46%)
Fe 03-065-9094 37.0 (18%), 41.8 (30%), 42.7 (100%), 43.8 (45.4%), 57.1 (14.9%)
Fe2C 03-065-1456 37.0 (18%), 41.8 (30%), 42.7 (100%), 43.8 (45.4%), 57.1 (14.9%)
Fe3C 01-089-3689 37.7 (20.5%), 41.4 (22.6%), 43.2 (100%), 57.3 (13.9%)
C 01-075-2078 26.6 (100%), 43.5 (11.3%)

Table 4
Key reactions occurring during coal pyrolysis and gasification.

Coal Pyrolysis Volatiles (1)


Pyrolysis zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
Coal! Cx Hy Oz þ ðH2 þ CO þ CO2 þ CH4 þ …Þ þ ðC þ AshÞ
|fflfflfflfflffl{zfflfflfflfflffl} |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflffl{zfflfflfflfflfflfflffl}
Tar component char
Small molecular gases
Secondary pyrolysis of volatiles Cx Hy Oz ! Crack Tars
þ ðH2 þ CO þ CO2 þ CH4 þ …Þ (2)
Steam reforming of volatiles Cx Hy Oz þ H2 O/CO þ H2 (3)
Steam-CH4-Reforming, CH4 þ H2 O/CO þ 3H2 (þ206 MJ/kmol) (4)
Water-Gas-Shift, CO þ H2 O/CO2 þ H2 (75 MJ/kmol) (5)
Cm Hn þ mH2 O/mCO þ ðm þ n=2ÞH2 (6)
Char gasification Water-Gas Reaction, C þ H2 O/CO þ H2 (þ131 MJ/kmol) (7)
Boudouard Reaction, C þ CO2 /2CO (þ172 MJ/kmol)
Methanation Methanation, C þ 2H2 /CH4 (75 MJ/kmol) (8)

Table 5
Key reactions between coal gas and iron ore.

3Fe2 O3 þ CO/2Fe3 O4 þ CO2 ; DH298


0 ¼ 43:2KJ=mol (9)
1:202Fe3 O4 þ CO/3:807Fe0:947 O þ CO2 ; DH298
0 ¼ 47:6KJ=mol (10)
Fe0:947 O þ CO/0:947Fe þ CO2 ; DH298
0 ¼ 16:7KJ=mol (11)
3Fe2 O3 þ H2 /2Fe3 O4 þ H2 O; DH298
0 ¼ 2:7KJ=mol (12)
1:202Fe3 O4 þ H2 /3:807Fe0:947 O þ H2 O; DH298
0 ¼ 88:1KJ=mol (13)
Fe0:947 O þ H2 /0:947Fe þ H2 O; DH298
0 ¼ 24:5KJ=mol (14)
0:947Fe þ H2 O/Fe0:947 O þ H2 ; DH298
0 ¼ 24:5KJ=mol (15)
3Fe0:947 O þ 0:788H2 O/0:947Fe3 O4 þ 0:788H2 ; DH298
0 ¼ 69:7KJ=mol (16)
Fe3 O4 þ 6CO/Fe3 C þ 5CO2 ; DH298
0 ¼ 411:8KJ=mol (17)
Fe3 C/3Fe þ C (18)
2CO/C þ CO2; DH298
0 ¼ 172:4KJ=mol (19)

2.3.1. Quality characteristics


In Taguchi method, signal to noise ratio (S/N ratio) is used to evaluate the quality characteristics of a product or process parameter. S/N
ratios aim to provide a healthy optimization by estimating the response (signal) variation due to uncontrollable parameters (noise) [16].
There are three types of S/N ratios defined in Taguchi method, including nominal-is-best, smaller-the-better and larger-the-better. In this
study, a better gasification performance corresponds to a higher yield of H2 and CO and higher H2/CO ratio in the gas product. Therefore, the
larger-the-better theorem was used to calculate the experimental results of the quality characteristics. The S/N ratio for larger-the-better is
defined in Equation (2).
!
S 1 Xn 1
¼ 10log i¼1 y2
(2)
N n i

where n is the number of tests and yi is the value of the ith experiment in each group.

2.3.2. Statistical analysis


Analysis of Variance (ANOVA) was used to evaluate the significance of all factors in this study, and the F-ratio method is used to
determine the error variance at a confidence level of 95%. The S/N ratio is effective to determine the optimized condition of each factor
and a set of optimized conditions for a bunch of factors. However, S/N ratio does not provide information on the contribution or impact of
each parameter on the quality characteristic which alternatively can be accomplished by analysis of variance. The percentage contribution
(also called contribution factor) is used to check the contribution of each parametric effect on the performance characteristic. The
percentage contribution was estimated by dividing the sum of square of the ANOVA model (SSt) by sum of squares of each parameter (SS)
as Equation (3).
Z. Yang et al. / Journal of the Energy Institute 92 (2019) 391e402 395

Table 6
Taguchi's L25 orthogonal design and experimental results.

Run number Level of factors Experiment results

A B C D H2 (mmol/g) CO (mmol/g) H2þCO (mmol/g) H2/CO


1 600 600 0 0 0.92 0.79 1.71 1.17
2 600 700 2:1 10 2.14 0.29 2.42 7.41
3 600 800 1:1 25 3.63 0.46 4.10 7.82
4 600 900 1:2 40 4.46 0.92 5.38 4.86
5 600 1000 1:4 55 7.14 2.37 9.50 3.02
6 700 600 2:1 25 4.86 0.21 5.07 23.07
7 700 700 1:1 40 4.67 0.50 5.17 9.33
8 700 800 1:2 55 4.64 0.72 5.36 6.47
9 700 900 1:4 0 4.39 1.64 6.03 2.67
10 700 1000 0 10 8.66 3.33 11.99 2.61
11 800 600 1:1 55 7.09 0.95 8.04 7.47
12 800 700 1:2 0 4.49 1.26 5.75 3.56
13 800 800 1:4 10 5.04 1.49 6.53 3.39
14 800 900 0 25 8.41 1.91 10.32 4.41
15 800 1000 2:1 40 16.63 0.96 17.59 17.36
16 900 600 1:2 10 6.99 1.66 8.64 4.22
17 900 700 1:4 25 8.50 2.14 10.63 3.98
18 900 800 0 40 9.12 2.21 11.33 4.12
19 900 900 2:1 55 13.83 1.14 14.97 12.16
20 900 1000 1:1 0 4.87 2.28 7.15 2.13
21 1000 600 1:4 40 9.11 2.57 11.68 3.55
22 1000 700 0 55 10.43 3.08 13.51 3.39
23 1000 800 2:1 0 1.02 0.68 1.69 1.50
24 1000 900 1:1 10 6.80 2.05 8.85 3.32
25 1000 1000 1:2 25 11.60 3.36 14.96 3.45

SS
Percentage contribution ¼ (3)
SSt

2.4. Characterization of iron ores

Powder X-ray diffraction (XRD) analysis was performed in a PANalytical/Philips X'pert Pro diffractometer (Netherland) with Cu-K alpha
radiation. The diffractometer was run at 40 KV and 40 mA. The 2q scan range of 10 e90 was used for all samples with a scan step size of
0.017. The iron ore powder was evenly distributed in a 10 mm  10 mm glass holder. XRD data was analyzed using X'pert HighScore
software package and standard PDF-2004 card. Based on published work [9,25,28e30], the major phases of iron during the reaction of iron
ore and tar include Fe2O3, Fe3O4, FeO, Fe, Fe2C, Fe3C, and C (Table 3). Semi quantitative analysis of the identified phases was performed using
the reference intensity ratio (RIR) calculation function in the software package. The morphology of iron ore samples was analyzed with a
Scanning Electron Micrograph (SEM, JSM-5610LV, JEOL, Japan) operating at 20 kV.

3. Results and discussion

3.1. Effect of temperature on gaseous products

Coal steam gasification occurs via four main steps: (1) coal pyrolysis, where coal decomposes into char, volatiles and gases; (2) secondary
pyrolysis of volatiles, where large molecules further degrade into smaller compounds and light gases; (3) steam reforming of volatiles,
where volatiles/gases are converted into CO and H2; and (4) char gasification, where char reacts with gasification agents, such as steam, CO2
and H2 to produce more gaseous products. The effects of steam concentration and temperature on non-catalytic coal gasification were
investigated in continuous feeding mode. As expected, the product gas mainly comprised H2, CO, CO2, CH4 and trace amount of C2þ. The
variation of gas yields and compositions versus reaction temperatures under N2 atmosphere and 10 vol.% steam are shown in Fig. 2 (a) and
(b), respectively. It was found that temperature had a profound effect on H2 and CO yields of coal. At the lowest reaction temperature
(600  C), the H2 yields obtained under N2 atmosphere and steam gasification were 2 mmol/g coal and 2.5 mmol/g coal, respectively, and
these values increased up to 12 mmol/g coal and 18 mmol/g coal, respectively, as the reaction temperature increased to 900  C. A CO yield of
2.0 mmol/g coal was observed under both N2 atmosphere and steam gasification at 600  C. The CO yield further increased to 6.0 mmol/g coal
and 8.0 mmol/g coal at N2 atmosphere and steam gasification at 900  C, respectively. However, the change of CH4 and CO2 yields with
temperature at both N2 atmosphere and steam gasification were negligible. The gas volume fractions at different gasification temperatures
is shown in Fig. 2(b). The H2 concentration showed an increasing trend whereas the concentrations of CH4 and CO2 showed a decreasing
trend with increasing temperature from 600  C to 900  C. The CO concentration, however, did not show any significant variation with the
increase in temperature. Coal steam gasification is a complex process which incorporates multiple steps, including coal pyrolysis, steam
reforming of volatiles and char gasification. From the major coal gasification reactions summarized in Table 4 [33,34], it can be seen that the
generation of H2 and CO were mainly attributed to endothermic reactions like water-gas reaction and steam reforming of methane, whereas
the generation of CH4 and CO2 were mainly attributed to exothermic reactions, such as water-gas shift and methanation. Therefore, a higher
gasification temperature favors the production of H2 and CO while it inhibits the production of CH4 and CO2.
396 Z. Yang et al. / Journal of the Energy Institute 92 (2019) 391e402

0.9
20
H2(10%) H2(0%) H2(10%) H2(0%)
CH4(10%) CH4(0%)

Gas composition (vol, N2 free basis)


CH4(10%) CH4(0%)
CO(10%) CO(0%) CO(10%) CO(0%)
15 CO2(10%) CO2(0%)
CO2(10%) CO2(0%)
Gas yield (mmol/g)

0.6

10

0.3

0 0.0
600 650 700 750 800 850 900 600 650 700 750 800 850 900
o
Temperature ( C)o Temperature ( C)

(a) (b)
Fig. 2. Effects of temperature on (a) gas yields and (b) gas compositions of non-catalytic gasification of MD with steam.

3.2. Effect of steam concentration on gaseous products

The gas yield obtained from gasification at 700  C (Fig. 3(a).) indicated that as the steam concentration increased from 10 vol.% to 50 vol.%,
the yield of H2 increased from 8.13 mmol/g to a maximum yield of 16.71 mmol/g at 40 vol.% and then decreased to 13.92 mmol/g when steam
concentration further increased to 50 vol.%. CO yield increased from 3.62 mmol/g (10 vol.%) to 7.63 mmol/g (30 vol.%), after which it
remained constant with further increase in steam concentration. The yields of CH4 and CO2 also increased slightly as the steam concen-
tration increased. As indicated in Fig. 3(b), higher gasification temperature (800  C) favored water-gas reaction and steam CH4 reforming
reactions, leading to the boost of H2 yield from 10.39 mmol/g (10 vol.%) to 23.10 mmol/g (50 vol.%). The CO yield peaked at 40 vol.% steam
(11.36 mmol/g) and then decreased as the steam concentration further increased to 50 vol.%. Interestingly, CO yield followed the same
pattern as H2 yield until the steam concentration reached 40 vol.%, where excess steam was consumed via water-gas shift reaction, resulting
in the consumption of CO and the production of H2. This indicates that a critical steam concentration occurred near 40 vol.%. The gas
composition as shown in Fig. 3(c) and (d), however, did not exhibit an obvious increase or decrease as the steam concentration varied,
indicating that the gas composition was not sensitive to a change in steam concentration. Since gas composition and gas yield reflect the
state of chemical equilibrium and reaction rate, respectively, the above results indicate that the increase of steam concentration only
affected reaction rate of coal gasification but barely affected the state of chemical equilibrium.

3.3. Effects of iron ore on gaseous products

Iron ore was added as a catalyst to observe its catalytic effects on gasification of coal at selected conditions (600  Ce1000  C at
50 vol.% steam and 0 vol.%e50 vol.% steam at 800  C). The gas yields as a function of steam concentration are shown at Fig. 4(a). The
addition of iron ore substantially promoted the H2 yield. For example, the H2 yield obtained under N2 atmosphere increased from
10.48 mmol/g to 13.49 mmol/g when iron ore was added. In addition, the H2 yield obtained under 40 vol.% steam concentration increased
from 23.10 mmol/g to 33 mmol/g. It is also worth noting that an inflection point occurred near 20 vol.% steam concentration, as indicated
by the sharp growth of H2 yield prior to 20 vol.% steam concentration and slow growth after 20 vol.% steam concentration. The increased
H2 yield indicates that iron ore exhibited catalytic effects on steam reforming of tar and steam gasification of char, as well as on dry
reforming of tar. In non-catalytic gasification (Fig. 3(b)), the CO yield was always higher than CO2 yield, suggesting that water-gas shift
reaction occurred slowly towards positive direction. However, the addition of iron ore resulted in a higher CO2 yield than CO yield under
all steam concentrations. In addition, the CO2 yields increased significantly while the CO decreased slightly as the steam concentration
increased under catalytic gasification. The trends for CO2 and CO yields from catalytic gasification can be attributed to the catalytic
activity of iron ore towards water-gas shift reaction. Iron-based catalysts have been widely used for their high activities in catalyzing
water-gas shift reaction and Fischer-Tropsch synthesis [35,36]. In addition, CO also acts as a reducing agent in iron ore reduction via
reactions summarized in Table 5 [25,37].
Fig. 4(b) showed the effect of temperature on different gas yields during catalytic gasification at a steam concentration of 50 vol.%. The H2
yield initially decreased as the temperature increased, with the lowest H2 yield of 26.5 mmol/g at 800  C. Thereafter the H2 yield increased
significantly as temperature increased. The overall H2 formation rate during catalytic steam gasification of coal was primarily determined by
the rates of three reactions: k1, positive rate of H2 formation during steam gasification of char, (2) k2, positive rate of H2 formation during
steam reforming of tar, and (3) k3, negative rate of H2 consumption during reduction of Fe2O3 by H2. The minimum value of H2 yield at a
temperature of 800  C implies that the overall rate increase of k1 þ k2 was lower than the rate increase of k3 at 700  C and 800  C. The CO
yield, which exhibited an opposite trend to H2 yield, reached a maximum of 2.64 mmol/g at 800  C and then decreased when temperature
further increased to 1000  C. This may be due to the balance of the rate of CO-producing (water gas reaction) and CO-consuming (reduction
of Fe2O3 by CO) reactions, the former exceeding the latter at 700  Ce800  C. The CO2 increased with increasing temperature as a result of
enhanced rate of Fe2O3 reduction by CO.
Z. Yang et al. / Journal of the Energy Institute 92 (2019) 391e402 397

Fig. 3. Effects of steam concentration on gas yields and compositions of non-catalytic MD steam gasification. (a) (c), gasified at 700  C (b) (d), gasified at 800  C.

Fig. 4. Gas yields of catalytic coal steam gasification as a function of (a) steam concentration (800  C) and (b) temperature (50 vol.% steam concentration).
398 Z. Yang et al. / Journal of the Energy Institute 92 (2019) 391e402

3.4. Effects of control factors on catalytic steam gasification

A Taguchi design based on L25 (54) orthogonal array was applied to investigate the effects of key operation parameters which can
potentially affect coal gasification performance, including upper zone temperature, lower zone temperature, steam concentration, and
catalyst loading. The combination of levels of factors and experimental results are summarized in Table 6. The maximum yield of H2 was
16.63 mmol/g, which occurred at Run 15, whereas the minimum yield of H2 (0.92 mmol/g) was observed at Run 1. In comparison, the yield of
CO varied from 0.21 mmol/g (Run 6) to 3.36 mmol/g (Run 25). The difference in observed responses indicated that performance of steam
reforming of coal was significantly affected by the combination of levels of factors.

3.4.1. S/N ratio


The mean value of S/N ratio at different levels of factors for H2 yield, CO yield, H2þCO yield, and H2/CO are shown in Fig. 5. As mentioned
earlier, a larger S/N ratio is a better quality characteristic for steam reforming of coal. The level with the highest S/N ratio represents an
optimal condition for the associated factor. Based on the results of S/N ratios, the optimum conditions for maximizing H2 yield, CO yield,
H2þCO yield and H2/CO can be represented by a combination of A (level 4) B (level 5) C (level 2) D (level 5), A (level 5) B (level 5) C (level 1) D
(level 5), A (level 4) B (level 5) C (level 2) D (level 5), and A (level 2) B (level 2) C (level 5) D (level 4), respectively.
The S/N ratio of increase in yield of H2 and CO increased with increasing upper zone temperature, suggesting that higher temperature
thermodynamically favors the coal pyrolysis (Reaction 1), secondary cracking of volatiles (Reaction 2) and steam reforming of volatiles
(Reaction 3). Alternatively, the S/N ratio plot for both H2 and CO yield presented an “N” shape with two inflection points at 700  C and 800  C
when lower zone temperature (reaction temperature of coal gas and iron ore) increased. This suggests that H2 and CO production during
catalytic coal gasification with iron ore is a balance of exothermic and endothermic reactions. The exothermic reaction (such as oxidization
of metal iron in the presence of steam (Reaction 15) and reduction of Fe0.945O into metal iron (Reaction 16)) favors the generation of H2 and
CO at lower equilibrium temperature, whereas higher equilibrium temperature prohibits the consumption of CO, thus promoting the
concentration of CO at equilibrium. As shown in Fig. 5(a) and (b), the S/N ratio of H2 and CO yield with respect to iron to coal ratio exhibited a
sharp increase at the iron to coal ratio of 1:4, followed by a steady increase trend as the iron to coal ratio increase to 2:1. The H2 and CO which

Fig. 5. Profiles of S/N ratio for (a) H2 yield (mmol/g), (b) CO yield (mmol/g), (c) H2þCO yield (mmol/g), (d) H2 yield/CO yield.
Z. Yang et al. / Journal of the Energy Institute 92 (2019) 391e402 399

was produced from coal steam gasification can be consumed as a reducing agent in the reactions with iron ore through reactions 9e14 and
17. As the fraction of iron in the mixture further increased, the quantity of H2 and CO derived from coal gasification decreased, meaning that
the availability of H2 and CO in the subsequent reduction/oxidation reactions with iron ore also decreased. The increasing trend toward iron
to coal ratio shown in Fig. 5(d) indicated that the consumption of CO was faster than that of H2. These results suggested that the addition of
iron ore was beneficial to improve the H2 concentration in the produced gas.
The significance of each factor was determined by calculating its effect value, that is, the difference between the maximum and minimum
level values, and the results are shown in the column plot in Fig. 6. Higher effect values indicate more pronounced impact of the associated
factors on experimental results. As for H2 yield, steam concentration presented the highest effect value (10.30), whereas iron to coal ratio
showed the lowest effect value (2.80). This indicated that steam concentration was the most significant operation parameter in determining
H2 yield, whereas iron to coal ratio played an insignificant role in H2 production. On the contrary, the highest effect value of 11.51 and the
lowest effect value of 2.53 were observed for iron ore to coal ratio and steam concentration, respectively in the case of CO yield, implying
that the impacts of steam concentration and iron to coal ratio on CO production were opposite to those on H2 production. As for H2þCO
yield, the effect value of factor C (3.72) was the lowest among all four factors, whereas the effect values of the other three factors were very
similar. This suggested that iron to coal ratio was the most insignificant factor in improving total gas yield. As for H2/CO, lower zone
temperature presented the lowest effect value (2.08), followed by factor upper zone temperature (6.69), iron ore to carbon ratio (9.84) and
steam concentration (10.04). This reflects that the lower zone temperature and steam concentration are the least and most influencing
factor, respectively, in determining H2/CO ratio during steam reforming of coal.

3.4.2. ANOVA
ANOVA was employed to evaluate the significance of operation parameters on performance of steam reforming of coal, and the results
were listed in Table 7. Percentage contributions (Percent P (%)) estimated the significance of the operation parameters on experiment results
of gasification. The error term refers to the errors that are not caused by main factors but by uncontrollable factors during the experiment.
Typically, the percentage of error should not exceed 50%, to ensure that the experiment results are trustable [38]. In this study, the estimated
percentage contribution of error for (a) H2 yield, (b) CO yield, (c) H2þCO yield, (d) H2 yield/CO yield was 12.44%, 3.33%, 9.97% and 10.82%,
respectively, implying that the experiment results were reliable. P-value (0.05) of main factors indicated that iron to coal ratio (P-
value ¼ .56) was not significant in production of H2, while steam concentration (P-value ¼ .24) was not significant in production of CO.
Interestingly, results of percentage contribution reflected that steam concentration with a contribution factor of 32.3% and iron to coal ratio
with a contribution factor of 38.62% were the most significant dominant factor affecting the production of H2 and CO, respectively. Iron to
coal ratio was also an insignificant factor for H2þCO yield. P-values of gasification and lower zone temperature were 0.089 and 0.52,
respectively, showing that both parameters were insignificant to H2/CO ratio. The percentage contribution of iron to coal ratio (46.05%) was
the highest among all the operation parameters, indicating that iron to coal ratio was the most influential parameter in determining the
value of H2/CO. These results matched closely with the S/N ratio results from Taguchi method.

3.5. Characterization of iron ores

3.5.1. Chemical composition of spent iron ores


In an earlier discussion, it was found that yields of H2 and CO were correlated to the form of iron within the iron ore sample. The resulting
iron ore samples from section 3.2 were subject to XRD analysis, and the main crystalline species detected were Fe3O4, FeO, Fe, Fe2C, Fe3C and
C. The relative content of these phases based on semi quantification are shown in Fig. 7. As the steam concentration increased from 10 vol.%
to 50 vol.%, the relative content of Fe3O4 increased from 30% to 95%, and as a result, the relative content of other forms of iron (Fe and FeO)
decreased significantly. This fact further confirmed that increasing steam concentration intensified the oxidation of iron compounds with
low valence to those with high valence (Reaction 15 and 16), resulting in an increase in H2 yield. In the meantime, carbon deposition was also
eliminated due to enhanced reforming reactions with steam. The effects of reaction temperature on composition of iron ore are shown in

Fig. 6. Significance of steam reforming parameters.


400 Z. Yang et al. / Journal of the Energy Institute 92 (2019) 391e402

Table 7
ANOVA for experiment results of steam reforming of coal.

Factors DOF Sum of Squares (SS) Variance (V) F-ratio (F) P-value Percent P (%)
(a) H2 yield (mmol/g)
Upper zone temperature ( C) 4 92.53 23.13 4.4 .036 27.35
Lower zone temperature ( C) 4 77.58 19.39 3.69 .055 22.93
Iron ore to coal ratio 4 16.81 4.20 0.8 .56 4.96
Steam concentration (%) 4 109.29 27.32 5.19 .023 32.30
Error 8 42.1 5.26 12.44
Total 24 338.31 100
(b) CO yield (mmol/g)
Upper zone temperature ( C) 4 6.09 1.52 17.18 .001 28.65
Lower zone temperature ( C) 4 5.63 1.40 15.89 .001 26.50
Iron ore to coal ratio 4 8.21 2.05 23.16 0 38.62
Steam concentration (%) 4 0.61 0.15 1.73 .24 2.87
Error 8 0.70 0.088 3.33
Total 24 21.27 100
(c) H2þCO yield (mmol/g)
Upper zone temperature ( C) 4 131.15 32.78 5.97 .016 29.75
Lower zone temperature ( C) 4 123.31 30.82 5.61 .019 27.97
Iron ore to coal ratio 4 26.28 6.56 1.2 .38 5.96
Steam concentration (%) 4 116.06 29.01 5.28 .022 26.33
Error 8 43.96 5.49 9.97
Total 24 440.76 100
(d) H2 yield/CO yield
Upper zone temperature ( C) 4 99.79 24.94 2.97 .089 16.08
Lower zone temperature ( C) 4 29.27 7.31 0.87 .52 4.71
Iron ore to coal ratio 4 285.77 71.44 8.51 .006 46.05
Steam concentration (%) 4 138.52 34.63 4.12 .042 22.32
Error 8 67.19 8.39 10.82
Total 24 620.54 100

Fig. 7. Compositions of iron ore obtained at (a): different steam concentration (Reaction conditions: 900  C) and (b): at different reaction temperatures (Reaction conditions: steam
concentration: 50%).

Fig. 7(b). The content of Fe3O4 first increased with increasing reaction temperature with a maximum of 95% at 900  C, and then dropped as
the reaction temperature increased to 1000  C. It is expected that increase in reaction temperature increased the reaction rates of oxi-
dization of iron with lower valence to Fe3O4 through reactions 15 and 16. However, high temperature might also lead to the destruction of
the backbones of iron ore, blocking the channels where oxidization reactions of iron species with steam occur.

3.5.2. Surface morphology of spent iron ores


Morphologies of fresh and spent iron ores produced at different temperatures and steam concentrations were analyzed with SEM and
the results are shown in Fig. 8. The fresh iron ores exhibited a non-porous structure with a clear profile (Fig. 8(a)). The spent iron ore
obtained at 700  C (Fig. 8(b)) demonstrated a porous structure where macro pores randomly scattered throughout the iron ore particles. As
the reaction temperature increased to 1000  C, the porous structure was further developed and more macrospores were observed than at
700  C. Micro morphologies of spent iron ores obtained from reacting under different steam concentrations (10 vol.%, 30 vol.% and 50 vol.%)
are shown in Fig. 8(d)e(f). At low steam concentration (10 vol.%), iron ores exhibited a compact surface structure. As the steam concen-
tration increased, iron ores surface became more porous and loose. The arrangement of pores and surface structure of iron ores may be
Z. Yang et al. / Journal of the Energy Institute 92 (2019) 391e402 401

Fig. 8. SEM of fresh and spent iron ores. (a) fresh iron ore, (b) (c) iron ore produced at 50 vol.% steam concentration and 700  C and 1000  C, respectively, (d) (e) (f) iron ore
produced at 900  C and 10 vol.%, 30 vol.% and 50 vol.% steam concentration, respectively.

associated with the removal of carbon deposits via carbon-steam reaction, making the iron ore surface more porous. This observation is
consistent with the XRD analysis of the iron ores as shown in Fig. 7(a).

4. Conclusions

The addition of iron ore significantly promoted the H2 yield during the catalytic steam gasification of coal. Taguchi method was employed
to investigate the effects of lower zone temperature, upper zone temperature, steam concentration and iron to coal ratio on yields of H2 and
CO. The analysis of signal-to-noise ratios suggest that steam concentration and iron to coal ratio was the determining factor of the yield of H2
and CO, respectively. The increase in steam concentration enhanced the oxidization of low valence iron to Fe3O4. The formation of Fe3O4 was
favored at reaction temperatures ranging from 600  C to 900  C. Further increase in reaction temperature led to the destruction of porous
structures in iron ore, thus inhibiting the oxidization of low valence iron to Fe3O4.

Acknowledgments

The authors wish to express their sincere thanks to the financial support from the National Natural Science Foundation of China
(51622604 and 51506071). The experiment was also assisted by Analytical and Testing Center in Huazhong University of Science & Tech-
nology (http://atc.hust.edu.cn), Wuhan 430074 China. The authors also appreciate Gayatri Yadavalli for language refining of this paper.

References

[1] P. Chiesa, S. Consonni, T. Kreutz, R. Williams, Co-production of hydrogen, electricity and CO2 from coal with commercially ready technology. Part A: performance and
emissions, Int. J. Hydrogen Energy 30 (2005) 747e767.
[2] J. Wang, K. Sakanishi, I. Saito, T. Takarada, K. Morishita, High-yield hydrogen production by steam gasification of HyperCoal (ash-free coal extract) with potassium
carbonate: comparison with raw coal, Energy Fuel 19 (2005) 2114e2120.
[3] N. Howaniec, A. Smolinski, Steam gasification of energy crops of high cultivation potential in Poland to hydrogen-rich gas, Int. J. Hydrogen Energy 36 (2011) 2038e2043.
[4] Y.H. Xiao, S.P. Xu, Y. Tursun, C. Wang, G.Y. Wang, Catalytic steam gasification of lignite for hydrogen-rich gas production in a decoupled triple bed reaction system, Fuel
189 (2017) 57e65.
[5] J. Wang, B. Xiao, S. Liu, Z. Hu, P. He, D. Guo, M. Hu, F. Qi, S. Luo, Catalytic steam gasification of pig compost for hydrogen-rich gas production in a fixed bed reactor,
Bioresour. Technol. 133 (2013) 127e133.
[6] S. Porada, A. Rozwadowski, K. Zubek, Studies of catalytic coal gasification with steam, Pol. J. Chem. Technol. 18 (2016) 97e102.
[7] J. Kopyscinski, R. Habibi, C.A. Mims, J.M. Hill, K2CO3-Catalyzed CO2 gasification of ash-free coal: kinetic study, Energy Fuel 27 (2013) 4875e4883.
[8] J. Wang, Y. Yao, J. Cao, M. Jiang, Enhanced catalysis of K2CO3 for steam gasification of coal char by using Ca(OH)2 in char preparation, Fuel 89 (2010) 310e317.
[9] S. Kudo, K. Sugiyama, K. Norinaga, C.Z. Li, T. Akiyama, J. Hayashi, Coproduction of clean syngas and iron from woody biomass and natural goethite ore, Fuel 103 (2013)
64e72.
[10] I.G. Lee, A. Nowacka, C.H. Yuan, S.J. Park, J.B. Yang, Hydrogen production by supercritical water gasification of valine over Ni/activated charcoal catalyst modified with Y,
Pt, and Pd, Int. J. Hydrogen Energy 40 (2015) 12078e12087.
[11] J. Tang, J. Wang, Catalytic steam gasification of coal char with alkali carbonates: a study on their synergic effects with calcium hydroxide, Fuel Process. Technol. 142
(2016) 34e41.
[12] G. Guan, G. Chena, Y. Kasai, E.W.C. Lim, X. Hao, M. Kaewpanha, A. Abuliti, C. Fushimi, A. Tsutsumi, Catalytic steam reforming of biomass tar over iron- or nickel-based
catalyst supported on calcined scallop shell, Appl. Catal. B Environ. 115e116 (2012) 159e168.
402 Z. Yang et al. / Journal of the Energy Institute 92 (2019) 391e402

[13] L. Wang, Y. Hisada, M. Koike, D. Li, H. Watanabe, Y. Nakagawa, K. Tomishige, Catalyst property of CoeFe alloy particles in the steam reforming of biomass tar and toluene,
Appl. Catal. B Environ. 121e122 (2012) 95e104.
[14] G.Q. Guan, M. Kaewpanha, X.G. Hao, A. Abudula, Catalytic steam reforming of biomass tar: prospects and challenges, Renew. Sustain. Energy Rev. 58 (2016) 450e461.
[15] M. Koike, D. Li, H. Watanabe, Y. Nakagawa, K. Tomishige, Comparative study on steam reforming of model aromatic compounds of biomass tar over Ni and Ni-Fe alloy
nanoparticles, Appl. Catal. A General 506 (2015) 151e162.
[16] C. Parsland, A.C. Larsson, P. Benito, G. Fornasari, J. Brandin, Nickel-substituted bariumhexaaluminates as novel catalysts in steam reforming of tars, Fuel Process. Technol.
140 (2015) 1e11.
[17] M.A. Uddin, H. Tsuda, S.J. Wu, E. Sasaoka, Catalytic decomposition of biomass tars with iron oxide catalysts, Fuel 87 (2008) 451e459.
[18] Y.W. Lu, Q.G. Yan, J. Han, B.B. Cao, J. Street, F. Yu, Fischer-Tropsch synthesis of olefin-rich liquid hydrocarbons from biomass-derived syngas over carbon-encapsulated
iron carbide/iron nanoparticles catalyst, Fuel 193 (2017) 369e384.
[19] T. Nordgreen, V. Nemanova, K. Engvall, K. Sjostrom, Iron-based materials as tar depletion catalysts in biomass gasification: dependency on oxygen potential, Fuel 95
(2012) 71e78.
[20] G. Wienhofer, F.A. Westerhaus, K. Junge, M. Beller, Fast and selective iron-catalyzed transfer hydrogenations of aldehydes, J. Organomet. Chem. 744 (2013) 156e159.
[21] B. Sun, K. Xu, L. Nguyen, M.H. Qiao, F. Tao, Preparation and catalysis of carbon-supported iron catalysts for Fischer-Tropsch synthesis, ChemCatChem 4 (2012) 1498e1511.
[22] Y. Tursun, J.G. Liu, S.P. Xu, L.G. Wei, W.J. Zou, Catalytic steam gasification of lignite with olivine as solid heat carrier, Fuel 112 (2013) 641e645.
[23] R. Michel, S. Rapagna, M. Di Marcello, P. Burg, M. Matt, C. Courson, R. Gruber, Catalytic steam gasification of Miscanthus X giganteus in fluidised bed reactor on olivine
based catalysts, Fuel Process. Technol. 92 (2011) 1169e1177.
[24] N.F. Othman, M.H. Bosrooh, Catalytic adaro coal gasification using dolomite and nickel as catalysts, in: Proceeding of 4th International Conference on Process Engi-
neering and Advanced Materials (Icpeam 2016), vol. 148, 2016, pp. 308e313.
[25] R.B. Cahyono, A.N. Rozhan, N. Yasuda, T. Nomura, S. Hosokai, Y. Kashiwaya, T. Akiyama, Catalytic coal-tar decomposition to enhance reactivity of low-grade iron ore, Fuel
Process. Technol. 113 (2013) 84e89.
[26] X.H. Zou, T.H. Chen, H.B. Liu, P. Zhang, D. Chen, C.Z. Zhu, Catalytic cracking of toluene over hematite derived from thermally treated natural limonite, Fuel 177 (2016)
180e189.
[27] F. Lind, M. Seemann, H. Thunmani, Continuous catalytic tar reforming of biomass derived raw gas with simultaneous catalyst regeneration, Ind. Eng. Chem. Res. 50
(2011) 11553e11562.
[28] Y. Mochizuki, N. Tsubouchi, T. Akiyama, Reduction behavior and crushing strength of carbon-containing iron ore sinters prepared from tar recovered from coke oven gas,
Fuel Process. Technol. 138 (2015) 704e713.
[29] R.B. Cahyono, A.N. Rozhan, N. Yasuda, T. Nomura, H. Purwanto, T. Akiyama, Carbon deposition using various solid fuels for ironmaking applications, Energy Fuel 27
(2013) 2687e2692.
[30] Y. Hata, H. Purwanto, S. Hosokai, J. Hayashi, Y. Kashiwaya, T. Akiyama, Biotar ironmaking using wooden biomass and nanoporous iron ore, Energy Fuel 23 (2009)
1128e1131.
[31] J. Hu, Y. Chen, K. Qian, Z. Yang, H. Yang, Y. Li, H. Chen, Evolution of char structure during mengdong coal pyrolysis: influence of temperature and K2CO3, Fuel Process.
Technol. 159 (2017) 178e186.
[32] D.D. Yao, C.F. Wu, H.P. Yang, Q. Hu, M.A. Nahil, H.P. Chen, P.T. Williams, Hydrogen production from catalytic reforming of the aqueous fraction of pyrolysis bio-oil with
modified Ni-Al catalysts, Int. J. Hydrogen Energy 39 (2014) 14642e14652.
[33] W.P. Walawender, D.A. Hoveland, L.T. Fan, Steam gasification of pure cellulose. 1. Uniform temperature profile, Ind. Eng. Chem. Process Des. Dev. 24 (1985) 813e817.
[34] M.J. Antal, W.H. Edwards, H.L. Friedman, F.E. Rogers, A study of the steam gasification of organic wastes, in: Project Report to EPA, Liberick W W, Project Officer,
Cincinnati, OH 45268, 1984.
[35] Y. Sekine, T. Chihara, R. Watanabe, Y. Sakamoto, M. Matsukata, E. Kikuchi, Effect of loading potassium and palladium over iron-based catalyst for low temperature water-
gas shift reaction, Catal. Lett. 140 (2010) 184e188.
[36] A.N. Pour, M.R. Housaindokht, S.F. Tayyari, J. Zarkesh, S.M.K. Shahri, Water-gas-shift kinetics over a Fe/Cu/La/Si catalyst in Fischer-Tropsch synthesis, Chem. Eng. Res. Des.
89 (2011) 262e269.
[37] M.F. Bleeker, S.R.A. Kersten, H.J. Veringa, Pure hydrogen from pyrolysis oil using the steam-iron process, Catal. Today 127 (2007) 278e290.
[38] M.H. Shahavi, M. Hosseini, M. Jahanshahi, R.L. Meyer, G.N. Darzi, Clove oil nanoemulsion as an effective antibacterial agent: Taguchi optimization method, Desalination
Water Treat. (2015) 1e12.

View publication stats

You might also like