Sei sulla pagina 1di 37

Subscriber access provided by University at Buffalo Libraries

Article
A two-dimensional CFD simulation of biomass gasification in a
downdraft fixed bed gasifier with highly preheated air and steam
Yueshi Wu, Qinglin Zhang, Weihong Yang, and Wlodzimierz Blasiak
Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/ef4003704 • Publication Date (Web): 02 May 2013
Downloaded from http://pubs.acs.org on May 12, 2013

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts
appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered
to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just
Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Energy & Fuels is published by the American Chemical Society. 1155 Sixteenth Street
N.W., Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 36 Energy & Fuels

1
2
3
4
5
6
7 A two-dimensional CFD simulation of biomass
8
9
10
11 gasification in a downdraft fixed bed gasifier with
12
13
14
15 highly preheated air and steam
16
17
18
19
Yueshi Wu,*,a Qinglin Zhang, b Weihong Yang, a and Wlodzimierz Blasiak a
20
21 a
22 KTH Royal Institute of Technology, School of Industrial Engineering and Management,
23
24 Department of Materials Science and Engineering, Division of Energy and Furnace
25
26 Technology, SE 100 44 Stockholm, Sweden
27
28 b
Plasco Energy Group Inc, 1000 Innovation Drive, Suite 400, Ottawa, Canada
29
30
31 *Corresponding author: Tel. +46 8 790 90 22; E-mail address: yueshiw@kth.se
32
33
34 Abstract
35
36
37 Biomass gasification is regarded as one of the most promising energy recovery
38
39 technologies for the widespread utilization of biomass. Mathematical models have been
40
41
42
developed to understand this process in downdraft fixed beds using zero-dimensional and
43
44 one-dimensional models, but only a limited number of two-dimensional models for downdraft
45
46 fixed-bed reactors can be found in the literature. In this study, a two-dimensional (2D)
47
48 computational fluid dynamics (CFD) model was developed to study the gasification process
49
50
in a downdraft configuration, considering drying, pyrolysis, combustion and the gasification
51
52
53 reactions. The gas and solid phases were resolved using an Euler-Euler multiphase approach
54
55 with exchange terms for the momentum, mass and energy. The standard k- turbulence
56
57 model was used in the gas phase. The model results were compared with existing data from a
58
59
60 1
ACS Paragon Plus Environment
Energy & Fuels Page 2 of 36

1
2
3 demonstration-scale fixed bed downdraft gasifier. The simulation results exhibit a reasonable
4
5 agreement with the experimental data. Parameter studies were performed based on the
6
7 developed model, which indicated that an external heat source for the High Temperature
8
9
10 Agent Gasification (HTAG) technology using super-heated air combined with steam resulted
11
12 in a limited combustion need in the gasifier and produced syngas with a high H2 fraction and
13
14 low tar content, which is environmentally preferable.
15
16
17
18
Keywords: biomass; high temperature gasification; downdraft fixed bed; 2D CFD model;
19
20
21 multiphase flow.
22
23
24
25 1 Introduction
26
27
28
Biomass is currently the largest source of renewable energy, accounting for
29
30 approximately 10-15% (or 45 ± 10 EJ)[1] of the world’s total energy supply. Markets for
31
32
33
biomass heat appliances have enjoyed healthy growth in recent years, particularly in Europe.
34
35 Gasification is regarded as one of the most promising energy recovery technologies for the
36
37 widespread utilization of biomass. A recent advance in developing countries is the
38
39 introduction of small-scale gasifier stoves for cooking[2]. The downdraft gasifier is a
40
41
comparatively cheap reactor that can be used in small-scale projects for heat and power.
42
43
44 Biomass gasifiers are complex facilities requiring time to be mounted and to be made
45
46 operational, which makes it difficult to study their various working conditions[3]. For such an
47
48 investigation, numerous models for downdraft fixed bed gasifiers have been developed.
49
50 These models of biomass gasification in downdraft fixed bed reactors can be categorized
51
52
53
into two groups: (1) thermodynamic equilibrium models and (2) kinetic or non-equilibrium
54
55 models. The thermodynamic equilibrium models, the so-called zero-dimensional (0D)
56
57 models, are widely used among researchers[4-8] to predict the composition of the produced
58
59
60 2
ACS Paragon Plus Environment
Page 3 of 36 Energy & Fuels

1
2
3 syngas and the equilibrium temperature by assuming that the chemical reactions reach
4
5 equilibrium, but these models cannot provide highly accurate results and also cannot provide
6
7 the concentration or temperature profiles inside the reactor as lack of the transformation
8
9
10 mechanisms. Because this approach is independent of the gasifier design, kinetic models,
11
12 which take into account the reaction kinetics and the transfer phenomena among the phases,
13
14 must be developed. This need spurred the development of one-dimensional (1D) biomass
15
16 gasification models [9-12]. These models simulated the variations in the physical and
17
18
chemical properties along the reactor height by considering the vertical movements. To
19
20
21 further this research, 2D models are needed to comprehensively understand the gasification
22
23 process and the effects of the reactor geometry. However, few works have performed 2D
24
25 simulations of downdraft fixed bed in the available literature.
26
27 In recent years, CFD has been used as a powerful tool to build 2D or even 3D models.
28
29
30 Most of the reported CFD simulations for biomass gasification have been of entrained flow
31
32 gasifiers[13, 14]. Some researchers used the CFD approach to simulate the fast pyrolysis of
33
34 biomass in fluidized bed gasifiers[15, 16]. Zhang et al.[17] performed the plasma gasification
35
36 process of municipal solid waste using a 2D CFD model for an updraft fixed bed gasifier.
37
38
Gerun et al.[18] built a 2D axisymmetric CFD model for a two-state downdraft gasifier but
39
40
41 only considered the oxidation zone.
42
43 In this study, a 2D CFD model was developed to study the biomass gasification process
44
45 in a downdraft fixed bed configuration, taking into account drying, pyrolysis, combustion and
46
47 the gasification reactions. The gas and solid phase were resolved using an Euler-Euler
48
49
50 multiphase approach with exchange terms for the momentum, mass and energy. All of the
51
52 chemical reactions were defined within ANSYS FLUENT 14.5 using the User Defined
53
54 Function (UDF). The model results were compared with existing data from a demonstration-
55
56 scale fixed bed downdraft gasifier using highly preheated air and steam, which was built at
57
58
59
60 3
ACS Paragon Plus Environment
Energy & Fuels Page 4 of 36

1
2
3 KTH-Royal Institute of Technology. The effect of the operating parameters on the
4
5 performance of gasifier was then investigated.
6
7
8
9
10
2 Facility and feedstock
11
12
13 2.1 Experimental setup
14
15 A demonstration-scale HTAG test facility has been built at KTH. This system was
16
17
described in previous publication[19], and a brief description of gasifier is provided below.
18
19
20 The feedstock was filled from the top of a vertically cylindrical reactor using a
21
22 continuous feeding system with four synchronized screws. The gasifying agent (the air/steam
23
24 mixture) was preheated to 1303 K by a regenerative preheater that used an external fuel
25
26 source to maintain a stable operation and was then introduced into the reactor from the side.
27
28
29
The grate stopped biomass/char particles, resulting in a charcoal bed. The pyrolyzed gases
30
31 were mixed with the gases produced by combustion and passed through the grate to generate
32
33 the produced gases, which were sampled at the outlet of the gasifier. The temperatures were
34
35 measured using thermocouple probes located at the centerline along the reactor’s height in
36
37
different reaction zones. A schematic of the downdraft fixed-bed gasifier is presented in
38
39
40 Figure 1.
41
42 Figure 1. A schematic of the downdraft fixed bed gasifier in the KTH laboratory
43
44
45
46
2.2 Fuel parameters
47
48 The feedstock used for the investigation was wood pellets 0.008 m in diameter with an
49
50 average length-to-diameter ratio, l/d, of 4. The properties of the feedstock are shown in Table
51
52 1. The bulk density of the wood pellets is 630-650 kg/m3.
53
54
55 Table 1. The characterization of the biomass feedstock
56
57
58
59
60 4
ACS Paragon Plus Environment
Page 5 of 36 Energy & Fuels

1
2
3 3 Numerical model
4
5
6 This study applied the Euler-Euler multiphase flow framework. All of the species were
7
8 divided into two phases: solid and gas. The conservation equations for the mass, momentum
9
10 and energy were solved for both phases. The exchanges of momentum (given by the drag
11
12 between solid phase and gas phase), mass (given by the heterogeneous chemical reactions)
13
14
and energy were allowed between phases and were described as sources terms in the
15
16
17 conservation equations. The standard    turbulence model was used in the gas phase. A
18
19 scheme of the multiphase model is shown in Figure 2.
20
21
22 Figure 2. A schematic of the CFD model
23
24
25 3.1 Governing equations
26
27 The mass, momentum and energy equations of the gas phase are described as follows:
28
29

30 (α g ρ gYi ) + ∇ ⋅ (α g ρ gYivrg ) = m& i + Si (1)
31 ∂t
32

33
34
(α g ρ g vrg ) + ∇ ⋅ (α g ρ g vrg vrg ) = −α g ∇p + ∇ ⋅ τ g + α g ρ g gr + K sg (vrs − vrg ) + m& sg vrsg (2)
∂t
35
36
37

38 (α g ρ g hg ) + ∇ ⋅ (α g ρ g vrg hg ) = −α g ∂p + τ g : ∇vrg − ∇ ⋅ qrg + S g + Qsg + m& sg hsg (3)
39 ∂t ∂t
40
41 The standard    turbulence model[20] was used in the gas phase. The ideal gas law
42
43 was used to calculate the gas density:
44
45
M (4)
46 ρg = p
47 RgTg
48
49 The governing equations for the solid phase are described as follows:
50
51

52 (α s ρsY j ) + ∇ ⋅ (α s ρ sY j vrs ) = m& j + S j (5)
53 ∂t
54
55 ∂ (6)
56 (α s ρ s vrs ) + ∇ ⋅ (α s ρ s vrs vrs ) = −α s∇p − ∇ps + ∇ ⋅ τ s + α s ρ s gr − m& sg vrsg
57 ∂t
58
59
60 5
ACS Paragon Plus Environment
Energy & Fuels Page 6 of 36

1
2

(α s ρ s hs ) + ∇ ⋅ (α s ρ s vrs hs ) = −α s ∂p + τ s : ∇vrs − ∇ ⋅ qrs + S s + Qgs − m& sg hsg
3 (7)
4 ∂t ∂t
5
6
The flow of the solid phase was treated as a granular flow. The solid phase stress consists
7
8
9 of the solid pressure and shear stress. According to Cowin[21], the flow of granular materials
10
11 in a fixed bed gasifier can be treated as a plastic flow. The Schaeffer model[22] can be used to
12
13 describe the solid pressure, ps , which can be written as[23]:
14
15
16 ps = α s p ∗ (8)
17
18
19 p∗ is expressed by an empirical power law:
20
21
22 (
p ∗ = A α g − α g∗ )
n (9)
23

24 where α g is the minimum fluidized volume fraction of the gas phase. The empirical values of
25
26
27 A = 10 25 (Pa) and n = 10 [23] were used, which allow to be used for plastic flow regime.
28
29 As the flow of the solid phase was a dense flow, the solid volume fraction of the solid
30
31
32
phase was near the packing limit, which is 0.5. Therefore, we only considered the frictional
33
34 viscosity for the shear stress portion. The shear stress can be calculated by[24]:
35
36 ps sin φ (10)
37 µs =
2 I2D
38
39
40 In a fixed-bed gasifier, the value of ∇ps is several orders of magnitude larger than the
41
42
43
gas-solid stress[25]. Therefore, the influence of the gas-solid stress on the solid phase can be
44
45 ignored. However, we considered this stress term in the gas phase, and it was calculated using
46
47 the Ergun equation. The interphase momentum exchange coefficient, K sg , is described as:
48
49
α (1 − α g )µ g
r r (11)
50 ρ gα s vs − vg
51 K sg = 150 s + 1.756
52 α g ds 2
ds
53
54 The interphase heat transfer is assumed to be a function of the temperature difference
55
56 between the two phases:
57
58
59
60 6
ACS Paragon Plus Environment
Page 7 of 36 Energy & Fuels

1
2
3 Qsg = −Qgs = k sg (T s−Tg ) (12)
4
5
where ksg is the heat transfer coefficient between the solid phase and the gas phase.
6
7
8 Convection is the main heat transfer mechanism within the fixed-bed reactor, and ksg is
9
10
11 related with the Nusselt number and is expressed as:
12
13 6κ gα sα g Nu s (13)
14 k sg = 2
ds
15
16
17 The Nusselt number was correlated by Gunn[26], which is applicable for void fraction of
18
19 0.35 – 1.0:
20
21
22 ( 2
)(
Nus = 7 − 10α g + 5α g 1 + 0.7 Re0s .2 Prg
0.33
) + (1.33 − 2.4α g
2
)
+ 1.2α g Re0s .7 Prg
0.33 (14)
23
24
25 3.2 Reaction models
26
27 The source terms for the governing equations are determined by the rates of the chemical
28
29
30 reactions in the gasifier. The biomass particle is dried after it is heated to the moisture
31
32 vaporization temperature and is continuously heated to a temperature of approximately 673 K,
33
34 at which the biomass particle is volatilized, resulting in volatile matter, char and ash. After
35
36 that, the char particle is burnt and gasified. Meanwhile, the gas phase reacts following a series
37
38
of homogeneous chemical reactions.
39
40
41 The feedstock is considered to be a mixture of moisture, volatile matter, charcoal and ash
42
43 as determined by the proximate analysis, following the equation:
44
45 → H 2 O (l ) + Volatile + Char + Ash
Biomass  (15)
46
47
48 The composition of the volatile matter is calculated by the ultimate analysis. Charcoal is
49
50 simplified as the fixed carbon and the composition of the ash is calculated from the elemental
51
52 balance.
53
54
55
56
57
58
59
60 7
ACS Paragon Plus Environment
Energy & Fuels Page 8 of 36

1
2
3 3.2.1 Drying
4
5 The moisture in the biomass is evaporated as the high temperature gasifying agent
6
7 transfers heat to the biomass.
8
9
10 H 2 O (l ) →
rd
H 2O( g ) (R-D-1)
11
12 The drying rate is determined by the following expression, and it is assumed that the
13
14 evaporation is assumed controlled by heat transfer:
15
16
17 rd = As k sg (Ts − Tevap ) / H evap (16)
18
19 where As is the surface area of the solid particles,  ; H evap is the heat of evaporation of
20
21
22 water, which is 40.65  /; and Tevap is the saturation temperature, set to 375.15 .
23
24
25
3.2.2 Pyrolysis
26
27
28
For simplification, the pyrolysis process is divided into primary pyrolysis and tar
29
30 cracking using a two-step pyrolysis model.
31
32 Primary pyrolysis:
33
34 r
Volatile →
p1
0.268CO + 0.295CO2 + 0.094CH 4 + 0.5 H 2 + 0.255 H 2O + 0.004 NH 3 + 0.0002 H 2 S + 0.2 Primary tar (P-1)
35
36
37 Tar cracking:
38
r
39 Primary tar →
p2
0.261Secon dary tar + 2.6CO + 0.441CO 2 + 0.983 CH 4 + 2 .161 H 2 + 0 .408 C 2 H 4 (P-2)
40
41 The pyrolysis reactions are considered to be under kinetic control. Tar is a complex
42
43
mixture of hundreds of different organic species; however, the global compositions are used
44
45
46 for both the primary tar and the secondary tar in this study. The primary tar is expressed as
47
48 . . , , which can be calculated from the elements balance of the primary
49
50 pyrolysis reaction. The secondary tar is assumed to be pure benzene, as experimental tests at
51
52
KTH showed that 85% of the secondary tar was benzene[19]. The compositions of the
53
54
55 product gas from the primary pyrolysis[27] and tar cracking[28] reactions are estimated on the
56
57 basis of the literature data obtained for wood.
58
59
60 8
ACS Paragon Plus Environment
Page 9 of 36 Energy & Fuels

1
2
3 The reaction rates are calculated by [29, 30]:
4
5  − 1.527 × 10 5  (17)
6 r p1 = 4 .38 × 10 9 (1 − α g )exp  C Volatile
 RT s 
7
8
9  − 1.08 × 10 5  (18)
r p 2 = 4.28 × 10 6 α g exp  C pri − tar
10  RT g 
 
11
12
13
14 3.2.3 Char reactions
15
16 The heterogeneous char reactions are considered using the unreacted shrinking core
17
18 model. It assumes that the char particles are spherical grains and that the reaction between the
19
20
21
char particles and the gas is initiated on the external surface. The reaction will move step-by-
22
23 step to the interior, which forms an interior unreacted core and an exterior ash layer. The
24
25 chemical reaction rates ( rm ) of the char reactions are determined by the film mass transfer
26
27
28 coefficient diffusion ( sm ) and the kinetic reaction rates ( rkm ). The chemical reaction rates are
29
30 written as:
31
32
33  1  Av ρ i (19)
rm =   m = 8 − 11
34
 vi M i  1 + 1
35
sm rkm ,
36
37
38 where i represents the gaseous reactants of reaction m .
39
40 The char combustion and gasification reactions considered in this model and the kinetic
41
42
43 reaction rates are summarized in Table 2. The film mass transfer coefficient is described
44
45 as[31]:
46
47 2.06G (20)
48 sm = Re −0.575 Sc −2 / 3
49 αsρg
50
51
where G is the gas mass flux in kg / m 2 ⋅ s .
52
53
54 Table 2. The kinetic reaction rates of the char reactions
55
56
57
58
59
60 9
ACS Paragon Plus Environment
Energy & Fuels Page 10 of 36

1
2
3 3.2.4 Homogeneous reactions
4
5 The gas phase species included in this model are O2, N2, H2, CO, CO2, H2O (g), CH4,
6
7 C2H4, H2S, NH3, primary tar and secondary tar.
8
9
An overview of chemical reactions in the gas phase and their kinetic reaction rates is
10
11
12 presented in Table 3.
13
14 Table 3. The kinetic reaction rates of the homogeneous reactions
15
16 The chemical reaction rates are determined by the minimum value of the kinetic reaction
17
18 rates ( rkm ) and the turbulent mixing rates ( rtm ):
19
20
21 rm = min(rkm , rtm ) , m = 1− 7 (20)
22
23
24 The turbulent mixing rates are estimated using the Eddy Dissipation (ED) model:
25
26 ε  Y Yj  (21)
27 rtm = 4.0 ρ min i , 
k  
28  vi M i v j M j 
29
30 where i and j denote the reactants of reaction m .
31
32
33
34
35 4 Results and discussion
36
37
38 4.1 Comparison with experimental data
39
40
41 The experimental data reported in our previous publication [32] are compared with the
42
43 simulation results. The operating conditions of the experiment are taken as the boundary
44
45 conditions for the simulation. The model uses a gasifying agent inlet on the topside and a
46
47 syngas outlet at the opposite side, under the grate. The biomass was fed from the top of the
48
49
50
gasifier in the experiment. As the main study is focused on the charcoal bed, in the model the
51
52 biomass input is relocated to be 0.45 m above the grate, which is the same position as the
53
54 height of the initial charcoal bed in the experiment.
55
56
57
58
59
60 10
ACS Paragon Plus Environment
Page 11 of 36 Energy & Fuels

1
2
3 For the comparison case, the mass flow rate of the feedstock is 60 kg/h. The hot gasifying
4
5 agent is feed at a mass flow rate of 106.3 kg/h and a temperature of 1303 K, in which the
6
7 steam/air mass ratio (S/A) is 0.89. In the real process, significant heat was lost before the
8
9
10 gasifying agent entered the reactor. Therefore, the temperature T1, which was measured after
11
12 the inlet of the gasifying agent and had a value of 973 K, is taken as the preheating
13
14 temperature in the model.
15
16 Figure 3 compares the temperature distribution for the center line along the vertical axis
17
18
predicted by the model with the experimental results.
19
20
21 Figure 3. The experimental and simulated temperature distributions for the centerline along
22
23 the vertical axis inside the gasifier
24
25
26 The predicted temperature was in reasonable agreement with the testing points, especially
27
28 at three important points: after the feed gas entered reactor (T1), the top of charcoal bed (T4)
29
30
31 and under the grate (T5). As expected, a significant temperature drop can be observed
32
33 underneath the top of charcoal bed due to the heat exchange between the cold feedstock and
34
35 hot gasifying agent and also because of the endothermic drying and pyrolysis reactions. The
36
37 temperature then increased, which showed the formation of the combustion zone. The
38
39
gasification reactions are mainly endothermic reactions, which might be the main reason for
40
41
42 the temperature drop after the combustion.
43
44 Figure 4 compares the produced gas compositions predicted by the model with the
45
46 experimental values. The species compared are the main combustible products, CO2 and tar
47
48 generated during the gasification process. It can be seen that both the predicted and
49
50
51 experimental data showed similar compositions. For both, the mole fractions of H2 and CH4
52
53 are approximately 26% and 4.1%, respectively, with a tar content of 4 g/Nm3. The predicted
54
55 produced gas has a higher CO content than found in the experimental data, but is the opposite
56
57 for CO2. This result might be because the air that entered the reactor along with the feed
58
59
60 11
ACS Paragon Plus Environment
Energy & Fuels Page 12 of 36

1
2
3 biomass reacted with CO to generate CO2. However, the deviation is less that 8%, with the
4
5 CO2 fraction showing the largest deviation. Therefore, this model is acceptable to analyze the
6
7 performance of the HTAG (air/steam mixture) process in the downdraft fixed bed gasifier.
8
9
10 Figure 4. A comparison of the simulated produced gas compositions with the
11
12 experimental values
13
14
15 4.2 Parameter study
16
17 The performance of the gasifier is directly influenced by the chosen operating conditions.
18
19
20
In this study, the examined parameters were the S/A ratio and preheating temperature of the
21
22 gasifying agent. Each operating parameter was varied while keeping the other one constant.
23
24 The mass flow rates of the feedstock and air were 60 kg/h and 56.3 kg/h, respectively, for all
25
26 of the cases. The basic conditions were an S/A ratio of 0.89 and a preheating temperature
27
28 1303 K for the gasifying agent.
29
30
31
4.2.1 Analysis of the basic case
32
33
34
The basic case was studied from two points: the gas temperature distributions in the
35
36 horizontal direction and the average gas compositions profile in the vertical direction.
37
38 Figure 5 shows the gas temperature distributions for different horizontal sections in
39
40 charcoal bed. The figure shows that the temperature distribution in a horizontal section is not
41
42 uniform, which is due to the uneven fluid field as the geometry of the gas inlet and outlet are
43
44
45 asymmetrical. The peak temperatures in the different sections appear at 0.23 m< x < 0.29 m.
46
47 In this model, the walls (except for the top wall) are to be considered adiabatic. The peak
48
49 temperature most likely could be explained by the char combustion that occurs in a certain
50
51 layer near the gas-bed interface. The average temperature of y = 1.13 m is much lower than
52
53
54
the other positions. At this point, the biomass meets with the hot gas and the drying reaction
55
56 occurs. The temperature is the highest at y = 1.0 m; here, the char combustion is enhanced.
57
58 The heat from the combustion was combined with the sensible heat and carried by the hot
59
60 12
ACS Paragon Plus Environment
Page 13 of 36 Energy & Fuels

1
2
3 agent; this energy was consumed by the pyrolysis at y = 1.1 m and gasification at y = 0.7 m-
4
5 0.9 m.
6
7 Figure 5. The gas temperature distribution for different horizontal sections in the
8
9
10 charcoal bed
11
12 Figure 6 shows the profiles of the average gas composition along the vertical direction of
13
14 the gasifier. The charcoal bed was located from 0.69 m to 1.14 m. The biomass met with the
15
16 hot agent at 1.14 m. The drying reaction was begun, and the moisture content in the biomass
17
18
particle was released, which can be observed from the first peak of the steam content. The
19
20
21 drying reaction could also explain the temperature drop in this area in Figure 5. The other
22
23 peak for steam denoted where the primary pyrolysis, which has steam as a product, took
24
25 place; this can also be confirmed by the appearance of H2 and CH4.
26
27 The gasification and combustion zone could be differentiated based on the presence or
28
29
30 absence of free oxygen. Oxygen was totally assumed at the y =1 m position, which indicated
31
32 the consummation of the combustion reactions. This was verified by the occurrence of the
33
34 highest temperature at the same place in Figure 5.
35
36 In reality, gasification and combustion typically occur at the same time. Figure 6 shows
37
38
that the rate of generation of CO is faster than that of CO2 after the point of y = 1.13. Some of
39
40
41 the reactions that can be used to explain this result include R-C-3, R-C-4 and R-G-7.
42
43 When no free oxygen exists, the only reaction that consumes CH4 is R-G-7. The mole
44
45 fraction of CH4 has a peak at approximately y = 0.95, when no free oxygen existed. At this
46
47 point, two reactions related to CH4 occur. The R-G-7 reaction consumes CH4, so we can say
48
49
50 that the CH4 peak is caused by the tar cracking reaction. After all of the reactions reached
51
52 equilibrium, the composition of the gases will not change.
53
54 Figure 6. Profiles of the average gas composition along the vertical axis
55
56
57
58
59
60 13
ACS Paragon Plus Environment
Energy & Fuels Page 14 of 36

1
2
3 4.2.2 Effect of the preheating temperature
4
5 Figure 7 shows the temperature of the gas phase and the mass concentration profiles of
6
7 the tars during gasification with different preheating temperatures for the feed gas, ranging
8
9
from 1103 K to 1503 K. The peak temperature inside the charcoal bed increased as the
10
11
12 preheating temperature increased. This is because the sensible heat carried by the hot feed gas
13
14 induced an increase in the temperature of the bed. The positive impacts of the higher bed
15
16 temperature were higher chemical reaction rates and enhanced heat transfer. At the same time,
17
18 the drying zone was significantly narrowed. A consequence of the higher bed temperature was
19
20
21
a relative decrease in the biomass residence time, which made the HTAG technology more
22
23 insensitive to variabilities in the heating value and moisture content of the biomass and the
24
25 particle size.
26
27 Figure 7 demonstrates that the primary tar began to be consumed (the secondary tar
28
29
began to appear) at a higher position when the preheating temperature was increased. This
30
31
32 finding is observed because the required activation energy for the tar cracking reaction was
33
34 reached more quickly using a higher preheating temperature. Therefore, for the same
35
36 residence time of biomass, more time remained for tar cracking. However, a high temperature
37
38 is also an important factor in increasing the tar combustion reaction rate. Consequently, the tar
39
40
41 content decreased with higher feed gas preheating temperatures.
42
43 Figure 7. The gas phase parameters for different preheating temperatures: (a) temperature (K),
44
45
(b) the mass concentration of the primary tar (g/Nm3), (c) the mass concentration of the
46
47
48 secondary tar (g/Nm3).
49
50
Figure 8 shows the influence of the feed gas preheating temperature on the composition
51
52
53 of the dry produced gases. It could be observed that for the same S/A ratio (0.89), the mole
54
55 fraction of H2 continuously increased from 25.8% to 29.1% within the preheating temperature
56
57 range of 1103 K to 1503 K. The same trend can be observed for the mole fraction of CO,
58
59
60 14
ACS Paragon Plus Environment
Page 15 of 36 Energy & Fuels

1
2
3 which increased from 43.9% to 46.9%. The observed increases are results of the pyrolysis and
4
5 tar cracking reactions, which are favored by the increase in the temperature because high
6
7 temperature would promote higher reaction rates and result in a greater production of the
8
9
10 combustible components of the produced gases. As expected, the mole fraction of CO2
11
12 significantly decreased as the preheating temperature increased. The high temperature
13
14 increased the activity of the char gasification reactions (R-C-3 and R-C-4). CO2 was
15
16 consumed by the gasification reactions, with corresponding increases in the fractions of CO
17
18
and H2. The mole fraction of CH4 decreased from 3.37% to 2.9%, which was affected by the
19
20
21 increased rate of reaction R-G-7 at higher temperature. However, the variation in CH4 was
22
23 relatively small compared to those of H2, CO and CO2.
24
25 The most pronounced effect of preheating could be observed in the increase in H2, which
26
27 is calculated by the increase in the H2/CO mole ratio from 0.58 to 0.62. At the same time, the
28
29
30 H2/CO2 mole ratio increased from 0.96 to 1.38. The CO/CO2 mole ratio increased from 1.63
31
32 at 1103 K to 2.22 at 1503 K. The results clearly indicate that the more favorable conditions
33
34 for reaction R-C-4 occur at higher preheating temperatures.
35
36
37 Figure 8. The influence of the feed gas preheating temperature on the composition of the
38
39 produced gases
40
41
42 4.2.3 Effect of the S/A ratio
43
44 The heat needed for the decomposition of the biomass can be supplied by the partial
45
46 oxidation of the feedstock by supplying air or other forms of external heat. In this work, the
47
48
49
partial oxidation of the biomass and the sensible heat of the preheated steam and air mixture
50
51 provide the heat required for the decomposition of biomass. The influence of the S/A ratio on
52
53 the HTAG process was studied by varying the steam feed rate while maintaining a constant
54
55 air feed rate. Figure 9 presents the distributions of the gas phase temperature and
56
57
concentration of the tars as the S/A ratio was varied from 0.71 to 1.07. Although at higher S/A
58
59
60 15
ACS Paragon Plus Environment
Energy & Fuels Page 16 of 36

1
2
3 ratios, more sensible heat was entering the reactor via the steam, the gas temperature inside
4
5 the bed did not change significantly. As gasification with steam is a highly endothermic
6
7 process, the sensible heat carried by the gasifying agent might be consumed by the steam
8
9
10 gasification reactions under the operating conditions in this work.
11
12 The results indicate that the increase in the S/A ratio is beneficial in reducing the tar
13
14 content. In this study, the S/A ratio was changed by varying the steam feed rate while
15
16 maintaining a constant air feed rate. A portion of the required heat came from the sensible
17
18
heat carried by the additional steam when the S/A ratio was increased. As a result, the saved
19
20
21 oxygen could react with the tar, which might be the reason for the tar reduction.
22
23 Figure 9. The gas phase parameters for different S/A ratios: (a) temperature (K), (b) the mass
24
25
26 concentration of the primary tar (g/Nm3), (c) the mass concentration of the secondary tar
27
28 (g/Nm3).
29
30
31 Figure 10 shows the compositions of the gases produced using a mixture of air and steam
32
33 at different S/A ratios as the gasification agent. When the S/A ratio increased from 0.71 to
34
35 1.07, the formation of H2 increased from 25.5 % to 27.9%. This could be explained as
36
37 reactions R-C-3 and R-G-5 involve steam and might be favored by increasing the steam in the
38
39
feed gas. In this same S/A ratio range, the mole fraction of CO exhibited a steadily decreasing
40
41
42 trend from 47.8% to 44.0%. Meanwhile, the mole fraction of CO2 showed the opposite trend.
43
44 These results could be explained by reaction R-G-5. This homogeneous so-called water-gas
45
46 shift reaction is a purely gaseous reaction, which takes place in the presence of steam. The
47
48 partial pressure of water vapor increased by adding steam, which forced the reaction to the
49
50
51 products. Additionally, the combination of reaction R-G-7 with reaction R-C-3 could also
52
53 explain the formation of CO. The mole fraction of CH4 exhibited an irregular trend as the
54
55 S/A ratio was varied and reached a maximum value of 3.2% when the S/A ratio was 0.89.
56
57
58
59
60 16
ACS Paragon Plus Environment
Page 17 of 36 Energy & Fuels

1
2
3 Reaction R-G-7 is reversible; therefore, when it reaches to a balance, adding steam will not
4
5 further affect the reaction.
6
7
8 Figure 10. The influence of the S/A ratio on the composition of the produced gases
9
10
11
12 5 Conclusions
13
14
15 A two-dimensional CFD model of a downdraft fixed bed gasifier was developed. The
16
17 simulation results were compared with experimental data and showed that the model can
18
19 reasonably predict the temperature profile and the composition of the gases during the HTAG
20
21 process.
22
23
24
The basic case was analyzed, and parameter studies were carried out on the temperature
25
26 and gas composition profiles to assess the effects of the preheating temperature and the S/A
27
28 ratio.
29
30 In the gasifier, the temperature distribution is not uniform in a given horizontal section.
31
32
The peak temperatures in different sections appear at 0.23 m< x < 0.29 m, which could be
33
34
35 explained by the occurrence of char combustion in a certain layer near the gas-bed interface.
36
37 The increase in the feed gas preheating temperature will lead a higher bed temperature
38
39 and a relatively decreased required biomass residence time, which makes the HTAG
40
41 technology more insensitive to variability. Additionally, the positive effects of higher
42
43
44
preheating temperatures are an increase in the combustible gases (H2, CO) and a decrease in
45
46 the tar produced.
47
48 The increase of the S/A ratio is also beneficial in reducing the tar content. Hence, at
49
50 higher S/A ratios, the mole fractions of H2 and CO2 were high and that of CO fraction was
51
52
low. The CH4 content showed a peak value that was determined by the balance of the
53
54
55 pyrolysis reactions (which produce CH4) and the methane - water shift reaction (which
56
57 consumes CH4).
58
59
60 17
ACS Paragon Plus Environment
Energy & Fuels Page 18 of 36

1
2
3 Conventional gasification technologies rely on the combustion inside the process itself to
4
5 sustain the reaction. However, the HTAG technology relies on an external heat source and
6
7 uses super-heated air combined with steam, which results in a limited need for combustion
8
9
10 within the gasifier and produces syngas with a high H2 fraction and low tar content, which is
11
12 environmentally preferable.
13
14
15
16 Acknowledgments
17
18
19 The financial support of this work by KIC InnoEnergy is gratefully acknowledged.
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 18
ACS Paragon Plus Environment
Page 19 of 36 Energy & Fuels

1
2
3 Nomenclature
4
5
6 A! pre-exponential factor of reaction i
7
8
9 #$ specific surface area %& '
10
11
12 #( surface area of solid particles % '
13
14
15  mass concentration % / '
16
17
18 )( particle diameter %'
19
20
21 * diffusion coefficient of the vapor in the bulk % /+'
22
23
24 ,- activation energy of reaction .
25
26
27 / gravitational acceleration %/+'
28
29
30 gas phase
31
32
33 h specific enthalpy %1/ '
34
35
36 H3$45 heat of evaporation of water % /'
37
38
39
40
 heat transfer coefficient %6/  '
41
42
43 . moisture
44
45
46 7 mass transfer rate % / +'
47
48
49 8 molar weight % /'
50
51
52 87 mass flow rate % /+'
53
54
55 9 pressure %:;'
56
57
58 9<.  =;< primary tar
59
60 19
ACS Paragon Plus Environment
Energy & Fuels Page 20 of 36

1
2
3
4
5
6 > heat flux %6/ '
7
8
9 ? interphase heat transfer %6/ '
10
11
12 < reaction rate %/ +'
13
14
15 <@- kinetic rate of chemical reaction . %/+'
16
17
18 <A- turbulent mixing rate of reactants in reaction . %/ +'
19
20
21 B universal gas constant %1/'
22
23
24
+ solid phase
25
26
27
28
+CD  =;< secondary tar
29
30
31 E source term
32
33
34 = time %+'
35
36
37 F temperature %'
38
39
40 G volatile from wood pellets species
41
42
43 G/ velocity %/+'
44
45
46 G- stoichiometric coefficient of reactant .
47
48
49 H- mass fraction of species .
50
51
52 I volume fraction
53
54
55 J density % / '
56
57
58
59
60 20
ACS Paragon Plus Environment
Page 21 of 36 Energy & Fuels

1
2
3 K stress tensor %:;'
4
5
6 L dynamic viscosity %:; +'
7
8
9 M angle of internal friction
10
11
12 N thermal conductivity %6/'
13
14
15
 turbulent dissipation rate % /+  '
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 21
ACS Paragon Plus Environment
Energy & Fuels Page 22 of 36

1
2
3 References
4
5
1. Saidur, R., et al., A review on biomass as a fuel for boilers. Renewable and Sustainable Energy
6
7 Reviews, 2011. 15(5): p. 2262-2289.
8 2. Janet I, S., Renewables 2012, in Global Status Report L. Mastny, Editor 2012: Paris, France. p.
9 33.
10 3. Vaezi, M., et al., Modeling Biomass Gasification: A New Approach to Utilize Renewable
11 Sources of Energy. ASME Conference Proceedings, 2008. 2008(48692): p. 927-935.
12 4. Jarungthammachote, S. and A. Dutta, Thermodynamic equilibrium model and second law
13 analysis of a downdraft waste gasifier. Energy, 2007. 32(9): p. 1660-1669.
14 5. Zainal, Z.A., et al., Prediction of performance of a downdraft gasifier using equilibrium
15 modeling for different biomass materials. Energy Conversion and Management, 2001. 42(12):
16 p. 1499-1515.
17
6. Sharma, A.K., Equilibrium modeling of global reduction reactions for a downdraft (biomass)
18
gasifier. Energy Conversion and Management, 2008. 49(4): p. 832-842.
19
20 7. Melgar, A., et al., Thermochemical equilibrium modelling of a gasifying process. Energy
21 Conversion and Management, 2007. 48(1): p. 59-67.
22 8. Huang, H.J. and S. Ramaswamy, Modeling biomass gasification using thermodynamic
23 equilibrium approach. Applied Biochemistry and Biotechnology, 2009. 154(1-3): p. 193-204.
24 9. Wang, Y. and C.M. Kinoshita, Kinetic model of biomass gasification. Solar Energy, 1993. 51(1):
25 p. 19-25.
26 10. Giltrap, D.L., R. McKibbin, and G.R.G. Barnes, A steady state model of gas-char reactions in a
27 downdraft biomass gasifier. Solar Energy, 2003. 74(1): p. 85-91.
28 11. Blasi, C.D., Dynamic behaviour of stratified downdraft gasifiers. Chemical Engineering Science,
29 2000. 55(15): p. 2931-2944.
30
12. Shetha, P.N. and B. Babub, Modeling and Simulation of Downdraft Biomass Gasifier.
31
13. Marklund, M., R. Tegman, and R. Gebart, CFD modelling of black liquor gasification:
32
33 Identification of important model parameters. Fuel, 2007. 86(12–13): p. 1918-1926.
34 14. Fletcher, D.F., et al., Computational fluid dynamics modelling of an entrained flow biomass
35 gasifier. Applied Mathematical Modelling, 1998. 22(10): p. 747-757.
36 15. Xue, Q., T.J. Heindel, and R.O. Fox, A CFD model for biomass fast pyrolysis in fluidized-bed
37 reactors. Chemical Engineering Science, 2011. 66(11): p. 2440-2452.
38 16. Papadikis, K., S. Gu, and A.V. Bridgwater, CFD modelling of the fast pyrolysis of biomass in
39 fluidised bed reactors. Part B: Heat, momentum and mass transport in bubbling fluidised beds.
40 Chemical Engineering Science, 2009. 64(5): p. 1036-1045.
41 17. Zhang, Q., et al., Eulerian Model for Municipal Solid Waste Gasification in a Fixed-Bed Plasma
42 Gasification Melting Reactor. Energy & Fuels, 2011. 25(9): p. 4129-4137.
43 18. Gerun, L., et al., Numerical investigation of the partial oxidation in a two-stage downdraft
44
gasifier. Fuel, 2008. 87(7): p. 1383-1393.
45
46 19. Donaj, P., et al., Effect of Pressure Drop Due to Grate–Bed Resistance on the Performance of a
47 Downdraft Gasifier. Energy & Fuels, 2011. 25(11): p. 5366-5377.
48 20. Jones, W.P. and B.E. Launder, The prediction of laminarization with a two-equation model of
49 turbulence. International Journal of Heat and Mass Transfer, 1972. 15(2): p. 301-314.
50 21. Cowin, S.C., A theory for the flow of granular materials. Powder Technology, 1974. 9(2-3): p.
51 61-69.
52 22. Schaeffer, D.G., Instability in the evolution equations describing incompressible granular flow.
53 Journal of Differential Equations, 1987. 66(1): p. 19-50.
54 23. Syamlal, M.R., W.; O’Brien, T. J.MFIX Documentation: Theory Guide; National Technical
55 Information Service: Springfield, 1993: VA.
56 24. S. Sarkar, L.B., Applications of a Reynold-Stress Turbulence Model to the Compressible Shear
57
Layer, 1990: NASA CR 18002. p. ICASE Report 90-18.
58
59
60 22
ACS Paragon Plus Environment
Page 23 of 36 Energy & Fuels

1
2
3 25. Kuipers, J.A.M., et al., A numerical model of gas-fluidized beds. Chemical Engineering Science,
4 1992. 47(8): p. 1913-1924.
5 26. Gunn, D.J., Transfer of heat or mass to particles in fixed and fluidised beds. International
6 Journal of Heat and Mass Transfer, 1978. 21(4): p. 467-476.
7 27. Di Blasi, C., et al., Product Distribution from Pyrolysis of Wood and Agricultural Residues.
8 Industrial & Engineering Chemistry Research, 1999. 38(6): p. 2216-2224.
9 28. Boroson, M.L., et al., Product yields and kinetics from the vapor phase cracking of wood
10
pyrolysis tars. AIChE Journal, 1989. 35(1): p. 120-128.
11
12
29. Di Blasi, C. and C. Branca, Kinetics of Primary Product Formation from Wood Pyrolysis.
13 Industrial & Engineering Chemistry Research, 2001. 40(23): p. 5547-5556.
14 30. Park, W.C., A. Atreya, and H.R. Baum, Experimental and theoretical investigation of heat and
15 mass transfer processes during wood pyrolysis. Combustion and Flame, 2010. 157(3): p. 481-
16 494.
17 31. Hobbs, M.L., P.T. Radulovic, and L.D. Smoot, Modeling fixed‐bed coal gasifiers. AIChE
18 Journal, 1992. 38(5): p. 681-702.
19 32. Donaj, P., Conversion of Biomass and Waste Using Highly Preheated Agents for Materials and
20 Energy Recovery, in Energy and Furnace Technology2011, Royal Institute of Technology:
21 Stockholm, Sweden.
22
33. Hobbs, M.L., P.T. Radulovic, and L.D. Smoot, Combustion and gasification of coals in fixed-
23
beds. Progress in Energy and Combustion Science, 1993. 19(6): p. 505-586.
24
25 34. Arthur, J.R., Reactions between carbon and oxygen. Transactions of the Faraday Society, 1951.
26 47: p. 164-178.
27 35. Yoon, H., J. Wei, and M. Denn, A model for Moving-Bed coal gasification reactors. AIChE
28 Journal, 1978. 24(5): p. 885-903.
29 36. Siminski V.L., W.F.J., Edelman R.B., Economos C., Fortune O.F. , Research on methods of
30 improving the combustion characters of liquid hydrocarbon fuels,AFAPL TR 72-74, vols. I and
31 II, , 1972, Air Force Aeroprolulsion Laboratory.
32 37. Varma, A.K., A.U. Chatwani, and F.V. Bracco, Studies of premixed laminar hydrogen  air
33 flames using elementary and global kinetics models. Combustion and Flame, 1986. 64(2): p.
34 233-236.
35
38. Howard, J.B., G.C. Williams, and D.H. Fine, Kinetics of carbon monoxide oxidation in
36
37
postflame gases. Symposium (International) on Combustion, 1973. 14(1): p. 975-986.
38 39. Macak, J. and J. Malecha, Mathematical Model for the Gasification of Coal under Pressure.
39 Industrial & Engineering Chemistry Process Design and Development, 1978. 17(1): p. 92-98.
40 40. Dryer, F.L. and I. Glassman, High-temperature oxidation of CO and CH4. Symposium
41 (International) on Combustion, 1973. 14(1): p. 987-1003.
42 41. Jones, W.P. and R.P. Lindstedt, Global reaction schemes for hydrocarbon combustion.
43 Combustion and Flame, 1988. 73(3): p. 233-249.
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 23
ACS Paragon Plus Environment
Energy & Fuels Page 24 of 36

1
2
3 Figure captions
4
5
6 Figure 1. A schematic of the downdraft fixed bed gasifier in the KTH laboratory
7
8 Figure 2. A schematic of the CFD model
9 Figure 3. The experimental and simulated temperature distributions for the centerline along
10 the vertical axis inside the gasifier
11
12 Figure 4. A comparison of the simulated produced gas compositions with the experimental
13 values
14 Figure 5. The gas temperature distribution for different horizontal sections in the charcoal
15 bed
16 Figure 6. Profiles of the average gas composition along the vertical axis
17
Figure 7. The gas phase parameters for different preheating temperatures: (a) temperature
18
19 (K), (b) the mass concentration of the primary tar (g/Nm3), (c) the mass concentration of the
20 secondary tar (g/Nm3).
21 Figure 8. The influence of the feed gas preheating temperature on the composition of the
22
produced gases
23
24 Figure 9. The gas phase parameters for different S/A ratios: (a) temperature (K), (b) the mass
25 concentration of the primary tar (g/Nm3), (c) the mass concentration of secondary the tar
26 (g/Nm3).
27
28 Figure 10. The influence of the S/A ratio on the composition of the produced gases
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 24
ACS Paragon Plus Environment
Page 25 of 36 Energy & Fuels

1
2
3 Table captions
4
5
6 Table 1. The characterization of the biomass feedstock
7
8 Table 2. The kinetic reaction rates of the char reactions
9 Table 3. The kinetic reaction rates of the homogeneous reactions
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 25
ACS Paragon Plus Environment
Energy & Fuels Page 26 of 36

1
2
3 Tables
4
5 Table 1. The characterization of the biomass feedstock
6 Proximate analysis Ultimate analysis (dry basis)
7
8 Moisture 8% C 50%
9
10 Volatiles 84% (dry basis) H 6.0-6.2%
11
12 Fixed carbon 15.5% (dry basis) O 43-44%
13
14 Ash 0.5% (dry basis) N <0.2%
15
16 S 0.01-0.02%
17
18
19
20 Table 2. The kinetic reaction rates of the char reactions
21
Reactions Kinetic reaction rate (m/s) Ref.
22
23 rk 8 = 2.30Ts exp (− 11100 / Ts )
R-C-1: 2C + O2 →
rk 8
2CO [33]
24
25
R-C-2: C + O2 →
rk 9
CO2 rk 8 [34]
26
27 O 2512exp %6420UF '
T
28 rk 9
29
30 R-C-3: C + H 2O r
k 10
→ CO + H 2 rk 7 = 5.714Ts exp(− 15600 / Ts ) [35]
31
32 R-C-4: C + CO2 r→
k 11
2CO rk 8 = 5.89 × 10 2 Ts exp (− 26800 / Ts ) [33]
33
34
35
36 Table 3. The kinetic reaction rates of the homogeneous reactions
37
Reactions Kinetic reaction rate (kmol/m3·s) Ref
38
39 Pr imary tar + O2 → H 2O + CO rk 1 = 59 .8T g P 0.3α g exp (− 12200 / T g )C pri − tar
rk 1 0.5
R-G-1: C O2 [36]
40
41
42
R-G-2: Secondary tar + O 2 →
rk 2
( )
H 2 O + CO rk 2 = 59 .8T g P 0.3α g exp − 12200 / T g C sec − tar 0.5 C O 2 [36]
43
44
R-G-3: H 2 + 1/ 2O2 → rk 3 = 3.53 × 10 8.4 α g exp (− 3670 / T g )C H 2 C O2
rk 3 1.1 1.1
45 H 2O [37]
46
CO + 1 / 2O2 → rk 4 = 1 .3 × 10 11 α g exp (− 15105 / T g )C CO C H 2 O C O 2
0 .5 0 .5
47 R-G-4: rk 4
CO2 [38]
48
49 R-G-5: CO + H 2 O ←
rk 5
→ H 2 + CO2  exp(− 7914 / Tg )CCO2 CH 2  [35,
rk 5 = 2.78α g exp(− 1511/ Tg ) CCOCH 2O − 
50  0.0265  39]
51
52
53 CH 4 + 1.5O2 → CO + 2 H 2O rk 6 = 1 .0 × 10 11 .7 α g exp (− 24357 / T g )C CH 4
0 .7 0 .8
R-G-6: rk 6
C O2 [40]
54
55 R-G-7 CH 4 + H 2O → CO + 4 H 2 rk 7 = 3 .0 × 10 8 α g T g exp (− 15083 / T g )C CH 4 C H 2 O
rk 7
[41]
56
57
58
59
60 26
ACS Paragon Plus Environment
Page 27 of 36 Energy & Fuels

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30 Figure 1. A schematic of the downdraft fixed bed gasifier in the KTH laboratory
84x57mm (150 x 150 DPI)
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Energy & Fuels Page 28 of 36

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28 Figure 2. A schematic of the CFD model
29 83x62mm (127 x 127 DPI)
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 29 of 36 Energy & Fuels

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Figure 3. The experimental and simulated temperature distributions for the centerline along the vertical axis
35 inside the gasifier
139x111mm (150 x 150 DPI)
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Energy & Fuels Page 30 of 36

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25 Figure 4. A comparison of the simulated produced gas compositions with the experimental values
26 84x44mm (150 x 150 DPI)
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 31 of 36 Energy & Fuels

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
Figure 5. The gas temperature distribution for different horizontal sections in the charcoal bed
27 85x48mm (150 x 150 DPI)
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Energy & Fuels Page 32 of 36

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30 Figure 6. Profiles of the average gas composition along the vertical axis
31 84x58mm (150 x 150 DPI)
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 33 of 36 Energy & Fuels

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
Figure 7. The gas phase parameters for different preheating temperatures: (a) temperature (K), (b) the
47
mass concentration of the primary tar (g/Nm3), (c) the mass concentration of the secondary tar (g/Nm3).
48 80x154mm (150 x 150 DPI)
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Energy & Fuels Page 34 of 36

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29 Figure 8. The influence of the feed gas preheating temperature on the composition of the produced gases
30 84x56mm (150 x 150 DPI)
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 35 of 36 Energy & Fuels

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
Figure 9. The gas phase parameters for different S/A ratios: (a) temperature (K), (b) the mass
47
concentration of the primary tar (g/Nm3), (c) the mass concentration of secondary the tar (g/Nm3).
48 79x151mm (150 x 150 DPI)
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Energy & Fuels Page 36 of 36

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29 Figure 10. The influence of the S/A ratio on the composition of the produced gases
30 85x55mm (150 x 150 DPI)
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment

Potrebbero piacerti anche