Full Length Article: Sciencedirect

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Fuel 280 (2020) 118582

Contents lists available at ScienceDirect

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

Full Length Article

Kinetic modeling of biomass gasification and tar formation in a fluidized bed T


gasifier using equivalent reactor network (ERN)

Bijoy Das, Atmadeep Bhattacharya1, Amitava Datta
Department of Power Engineering, Jadavpur University, Salt Lake Campus, Kolkata 700098, India

A R T I C LE I N FO A B S T R A C T

Keywords: A chemical kinetic based simulation model of a bubbling fluidized bed biomass gasifier has been developed using
Biomass gasification an Equivalent Reactor Network (ERN). The pyrolysis zone is modeled using a thermodynamic equilibrium model
Pyrolysis and a kinetic-controlled perfectly stirred reactor (PSR) placed in succession. The gasification zone and the
Kinetic modeling freeboard are simulated as perfectly stirred reactor (PSR) and plug flow reactor (PFR), respectively. A detailed
Tar
chemical reaction mechanism is used to describe the chemistry of reaction during devolatilization and gasifi-
Equivalent reactor network
cation stages. The formation and evolution of gaseous and tar species in different zones of the gasifier are
estimated and the major reaction paths have been identified. The influence of equivalence ratio (ER) on the
gasifier performance and tar species formation has been studied. Out of the three ERs considered (0.27, 0.24 and
0.19), the best performance of the gasifier is exhibited at ER = 0.24 in terms of cold gas efficiency.

1. Introduction The fuel gas (named producer gas or syngas), produced through
gasification of biomass in air, contains hydrogen (H2), carbon monoxide
Biomass is a widely available, carbon–neutral energy resource that (CO) and methane (CH4) as the combustible components; in addition to
currently accounts for more than 10% of the world's total energy con- the other non-combustible components like carbon dioxide (CO2) and
sumption [1]. One convenient way of extracting energy from the bio- nitrogen (N2). The syngas composition varies significantly with the
mass is through its gasification and then using the resultant fuel gas biomass feedstock, gasifier type and gasifier operating parameters [6].
either in internal combustion engine or in gas turbine engine to produce It is an established fact that the performance of both the IC engine and
electricity. However, as the energy content and density of the biomass the IGCC plant depends strongly on the quality of the fuel gas produced
feedstock are low, the cost of the biomass transportation should be from biomass gasification [7,8]. The producer gas also contains tar,
considered in the economic assessment of biomass conversion tech- which is a complex mixture of hydrocarbons heavier than benzene [9].
nologies [2]. Therefore, the power generation from gasification of Tar is produced in the gasifier primarily during the pyrolysis of bio-
biomass in small scale decentralized units is often a more economical mass; and then it is subjected to secondary reactions. Different com-
option than that in a large scale centralized plant [3,4]. Furthermore, pounds present in tar have been listed by Milne et al. [10] following the
the emission indices of carbon monoxide (CO) and soot from a gasifier- work of Evans and Milne [11]. Tar is an unwanted constituent of the
engine combined system operating on woody biomass remains negli- producer gas, which adversely affects the quality of the fuel for its use
gibly small [5]. Besides the gasifier and reciprocating internal com- in engines and turbines. The tar constituents may condense to block the
bustion engine couplings, there are a few works available in the lit- fuel lines and particle filters in engines. They also result in the increased
erature on the gasifier and gas turbine combinations as well. Such emission of soot and UHC (unburned hydrocarbon) in the exhaust.
combinations may also be used in conjunction with steam turbines in Therefore, reduction of tar in the producer gas remains an important
the integrated gasification combined cycle (IGCC) plants. A detailed objective in the design and development of the biomass gasifier unit. A
review of the usefulness of the IGCC technology involving biomass can recent review by Islam [12] has listed numerous catalytic materials that
be found in Parraga et al. [6]. It has been identified in this review that are being researched upon for the tar removal process in biomass ga-
the biomass gasification process is quite complex and therefore its sification.
modeling is a challenging task. Another method of reducing the quantity of tar in the producer gas


Corresponding author.
E-mail address: amdatta_ju@yahoo.com (A. Datta).
1
Presently at Department of Mechanical Engineering, Aalto University, Finland.

https://doi.org/10.1016/j.fuel.2020.118582
Received 6 April 2020; Received in revised form 31 May 2020; Accepted 29 June 2020
Available online 11 July 2020
0016-2361/ © 2020 Elsevier Ltd. All rights reserved.
B. Das, et al. Fuel 280 (2020) 118582

is by in-situ control of tar formation through appropriate choices of the thermodynamics and Sandia PSR [34] for kinetic simulations – for the
gasifier type and operating parameters [13,14]. Several gasification precise prediction of the biomass pyrolysis products. It was demon-
parameters, like the type of biomass feedstock, gasification agent and strated in this work that the combination of thermodynamic equili-
equivalence ratio in the gasifier, gasification temperature and pressure, brium model in HSC chemistry and PSR can successfully predict the
are found to influence the formation of tar during biomass gasification. pyrolysis products. Dupont et al. [35] incorporated only the kinetic
The fluidized bed gasification (FBG) process leads to efficient tar processes to model the biomass-steam gasification considering the
cracking due to high gasifier temperature causing a reduction in the biomass transformation process to occur in two successive steps – (i)
quantity of tar in the gas [15]. At the same time, bubbling fluidized bed pyrolysis of biomass and (ii) steam gasification of solid residue. They
gasifiers are among the simplest (in terms of scalability and controll- observed that the gasification process is controlled by the chemical
ability) and the most cost-effective technology of gasification with kinetics of reactions. They further compared the time scales of different
flexibility towards a wide variety of biomass feedstock [16,17]. processes involved in gasification and showed that the pyrolysis time is
Experimental determination of tar in producer gas is a resource much less than the gasification reaction time and the relative time
intensive exercise that necessitates successful sampling and highly re- scales depend on the gasification temperature. Hafner et al. [36] used a
liable measurement [18,19]. On the other hand, a suitable model of tar detailed reaction mechanism for high temperature gasification of
can predict the formation and transformation of tar compounds in ethylene glycol produced by fast pyrolysis of biomass in an entrained
different zones of the gasifier and also evaluate the composition in the bed gasifier. They validated their simulation results using experimental
final producer gas. The model can also be used to optimize the oper- data from the literature for concentration profile, laminar flame speeds,
ating parameters for controlling the tar quantity in the fuel. However, it ignition delay times as well as gasification result in complex turbulent
is necessary to couple the tar formation model with the overall model of reactive flow. A multi-physics model by Qiao et al. [37], considering
the gasifier, which will predict the composition of the producer gas detailed gas phase chemistry and particle phase reactions, simulated the
formed in gasification. gasification processes in a well-stirred reactor. They studied the effects
Tar is a complex mixture of condensable hydrocarbon compounds, of particle size, pressure, porous structure, O2 concentration, radiative
classified as primary, secondary and tertiary, according to the appear- heat transfer and H2 addition on gasification performance. Another
ance of the molecules and based on its evolution according to tem- detailed kinetic model for bio-oil gasification was developed by Khan-
peratures [20]. Many researchers have represented tar by an equivalent shan and West [38] using separate plug flow (PFR) and continuous
single hydrocarbon (like CHpOq) to simplify the modeling of its for- stirred tank (CSTR) reactors. They used the sensitivity analysis to
mation [21,22]. Such assumption enables to include tar in the equili- identify the most valuable information for improving the mechanism
brium modeling of the gasification process [23]. However, the equili- predictions. A reactor network modeling (RNM), comprising a CSTR
brium models have the limitations in over-predicting the gas and PFR, was adopted by Stark et al. [32] for air-blown gasification in
production, tar content and the concentration of the combustible gas- fluidized bed gasifier. The biomass mechanism developed by the CRECK
eous components in the producer gas [24,25]. An improved modeling of group was adopted in the model to simulate the chemical reactions. The
tar formation and destruction considers tar to be composed of a few concentrations of major gaseous species and tar compounds, produced
lumps of hydrocarbons involving a set of kinetic equations [26]. Palma during biomass gasification, were predicted and validated. It was found
[27] proposed a brief chemical kinetic model on tar formation from that the RNM model can predict the major syngas species and one-ring
lignin in a fluidized bed gasifier. This model successfully gave quali- aromatic tars as well as heterocyclic aromatic tars with good accuracy.
tative chemical insights into the tar formation process. In another work, It has also been shown in the work that the char conversion following
the same author [13] has reviewed detailed kinetic mechanisms for tar pyrolysis is mostly by elutriation from the bed and carbon conversion
formation and evolution in biomass gasification comprising large through char gasification and oxidation plays insignificant role.
number of elementary reaction steps. Moreover, the need for better It is clear from the above-mentioned literature survey that accurate
understanding of detailed chemical kinetics of the tar formation me- prediction of tar species with detailed chemical kinetics in the biomass
chanism has been highlighted in this review work. gasification process needs more attention. The transformation of bio-
The use of computational fluid dynamics (CFD) in the simulation of mass in a gasifier takes place through two successive steps – pyrolysis
a fluidized bed gasifier can provide a deeper insight into the transport and gasification, and chemical kinetics play an important role in both of
processes, predicting the distributions of the velocity, pressure, tem- them. In a fluidized bed gasifier, the vigorous mixing due to movements
perature, species concentrations and other hydrodynamic parameters of bubbles and bed solids creates an environment of uniform tem-
inside the gasifier. Ismail et al. [28] adopted the Eulerian–Eulerian 2D perature and concentrations. Thus, the continuously stirred reactor
CFD model to simulate the fluidized bed biomass gasification and in- model applies well within the bed. To the best of our knowledge, no
vestigated the concentrations of individual species and cold gas effi- work has been reported in the literature encompassing, successively,
ciency at different equivalence ratio and biomass moisture content. As the detailed kinetic model both for the pyrolysis and gasification pro-
an alternative to this approach, Ku et al. [29] simulated the gas phase cesses. Therefore, the present work deals with the chemical kinetic
by Eulerian method and the particle phase by discrete element method analysis of the fluidized bed gasification process in an equivalent re-
(DEM), thereby tracing the individual particles to accurately predict the actor network (ERN) that can physically describe the transport pro-
microscopic information like the particle trajectories, particle–particle cesses in the gasifier. This approach will require relatively small com-
interactions, particle entrainments, and particle reactions. During the putational demands of mathematical model with large chemical
CFD simulation in the reactor, the associated chemical kinetic me- reaction mechanism. The chemical kinetic mechanism has been opti-
chanism consumes the major part of the computational resources [30]. mized in the parametric range under consideration using sensitivity
Therefore, it is a common practice to use global reactions in CFD si- analysis. The mechanism has been validated using the experimental
mulations in order to minimize the computational cost [31]. However, data from the literature. The tar formation and destruction pathways
the detailed chemical reaction path from the input biomass to the have been analyzed and the concentrations of important tar compo-
product gas is missing in the global reaction pathway. Simplistic global nents in different zones of the gasifier and finally in the producer gas
mechanisms fail to evaluate the tar composition correctly and cannot have been reported at various operating conditions.
predict the component-wise distribution of tar [32].
To overcome the shortcomings of the equilibrium modeling and 2. Simulation methodology
global mechanism based CFD simulation, Lee et al. [33] combined
thermodynamic and kinetic based models for the pyrolysis of biomass. The present analysis deals with the fluidized bed gasification of dry
They applied two computational codes – HSC Chemistry for and ash free woody biomass in a stream of air. Fig. 1(a) shows the

2
B. Das, et al. Fuel 280 (2020) 118582

Fig.1. Schematic diagram of (a) the bubbling fluidized bed biomass gasifier showing different zones in it; (b) Equivalent reactor network for the gasifier.

schematic of a bubbling fluidized bed biomass gasifier with the biomass 1. Pyrolysis zone has been modeled as a combination of thermo-
feedstock fed from the top and the gasification air introduced from the dynamic equilibrium model and kinetic model [33]. The position of
bottom. Such a scheme has been adopted in the present work from the the pyrolysis zone is adjacent to the fuel entry region (Fig. 1 (a)).
experimental study of Kim et al. [39]. It may be mentioned in this re- The oxygen present in air fed from bottom of the gasifier does not
gard that similar fuel and oxidizer feeding strategies in bubbling flui- reach the fuel entry point due to the sub-stoichiometric environment
dized bed biomass gasifiers are quite commonly reported in the lit- inside the reactor.
erature [40–42]. The air, which is sub-stoichiometric in quantity, 2. The volume of the pyrolysis zone has been assumed as 2% of the
maintains fluidization of the bed in the bubbly flow regime, where total gasifier volume from a scale analysis considering the fact that
vigorous bubble movement inside the bed ensures good mixing. The the pyrolysis takes place within a very short time upon entry of the
oxygen in air gets fully consumed in the chemical reactions occurring in biomass into the gasifier [43,35,27]. Devolatilization occurs in-
the bed. The biomass, entering the gasifier, gets pyrolyzed due to the stantaneously during pyrolysis so that the thermodynamic equili-
high temperature and in absence of oxygen, forming the gaseous brium exists among the product gases from the devolatilization
components, tar and char. Further gasification reactions continue in the process. The devolatilization products further react among them-
fluidized bed and finally in the freeboard above the bed, as shown in selves following a kinetic scheme within the pyrolysis zone. During
Fig. 1(a). In order to compute the composition of the output gas, and such reactions, the homogeneous gaseous interaction is much faster
concentration of tar components an equivalent reactor network (ERN) than the heterogeneous surface reaction [32]. The small volume of
has been formed (Fig. 1 (b)) considering three zones – i.e. pyrolysis the pyrolysis zone and the turbulence in the flow results in a well
zone, gasification zone and freeboard zone. The following assumptions stirred situation during the kinetic reaction which has been modeled
have been made in developing the model: using a perfectly stirred reactor (PSR).
3. The gasification reactions occur in the fluidized bed which is

3
B. Das, et al. Fuel 280 (2020) 118582

approximated as a perfectly stirred reactor (PSR) considering the

Fixed bed gasification involving


very high rate of mixing due to the hydrodynamics of the bed.
4. The freeboard is modeled as a plug flow reactor (PFR) due to little

Detailed chemical kinetics

Detailed chemical kinetics


axial mixing or negligible mass diffusion in the flow direction

Sommariva et al. [49]


5. As the model predictions are validated against the experimental
results of Kim et al. [39], the temperatures at different zones of the
gasifier are assumed to be the same as those in the experiments. The

solid fuels
temperature values of different zones were set by averaging the
experimental thermocouple readings of Kim et al. [39] in the re-
spective zones.
6. The walls of the gasifier are considered as non-catalytic. All gaseous
species are assumed to be ideal with variable specific heat.

Devolatilization and gas phase reactions in


The thermodynamic equilibrium calculation for the simulation of
initial part of the pyrolysis process has been done with the HSC
Chemistry 6.0 tool [44]. The biomass entering the pyrolysis zone has

Detailed chemical kinetics


elements C (51.02%), H (7.16%), O (41.73%), N (0.09%) and S
(0.004%), following the ultimate analysis of the feedstock [39]. During
the evaluation of equilibrium species composition, the Gibbs free en-

biomass pyrolysis
Ranzi et al. [48]
ergy has been minimized for the mixture. The species considered during
the equilibrium calculation are H, H2, C, CO, CO2, CH4, HCN, NH3,
C2H2, C2H6, C2H4, C3H8, N, N2, NO, N2O, O, O2, H2O, H2O2, HO2, NO2,

——
NO3. However, as the hydrocarbons heavier than CH4 are not ther-
modynamically stable species, their existence in the pyrolyzed gas
cannot be predicted from the equilibrium calculation [45]. Therefore,

Thermodynamic equilibrium and well-stirred reactor using


the thermodynamic equilibrium calculation returns only H2, C, CO,
CO2, CH4 and H2O in significant concentrations. The components from
the HSC calculation are further reacted in a perfectly stirred reactor
(PSR1 in Fig. 1(b)), which is also a part of the pyrolysis process, ac-
counting the kinetic rates of chemical reactions.
Table 1 summarizes the simulation methodologies adopted in the
present work for the modeling of pyrolysis and gasification zones in a
biomass gasifier and comparing them with methodologies adopted by
skeletal GRI-Mech mechanism

other researchers. It may be seen from the table that Lee et al. [33] used
similar strategy to predict the gaseous products from biomass pyrolysis.
However, the present treatment of biomass pyrolysis differs from the
Biomass pyrolysis

work of Lee et al. in the following respects. The carbon (C) in the
Lee et al. [33]

equilibrium products from HSC calculation represents both volatile and


solid phase carbon with the relative amount dependent on temperature.
Description of pyrolysis and gasifier models adopted in the present work from literature.

For the pyrolysis temperature considered in the present work, the solid
——

carbon in the equilibrium product is found to be 65–68% of the total


carbon. The solid phase carbon remains as charcoal. Lee et al. [33]
Thermodynamic equilibrium and well-stirred reactor using

Well-stirred and plug flow reactor using detailed chemical

eliminated the solid phase carbon from the equilibrium product mixture
and considered only the pyrolyzed gas components and the volatile
carbon to react in the PSR through kinetic reactions. However, the
charcoal formed after devolatilization of woody biomass is a micro-
Bubbling fluidized bed biomass gasification

porous solid with a very high surface area per unit volume. Therefore,
without neglecting the solid phase carbon from further reaction, we
have assumed it to react homogeneously over the whole volume of the
charcoal, with access of gaseous components through the large number
of micropores and mesopores. Thus, the entire carbon obtained from
detailed chemical kinetics

the equilibrium calculation continues to react in the PSR1 of the pyr-


olysis zone. Moreover, as mentioned in Table 1, Lee et al. [33] con-
sidered the GRI-Mech chemical mechanism and the associated kinetic
Present Work

parameters to solve the chemistry in the perfectly stirred reactor during


pyrolysis. GRI-Mech was developed for the oxidation of natural gas and
kinetics

does not employ well for the reactions involving higher hydrocarbons.
So, it will not be able to correctly account the wide variety of com-
pounds in the biomass tar formed on pyrolysis. We, therefore, adopted a
Investigated phenomenon

chemical mechanism more suited for biomass devolatilization and


pyrolysis.
Pyrolysis model

The pyrolysis is followed up by the gasification process, modeled


Gasifier model

using a perfectly stirred reactor (PSR2), occurring in the fluidized bed


(Fig. 1(b)), with the admission of fluidization air of sub-stoichiometric
Table 1

quantity. The PSR is characterized by the volume of the reactor, mass


flow rate of fluid through the reactor, inlet conditions (temperature and

4
B. Das, et al. Fuel 280 (2020) 118582

composition) and temperature of the reactor. In the PSR, the mixing at atmospheric pressure. The temperatures during gasification have
rates are considered to be very fast and the chemical conversions from been considered as 769 °C, 765 °C and 739 °C for the 0.27, 0.24 and
the reactants to the products are controlled by the rates of kinetics of 0.19, respectively. All these temperatures correspond to the average
reactions. The Modeling of both PSR1 and PSR2 has been done by experimental temperature data given in Kim et al. [39].
solving a system of nonlinear algebraic equations using Chemkin-Pro
package to evaluate the species concentrations within the reactor
(which are equal to those at the outlet) considering homogeneous re- 3. Results and discussion
actions in the gas phase. The equations are solved by the damped
modified Newtown algorithm. The mathematical details of the perfectly 3.1. Analysis of syngas production
stirred reactors used in the present work may be found in [34,46].
It may be seen from Fig. 1 (b) that the products of PSR2 enter the The developed model is used to simulate the gasification of woody
freeboard zone above the bed. Low solid concentration, perfect mixing, biomass in a bubbling fluidized bed gasifier and predict the composi-
uniform properties in the radial direction and change of concentration tion of gaseous constituents and tar in the producer gas. A comparison
along the axis of flow make the freeboard zone to be suitably modeled of the mole fractions of the major gaseous constituents, like H2, CO, CO2
as PFR. A set of ordinary differential equation is solved for conservation and CH4, in the producer gas predicted from the model against those
of mass, energy and momentum for a differential slice in the flow di- reported from the experiments [39], at different ER, is shown in
rection to solve the PFR model. The relevant equations pertaining to the Fig. 2(a-d) at identical pressures, temperatures, equivalence ratios and
PFR in the freeboard zone has been solved using Senkin program [47] biomass feed rates (Table 2). However, the experimental data on tar
in the Chemkin-Pro package. The mathematical details of PFR used in quantity at the gasifier outlet has not been provided by Kim et al. [39].
this simulation are available in [46]. Therefore, the tar concentration in the producer gas predicted using the
The kinetic model of multistep chemical reactions focusses on the present model has been compared with the experimental data from
pyrolysis reactions forming gaseous species, tars and solid carbon, Campoy et al. [50] for the same biomass fuel and equivalence ratios
secondary gas phase reactions of the gaseous and tar species, and re- (ER = 0.27 and 0.19) at atmospheric pressure. In the present work, the
action of the char with the gaseous species. Ranzi and his coworkers at tar species considered at the gasifier outlet are styrene, naphthalene,
the CRECK modeling group have done a lot of work and proposed phenylacetylene, anthracene, 7,12-dimethylbenz(a)anthracene and its
chemical mechanisms with kinetic parameters for the simulation of isomers, corannulene and its isomers, pyrene and its isomers, biphenyl,
pyrolysis and gasification reactions of biomass [48,49]. It is shown in phenol, toluene, napthol, acenaphthylene and its isomers, indene and
Table 1 that the detailed chemical kinetic model has been adopted in xylene. It may be noted in this context that while the simulated gasifier
the present work for the gas phase reactions in both the pyrolysis and conditions are according to Table 2, the experimental conditions in
gasification zone. However, in the present work, we have optimized the Campoy et al. [50] are the following: biomass feeding rate 15 kg/hr,
detailed CRECK mechanism containing 451 species and 17,848 reac- gasifier bed temperature 805 °C (for ER = 0.27) and biomass feeding
tions (version 1412, available at the website http://creckmodeling. rate 20.5 kg/hr, gasifier bed temperature 780 °C (for ER = 0.19).
chem.polimi.it/) through sensitivity analysis [30] to generate a me- It is observed from the model predictions that the mole fraction (in
chanism to predict the formation of the major gaseous and tar con- percentage) of H2 in the producer gas increases monotonically from
stituents in the producer gas in the parametric range suitable for flui- 13.44% to 16.26% (Fig. 2a) and that of CO from 13.67% to 16.32%
dized bed gasification. The sensitivity analysis has been performed on (Fig. 2b) as the ER is decreased from 0.27 to 0.19. On the other hand,
the mole fractions of the four major constituent species (H2, CO, CO2 the mole fraction of CH4 increases marginally from 4.88% to 5.35%
and CH4) pertaining to the producer gas at the gasifier outlet one at a (Fig. 2c) with the corresponding change of ER. The concentration of
time. Finally, 42 reactions that led to the wrong prediction of the target CO2 first decreases from 16.70% to 15.44% with ER changing from 0.27
species have been identified and removed from the initial mechanism. to 0.24 and then increases to 16.26% as ER decreases further to 0.19
The mechanism with the remaining 17,806 reactions has been used in (Fig. 2d). The simulation results demonstrate a good agreement with
the work. Appendix A of the supplementary material lists the excluded the experimental result of Kim at al. [39] as observed from the figures.
reactions in the present work. The prediction of CO shows a maximum deviation of 5.15% at ER of
The model predictions have been validated against the experimental 0.24 and a minimum of −0.94% at ER of 0.27 from the corresponding
data from Kim et al. [39]. The total height and the internal diameter of experimental results. Likewise, the deviations for all other species and
the gasifier are 3.8 m and 0.4 m, respectively. In the model, the bed at all three ERs remain well within ± 10%. On a similar note, it may
region is considered to extend up to 2 m height above which the free- be observed from Fig. 2e that the present model predicts the tar content
board zone has the height of 1.8 m. The corresponding volumes are quite accurately as well. One reason of the deviation between the ex-
used for the PSR2 and PFR zones. The simulations have been done for perimental data and the numerical prediction, whatever is observed in
three different equivalence ratios (ER = Actual air–fuel ratio/ Stoi- the figure, is the difference in operating conditions between the ex-
chiometric air–fuel ratio) of 0.27, 0.24 and 0.19. The details of biomass periments and simulation.
and air feeding rates at different equivalence ration are shown in It is seen from Fig. 2 that the mole fractions of CO and H2 increase
Table 2. The values in the table correspond to those in the experimental continually with the decrease in equivalence ratio. At a lower equiva-
work of Kim et al. [39]. As seen from the table, the simulation of the lence ratio, as the environment within the gasifier becomes less oxi-
pyrolysis zone has been performed at atmospheric pressure and tem- dizing, more incomplete combustion products, like CO and H2, remain
peratures of 650 °C, 681 °C and 660 °C for ER 0.27, 0.24 and 0.19, in the product gas. Similar trends in the variations of syngas composi-
respectively. The modeling of the gasification zone has also been done tion with the change in ER were observed by earlier researchers [23,51]
in their gasification studies. The opposing trend in the variation of CO2

Table 2
The gasifier conditions used in simulation for comparison with experimental results [39] in Fig. 2 (a-d). All simulations are done at atmospheric pressure.
Case ER Biomass feeding rate (kg/hr) Air flow rate (Nm3/hr) Pyrolysis Temperature (°C) Gasification Temperature (°C)

Case I 0.27 34 37 650 769


Case II 0.24 55 54 681 765
Case III 0.19 55 43 660 739

5
B. Das, et al. Fuel 280 (2020) 118582

(a) (b)
H2 (Present Simulation) CO (Present Simulation)
20 20
H2 (Experiment [39]) CO (Experiment [39])

H2 Mole Fraction (dry vol.%)

CO Mole Fraction (dry vol.%)


15 15

10 10

5 5

0 0
Case I Case II Case III Case I Case II Case III

(c) (d) CH4(Present Simulation)


20 CO2 (Present Simulation)
8
CO2 (Experiment [39]) CH4 (Experiment [39])

CH4 Mole Fraction (dry vol.%)


CO2 Mole Fraction (dry vol.%)

15 6

10 4

5 2

0 0
Case I Case II Case III Case I Case II Case III

(e) 40 Tar (Present Simulation)


Tar (Experiment [50])
Tar Concentration (g/Nm3)

30

20

10

0
ER 0.27 ER 0.19
Fig. 2. Comparions of the predicted mole fractions of syngas components from the present model with the experimental data from the literature [39]; (a) H2, (b) CO,
(c) CO2, (d) CH4 and (e) tar for different equivalence ratios as in table 1.

mole fraction with the change in ER from 0.19 to 0.24 and then from (Fig. 3(a)). On the other hand, the reaction of H2 with phenyl (C6H5),
0.24 to 0.27, observed in the present work, is also in line with the re- vinyl (C2H3) and methoxy (CH3O) radicals are significant in the con-
sults of biomass gasification reported earlier by Lv et al. [52]. Gasifi- sumption of H2. The reaction of cyclopentadiene (CYC5H6) with H ra-
cation temperature along with the ER influences the chemical reactions dical also has a positive reaction rate for the generation of hydrogen,
to alter the concentrations of the components in the producer gas. although to a much lesser extent. Similarly, the reaction of hydrogen
In order to understand the chemical conversion paths enabling the with styrene radical (C6H5C2H2) consumes H2 to some extent. The
formation of the major combustible components of the syngas (H2 and formation of CO mostly takes place through the oxidation of formyl
CO), the rate of production (ROP) analysis has been done in the flui- radical (HCO), showing the maximum rate of production of CO
dized bed zone of the gasifier (i.e. in PSR2) for H2 and CO. The rate of (Fig. 3(b)). Reactions of benzene (C6H6), acetyl radical (CH3CO) and
production of H2 and CO in the major reaction steps, at ER = 0.19, are cyclopentadienyl radical (CYC5H5) with other radicals also produce CO,
shown in Fig. 3(a) and 3(b), respectively. It is to be noted that in the though to a much lesser extent. The reactions of CO with hydroperoxyl
reaction zone, there can be a host of reactions involving a particular radical (HO2) and hydroxyl radical (OH) are significant for CO de-
species. Out of them, some reactions produce the particular species struction. Both these reaction steps lead to the formation of CO2 in the
resulting in a positive rate of production while, other reactions consume bed. All the results show the dominating effects of different radicals
it causing a negative rate of production. We have shown six most sig- towards the production of H2 and CO in the syngas. The temperature
nificant reactions, for each of H2 and CO, in terms of the absolute value existing in the gasifier plays an important role in controlling the ki-
of rate of production in Fig. 3. It is observed that the reaction of for- netics of the reactions and influencing the formation of the syngas
maldehyde (CH2O) with the H radical is the most dominant path in the constituents. In the present simulation, the temperature has been set in
formation of H2, contributing to its maximum rate of production agreement with the experiment. These results indicate the inadequacy

6
B. Das, et al. Fuel 280 (2020) 118582

6. C6H5C2H2+H2=>C6H5C2H3+H
(a) ROP H2 5. H+CYC5H6=>H2+CYC5H5

4. H+CH3OH<=>H2+CH3O

3. C2H3+H2=>C2H4+H

2. H+C6H6<=>H2+C6H5

1. H+CH2O<=>H2+HCO

-0.4 -0.2 0 0.2 0.4 0.6 0.8 1 1.2


Rate of Production (mole/cm3-s)×107

6. HO2+CYC5H5=>C4H5+CO+OH
(b) ROP CO 5. CH3CO+M<=>CH3+CO+M

4. OH+C6H6=>CYC5H6+CO+H

3. CO+OH<=>CO2+H

2. CO+HO2<=>CO2+OH

1. O2+HCO<=>HO2+CO

-3.00 -2.00 -1.00 0.00 1.00 2.00 3.00 4.00


Rate of Production (mole/cm3-s)×107
Fig. 3. Rate of Production for (a) H2 and (b) CO at 739.4° C gasifier temperature and 0.19 ER.

of equilibrium model in simulating the gasifier operation. zone. This is due to the release of the hydrocarbon species during the
devolatilization of biomass and subsequent reactions among them in
3.2. Analysis of tar production the pyrolysis zone. Such observations have been made previously
during biomass pyrolysis studies [53–55]. The concentrations of the tar
The most important contribution of the present work is the evolu- species decrease in the gasification zone through various cracking re-
tion and prediction of tar components in different parts of the gasifier actions [56,57]. Thereafter, not much change in the concentrations has
and finally in the producer gas. A significant amount of tar components been found to occur in the freeboard region. Fig. 4 shows that benzene
is formed in the pyrolysis products following the decomposition of the is having much higher concentrations than any of the other four species
biomass. The tar produced during pyrolysis further undergo cracking in plotted in the figure. This observation is in line with the experimental
the gasification zone and produce a mixture of non-condensable gases, observation made by Dufour et al. [58] for the pyrolysis of solid bio-
light hydrocarbons and unconverted tar fractions. The formation of the mass. The mole fraction of benzene in the pyrolysis zone is 4.8%, which
tar components depends on the gasifier operating parameters such as drops down to 2.3% during the gasification (Fig. 4(a)). Styrene is the
pressure, temperature, equivalence ratio and also on the residence time second most significant species having mole fraction of 0.75% in the
over which the chemical reactions occur. It has been observed in the pyrolysis zone (Fig. 4(c)). The concentrations of toluene (Fig. 4(b)) and
literature that for the operating temperature range of the gasifier, as naphthalene (Fig. 4(e)) in the pyrolysis zone remains at 0.13% and
considered in this work, the predominant tar species are the aromatics 0.17%, respectively, while xylene has mole fraction less than 0.03%
[14]. during the pyrolysis (Fig. 4(d)). The mole fractions of styrene, naph-
thalene, toluene and xylene are 0.17%, 0.07%, 0.025% and 12 ppm,
3.2.1. Identification of main tar forming reactions respectively, after gasification. Finally, not much change in the con-
Benzene (C6H6) is the aromatic species with single aromatic ring centrations of the species occur in the freeboard and the mole fractions
formed during the pyrolysis of biomass. Other aromatic species are of the tar components in the final producer gas remain only slightly
formed from it through a sequence of reactions. Among a host of dif- lower than those in the gasifier. The near constant major species con-
ferent hydrocarbon species in the chemical mechanism, we have iden- centration along the freeboard zone of a bubbling fluidized bed gasifier
tified five species, namely benzene, toluene, styrene, xylene and has been previously reported in the experimental study by Radmanesh
naphthalene, [13,27] and plotted their concentrations in the pyrolysis, et al. [43] as well.
gasification and freeboard regions to show the evolution of the tar The detailed chemical mechanism considered in different zones of
species in the gasifier (Fig. 4(a)-(e)). All the results of Fig. 4 are for the gasifier enables to identify the major reaction steps through which
ER = 0.19 and with the corresponding flow rates listed in Table 2. the important tar species are formed or consumed. As the freeboard
Moreover, to understand the evolution path of the tar species, we have zone does not contribute in changing the concentrations of the tar
also plotted the most significant reaction steps in the pyrolysis and species to any significant extent, only the pyrolysis and gasification
gasification zones causing the formation or destruction of each of the reactions have been focused upon in the present work. Fig. 5(a) and
five tar species as mentioned above (Fig. 5) at the same ER. 5(b) show the ten most significant reactions involving benzene during
The mole fractions of the species, shown in Fig. 4, indicate that the the pyrolysis and gasification, respectively, on the basis of the magni-
concentrations of the tar species remain the highest in the pyrolysis tude of the rate of production of benzene. The formation of benzene in

7
B. Das, et al. Fuel 280 (2020) 118582

6 0.15
(a) Benzene (b) Toluene
5

Mole Fraction (%)

Mole Fraction (%)


4 0.1

2 0.05

0 0
Pyrolysis Gasification Freeboard Pyrolysis Gasification Freeboard

1 0.032
(c) Styrene (d) Xylene
0.8
0.024
Mole Fraction (%)

Mole Fraction (%)


0.6
0.016
0.4

0.008
0.2

0 0
Pyrolysis Gasification Freeboard Pyrolysis Gasification Freeboard

0.2

(e) Napthalene
0.15
Mole Fraction (%)

0.1

0.05

0
Pyrolysis Gasification Freeboard
Fig. 4. Predicted mole fractions of (a) benzene, (b) toluene , (c) styrene, (d) xylene and (e) napthalene at the outlets of different zones in the gasifier at bed
temperature of 739.4° C and ER 0.19.

the pyrolysis zone is mainly contributed by the reactions zone. The analysis of the rate of production of benzene in the gasifi-
C2H3 + C4H4= > C6H6 + H, H + LC6H6 < = > C6H6 + H and cation zone (Fig. 5(b)) indicates the most significant reaction to be that
C6H5 + CH2CO= > C6H6 + HCCO, having production rates of between phenyl group and hydrogen gas to form benzene (Reaction 1).
11.26 × 10-7 mol/cm3-s, 10.14 × 10-7 mol/cm3-s and 9.34 × 10-7 mol/ However, the production rate of benzene in the above reaction is sub-
cm3-s, respectively (Fig. 5(a)). Colket [59] identified the first of the sided by the two main consumption reactions in which CO and CH4 are
above three reactions, between vinyl radical (C2H3) and vinylacetylene formed (Reactions 2 and 3). While the rate of production of benzene in
(C4H4), as the main route to the formation of benzene during pyrolysis reaction 1 occurs at a rate of 0.22 × 10-7 mol/cm3-s, the consumption of
in the prescribed temperature range. The importance of H radical in the benzene through the next two reactions occurs at a cumulative rate of
reactions of benzene is evident from the list of reactions in Fig. 5(a). H 0.27 × 10-7 mol/cm3-s. Likewise, considering the whole set of reactions
radical reacts with toluene (C7H8) to form benzene by separating a of benzene in the gasification zone, the net rate of production is ne-
methyl radical (Reaction 5). Another significant reaction in benzene gative, lowering the concentration of benzene from its value in the
formation (Reaction 6) is the acetylene (C2H2) addition to n-butadienyl pyrolysis zone. The volumetric rate of production of benzene plotted in
(C4H5) followed by cyclization and hydrogen loss as suggested by Cole Fig. 5(a) and 5(b) for the pyrolysis and gasification zones, respectively,
et al. [60]. On the other hand, consumption of benzene occurs during show a two order of magnitude difference in their values. This is be-
pyrolysis to form naphthalene (C10H8) through reactions 4, 8 and 10; cause of the shorter volume of the pyrolysis zone in which the reactions
styrene (C6H5C2H3) through reaction 7 and methane (CH4) through get completed indicating the fast rate of reaction occurring there. Si-
reaction 9. milar order variation in the rates of production of other species has also
In the gasification zone of the fluidized bed, the mole fraction of been observed between the pyrolysis and gasification zones.
benzene drops down considerably compared to that in the pyrolysis The ten important reactions involving toluene in the pyrolysis and

8
B. Das, et al. Fuel 280 (2020) 118582

(a) Pyrolysis 10. C4H3+C6H6=>C10H8+H


9. CH3+C6H6=>CH4+C6H5
8. CH2C3H5+C6H6=>2H2+H+C10H8
7. H+C6H5C2H3<=>C6H6+C2H3
6. C4H5+C2H2<=>C6H6+H
5. H+C7H8<=>C6H6+CH3
4. C4H5+C6H6=>C10H8+H2+H
3. C6H5+CH2CO=>C6H6+HCCO
2. H+LC6H6<=>C6H6+H
1. C2H3+C4H4=>C6H6+H

-10.00 -5.00 0.00 5.00 10.00 15.00


Rate of Production (mole/cm3-s)×107

(b) Gasification 10. C6H5+CH2O=>C6H6+HCO


9. HO2+C6H6=>H2O2+C6H5
8. H+LC6H6<=>C6H6+H
7. C2H3+C4H4=>C6H6+H
6. C6H5+CH4=>C6H6+CH3
5. H+C6H5OH<=>C6H6+OH
4. C4H5+C6H6=>C10H8+H2+H
3. CH3+C6H6=>CH4+C6H5
2. OH+C6H6=>CYC5H6+CO+H
1. H+C6H6<=>H2+C6H5

-0.15 -0.10 -0.05 0.00 0.05 0.10 0.15 0.20 0.25


Rate of Production (mole/cm3-s)×107

(c) Pyrolysis 10. C7H8<=>CH3+C6H5


9. C2H3+C7H8=>C2H4+C7H7
8. C6H4CH3+H2=>C7H8+H
7. CH3+C7H8<=>CH4+C7H7
6. H+C7H8=>H2+C6H4CH3
5. H+XYLENE<=>C7H8+CH3
4. 2SC4H7=>CH3+H+C7H8+H2
3. H+C7H8=>H2+C7H7
2. H+C7H8<=>C6H6+CH3
1. C7H8<=>C7H7+H

-8.00 -6.00 -4.00 -2.00 0.00 2.00 4.00 6.00 8.00


Rate of Production (mole/cm3-s)×107
Fig. 5. Rate of production in the pyrolysis and gasification zone for major tar components; (a) & (b) benzene (c) & (d) toluene, (e) & (f) styrene, (g) & (h) xylene and
(i) & (j) naphthalene at bed temperature of 739.4° C and ER 0.19.

gasification zones, in terms of the magnitude of rate of production, are the pyrolysis zone and the toluene mole fraction at the exit of the
shown in Fig. 5(c) and 5(d), respectively. It is observed that the first pyrolysis zone remains at 0.13%. The mole fraction of toluene drops
reaction in the pyrolysis zone produces toluene while the next two considerably (by 82%) in the gasification zone as toluene reacts with
important reactions consume toluene (Fig. 5(c)). Reactions 4 and 5 of different radicals like H, CH3, OH and HO2. H-abstraction reactions are
the pyrolysis zone again produce toluene. The net contribution of the found to play important role in many of the destruction reactions of
ten significant toluene reactions in the pyrolysis zone is a rate of pro- toluene with the radicals. The cracking reactions of toluene in the ga-
duction of 1.09 × 10-7 mol/cm3-s, while the other reactions are even sification zone results in the formation of CH4 and H2 as in reactions 2
less contributing. This rate of production is much smaller in comparison and 3, respectively, (Fig. 5(d)).
to that of benzene for which the net rate of production of the ten im- The formation of styrene (C6H5C2H3) in the pyrolysis zone mainly
portant reactions in the pyrolysis zone is 21.18 × 10-7 mol/cm3-s. As a occurs through the reaction of benzene with vinyl radical (C2H3)
result, the toluene concentration never increases to a very high value in marked as shown in reaction 1 of Fig. 5(e). The reactions

9
B. Das, et al. Fuel 280 (2020) 118582

(d) Gasification 10. OH+C7H8=>C6H4CH3+H2O


9. H+C7H8=>H2+C6H4CH3
8. C7H8<=>C7H7+H
7. H+C7H8<=>C6H6+CH3
6. HO2+C7H8=>H2O2+C7H7
5. OH+C7H8=>C7H7+H2O
4. C6H4CH3+H2=>C7H8+H
3. H+C7H8=>H2+C7H7
2. CH3+C7H8<=>CH4+C7H7
1. O2+C7H8<=>HO2+C7H7

-0.02 -0.01 -0.01 0.00 0.01 0.01 0.02


Rate of Production (mole/cm3-s)×107

(e) Pyrolysis 5. C2H3+C6H5C2H3=>C2H4+C6H5C2H2


4. C6H5C2H2+CH2CO=>C6H5C2H3+HCCO
3. C4H5+C4H4=>C6H5C2H3+H
2. H+C6H5C2H3=>H2+C6H5C2H2
1. H+C6H5C2H3<=>C6H6+C2H3

-2.00 -1.00 0.00 1.00 2.00 3.00 4.00 5.00


Rate of Production (mole/cm3-s)×107

(f) Gasification 5. CH3+C6H5C2H3=>CH4+C6H5C2H2


4. C6H5C2H2+C6H6=>C6H5C2H3+C6H5
3. HO2+C10H7OH=>C6H5C2H3+2CO+OH
2. C6H5C2H2+H2=>C6H5C2H3+H
1. OH+C6H5C2H3=>H2O+C6H5C2H2

-0.16 -0.12 -0.08 -0.04 0.00 0.04 0.08


Rate of Production (mole/cm3-s)×107

(g) Pyrolysis 5. CH3+XYLENE=>CH4+RXYLENE


4. H+XYLENE<=>C7H8+CH3
3. 2SC4H7=>2H+XYLENE+H2
2. H+XYLENE=>H2+RXYLENE
1. XYLENE<=>RXYLENE+H

-6.00 -4.00 -2.00 0.00 2.00 4.00 6.00 8.00


Rate of Production (mole/cm3-s)×107
Fig. 5. (continued)

C4H5 + C4H4= > C6H5C2H3 + H and other reactions for which the styrene concentration falls considerably in
C6H5C2H2 + CH2CO= > C6H5C2H3 + HCCO also contribute towards the gasification zone.
the formation of styrene, though to a lesser extent than the first reac- The formation rate of xylene in the pyrolysis zone is low and most of
tion. Styrene is consumed in the pyrolysis zone through its reaction the xylene formed there is cracked in the gasification zone by H, OH
with H-radical (reaction 2) and vinyl radical (reaction 5). However, the and CH3 radicals (Fig. 5(g) and 5(h)). Only 3% of the xylene produced
rate of consumption of styrene in these reactions remains far less than in the pyrolysis, is present at the gasifier outlet. Naphthalene is another
its rate of generation in reaction 1. Thus, the styrene concentration in important tar compound formed in the pyrolysis zone mainly from the
the pyrolyzed product remains at 0.75%, which is higher than toluene single ring aromatic benzene (Fig. 5(i)) and also from tetralin and
and naphthalene. Styrene mole fraction drops down to 0.17% in the naphthalene radicals. Naphthalene also gets consumed through the
gasification zone mainly due to its reaction with OH radical producing reaction with H radical in this zone. Combination of all these reactions
water vapour and C6H5C2H2 (Fig. 5(f)). The consumption rate of leaves the pyrolyzed gas with 0.16% naphthalene, which drops down in
styrene in this reaction far outweighs the formation of rate of styrene in the gasification zone to near about 0.06%. The important cracking

10
B. Das, et al. Fuel 280 (2020) 118582

(h) Gasification 5. CH3+XYLENE=>CH4+RXYLENE


4. H+XYLENE<=>C7H8+CH3
3. O2+XYLENE<=>HO2+RXYLENE
2. OH+XYLENE=>H2O+RXYLENE
1. H+XYLENE=>H2+RXYLENE

-4.00 -3.00 -2.00 -1.00 0.00 1.00 2.00 3.00


Rate of Production (mole/cm3-s)×1010

(i) Pyrolysis 10. C2H3+C6H5C2H=>C10H8+H


9. C10H7+CH2CO=>C10H8+HCCO
8. H+C10H8=>H2+C10H7
7. C4H3+C6H6=>C10H8+H
6. CH2C3H5+C6H6=>2H2+H+C10H8
5. C10H7+H2=>C10H8+H
4. RTETRALIN=>C10H8+H2+H
3. C4H5+C6H6=>C10H8+H2+H
2. H+C10H8=>0.5RTETRALIN+0.5C10H7
1. H+C10H8=>C6H5+C4H4

-15.00 -10.00 -5.00 0.00 5.00 10.00


Rate of Production (mole/cm3-s)×107

(j) Gasification 10. C4H3+C6H6=>C10H8+H


9. C4H2+C6H6=>C10H8
8. H+C10H8=>H2+C10H7
7. O2+C14H9=>C10H8+2CO+C2H
6. OH+BIN1A=>0.4H2+0.2C10H8+0.85BIN1A+HCO
5. C10H7+H2=>C10H8+H
4. C4H5+C6H6=>C10H8+H2+H
3. RTETRALIN=>C10H8+H2+H
2. H+C10H8=>0.5RTETRALIN+0.5C10H7
1. H+C10H8=>C6H5+C4H4

-0.06 -0.04 -0.02 0.00 0.02 0.04


Rate of Production (mole/cm3-s)×107
Fig. 5. (continued)

reactions contributing to the fall in naphthalene concentration in the alone. The gasifier temperatures have been maintained in line with the
gasification zone are again affected by H radicals (Fig. 5(j)). experiments of Kim et al. [39] and are fixed at 769.3 °C, 765 °C and
739.4 °C, respectively, for ER = 0.27, 0.24 and 0.19. It is observed from
the figure that with the decrease in equivalence ratio, as the gasifier
3.2.2. Effects of equivalence ratio on tar species
environment becomes richer in fuel, the mole fractions of most of the
The effect of equivalence ratio on the concentration of selected tar
tar species increase. This happens due to less availability of oxygen to
species in the producer gas formed in the gasification of woody biomass
react with the tar compounds inside the gasifier. The mole fraction of
in a bubbling fluidized bed gasifier is shown in Fig. 6(a)- (f). The mole
benzene far exceeds the other hydrocarbons in the producer gas as
fractions of benzene, toluene, xylene, styrene, naphthalene, a few 2–3
observed from the figure. The mole fraction of benzene in the producer
rings aromatics, and a few other 4–7 rings aromatics are plotted in the
gas at ER = 0.27 is 1.8%, while at ER = 0.19 the mole fraction of it
figure for three equivalence ratio values (ER = 0.27, 0.24 and 0.19).
goes up to 2.3%. Styrene and naphthalene are two most abundant
For the comparative study, the equivalence ratio is varied only by
species after benzene. While styrene mole fraction at ER = 0.19 is
changing the air flow rate while keeping a constant biomass feeding
0.16% that of naphthalene at the corresponding condition is 0.07%.
rate of 55 kg/hr. Thus, the operating conditions for the results in Fig. 6
Toluene is found to have a low concentration. Interestingly, some of the
do not fully agree with those in Table 2, where for the ER = 0.27 case,
2–3 ring and 4–7 ring tar components show a decrease in their mole
the biomass feeding rate is different. We have done this as biomass feed
fractions when the ER is changed from 0.27 to 0.24; but their con-
rate is also an influencing parameter on tar formation and varying it
centrations increase when the environment in the gasifier is made
along with the equivalence ratio will not indicate the effect of the latter

11
B. Das, et al. Fuel 280 (2020) 118582

3.00 0.2
(a) Benzene (b) Styrene

1- Ring Tar Mole Fraction (%)


Benzene Mole Fraction(%)
2.50 Toluene
0.16
2.00
0.12
1.50
0.08
1.00

0.50 0.04

0.00 0
ER 0.27 ER 0.24 ER 0.19 ER 0.27 ER 0.24 ER 0.19

3.0E-4 0.1
(c) (d) Napthalene

2-3 Ring Tar Mole Fraction (%)


Xylene Phenanthrene and anthracene
2.5E-4
Xylene Mole Fraction (%)

0.08 Biphenyl
2.0E-4
1.5E-4 0.06

1.0E-4 0.04
5.0E-5
0.02
5.0E-19
ER 0.27 ER 0.24 ER 0.19
-5.0E-5 0
ER 0.27 ER 0.24 ER 0.19

0.01 0.06
(e) Acenaphthalene and isomers (f) Corannulene and isomers
4-7 Ring Tar Mole Fraction (%)
2-3 Ring Tar Mole Fraction (%)

Fluorene
0.008 0.05 7,12-Dimethylbenz[a]anthracene
Diphenylmethane
and isomers
0.04 Pyrene and isomers
0.006
0.03
0.004
0.02
0.002
0.01

0 0
ER 0.27 ER 0.24 ER 0.19 ER 0.27 ER 0.24 ER 0.19
Fig. 6. Predicted mole fraction of tars; (a) benzene, (b) styrene & toluene, (c) xylene, (d) &(e) 2–3 ring tars, and (f) 4–7 ring tars at different equivalence ratios with
fixed biomass flow rate of 55 kg/hr and gasifier temperature of 769.3 °C, 765 °C and 739.4 °C for ER of 0.27, 0.24 and 0.19 respectively.

richer with ER = 0.19. This trend is similar to that of CO2 in the pro- computed for the three ER values to observe the variation. The aromatic
ducer gas with the variation in ER and can be attributed to the re- species with heavier molecular weight than benzene have been con-
sidence time and oxidizing atmosphere within the gasification zone of sidered in the total quantity of tar due to the ability of benzene to retain
the reactor. It is to be mentioned here that in the present computation its gaseous form at the exit of the gasifier [27]. The tar species con-
the temperature of the gasifier has been defined in each case without sidered here are the same as in Fig. 2 (e). The tar concentration in the
computing the value from the energy conservation equation at the gas at ER = 0.19 is found to be 29.6 g/Nm3. However, the corre-
corresponding ER. The gasification air flow rate through the bed is sponding values at ER = 0.24 and 0.27 are evaluated as 18.97 g/Nm3
lower at lower ER, maintaining a less oxidizing (i.e. richer) environ- and 18.76 g/Nm3, respectively. This nature of variation of the total tar
ment in the bed. The kinetic rates of reactions change accordingly due with ER is consistent with the experimental results of Kinoshita et al.
to the change in concentrations of the reacting species. At the same [62].
time, the reduced flow rate of air through the bed increases the re-
sidence time of the species within the bed enabling them to participate
3.3. Overall gasifier performance
in chemical reactions over longer duration. The combined effect of
these two results in non-monotonic variation of the mole fraction of
Table 3 gives a general description of the important performance
certain tar species with the change in ER. Similar trends in the results
characteristics of the gasification process predicted from the model. The
were earlier reported from the fluidized bed gasification study of
experimental data from Kim et al. [39] have also been shown for
sawdust biomass by Horvat et al. [61].
comparison at ER = 0.24 and 0.19 and good agreements are observed.
The total tar concentration formed in the gas (g/Nm3) has been
The results of ER = 0.27 have not been compared as it has already been

12
B. Das, et al. Fuel 280 (2020) 118582

Table 3
Gasifier performance at different equivalence ratios (ER) for biomass flow rate of 55 kg/hr and bed temperature of 769.3 °C, 765 °C and 739.4 °C for ER of 0.27, 0.24
and 0.19 respectively. The experimental data [39] at identical conditions is shown for comparison at ER = 0.24 and 0.19.
Parameter ER 0.27 ER 0.24 ER 0.19
Present simulation Present simulation Kim et al. [39] Present simulation Kim et al. [39]

CO (Volume %, dry) 13.97 15.77 15 16.32 16.1


CO2 (Volume %, dry) 16.10 15.44 16 15.93 16.4
H2 (Volume %, dry) 13.97 15.56 16.1 16.26 16.5
CH4 (Volume %, dry) 4.48 4.79 4.6 5.35 5.3
Tar (gm/Nm3) 18.77 18.97 – 29.60 –
LHV (MJ/Nm3) 5.33 5.88 5.3 6.26 5.7
Dry gas production (Nm3/kg fuel) 2.18 2.09 – 1.84 –
CCE (%) 82.06 82.57 – 76.24 –
CGE (%) 64.62 68.42 – 64.02 –

mentioned in Section 3.2.2 that the biomass feeding rate for this con- stirred reactor (PSR1). Subsequently, the gasification zone and free-
dition has been changed from that of Kim et al. [39]. The results show board have been modeled using a second perfectly stirred reator (PSR2)
that the combustible gaseous components in the producer gas, namely and a plug flow reactor (PFR). The kinetic scheme of biomass proposed
CO, H2 and CH4, increase with the decrease in the ER. On the other by the CRECK group has been used after eliminating 42 reactions
hand, the tar concentration in the producer gas increases considerably identified through a sensitivity analysis. The composition of the pro-
at ER = 0.19, compared to the other two cases. Lower Heating Value ducer gas has been compared against the experimental data from the
(LHV) of the producer gas is an important indicator of the quality of the literature and a good agreement has been reached.
fuel produced. It is observed from the results in Table 3 that the LHV of The following conclusions can be drawn from the study.
the fuel increases with lower ER, due to increased concentration of the
combustible components in the gas. On the contrary, lower ER leads to • Active radicals play an important role in determining the rates of
lower rate of dry producer gas generation. The carbon conversion ef- different gasification reactions in the fluidized bed. Kinetic para-
ficiency (CCE), representing the fraction of the number of carbon moles meters of the reaction equations decide the composition of the
in the biomass feedstock which has been transferred to the gaseous producer gas mixture.
components of the producer gas, decreases at ER = 0.19. It is to be • Hydrocarbon species of tar are formed in the pyrolysis zone and
mentioned here that in calculating the carbon conversion efficiency we have their highest concentration there. Thereafter, the tar con-
have considered the carbon in CO, CO2, CH4, C2H4 and C2H6 in the centrations fall in the gasification zone of the fluidized bed. Not
producer gas from the carbon in biomass. The reduction in carbon much change in the concentrations of species is observed in the
conversion efficiency at ER = 0.19 is due to the carbon transferred in freeboard zone of the gasifier.
the increased quantities of tar compounds. • Benzene is having the maximum concentrations among different
Cold gas efficiency (CGE = dry gas generation, Nm3/kg fuel × LHV aromatic hydrocarbons. Styrene and naphthalene are the next two
of producer gas, MJ/Nm3/LHV of Biomass feedstock, MJ/kg) defines significant hydrocarbons. The concentrations of different hydro-
the fraction of the stored energy in the biomass which is transferred to carbons depend on their formation and destruction rates in various
the producer gas fuel on gasification. The results show that among the zones. The rate of reaction in the pyrolysis zone is significantly
three ER values considered, the maximum cold gas efficiency is ob- larger than that in the gasification zone.
tained at the middle value of ER = 0.24. It shows that there is an op- • For most of the tar species, the concentrations decrease continually
timum equivalence ratio of operation of the gasifier which will give the with the increase in ER. The oxidizing atmosphere created at higher
most efficient energy conversion performance. It may be mentioned in ER reduces the concentration of the hydrocarbons.
this regard, that the presence of optimum equivalence ratio in the • However, for some 4–7 ring aromatic species (such as pyrene and its
context of cold gas efficiency of bubbling fluidized bed gasifiers has isomers, corannulene and its isomers and 7,12-dimethylbenz(a)an-
been reported before in both theoretical [63] and experimental [64] thracene), the concentration becomes the minimum at an optimum
studies. ER in between. Such a non-monotonic trend with ER originates from
It may be summarized from the above results that the detailed the combined effects of changes in chemical kinetics and residence
chemical kinetic analysis of biomass gasification process in bubbling time inside the reactor.
fluidized bed reactors brings about the identification of important re- • The tar concentration in the producer gas drops considerably with
action pathways leading to the prediction of syngas composition. the increase in ER from 0.19 to 024. However, with further increase
Therefore, the present study provides accurately validated explanations in ER to 0.27, the drop in tar concentration is only marginal.
on the formation and destruction routes of major pyrolysis and gasifi- • Increase in ER from 0.19 to 0.27 decreases the LHV of the producer
cation products like H2, CO and tar. The approach developed in this gas due to less concentrations of the combustible components, like
work is the first of its kind and is expected to provide more accurate H2, CO and CH4, in the oxidizing environment of the gasifier.
chemical information on tar formation in bubbling fluidized bed bio- However, the dry gas production rate increases as the ER is in-
mass gasifiers. Moreover, as an additional advantage, the proposed creased. The cold gas efficiency is found to be the maximum at
approach demands very little amount of computational resource due to ER = 0.24 indicating it to be the optimum equivalence ratio leading
the exclusion of detailed fluid dynamic analysis. to the best performance of the gasifier.
• Finally, the proposed modelling approach provides accurate and
4. Conclusions computationally inexpensive way of evaluating major products like
H2, CO and tar constituents in the fluidized bed biomass gasification
A kinetic model of fluidized bed gasifier has been developed based process.
on an Equivalent Reactor Network (ERN). The gasifier is divided into
three zones – pyrolysis zone, gasification zone in the bed and freeboad CRediT authorship contribution statement
zone. Pyrolysis of the biomass is modeled with a thermodynamic
equilibrium model followed by a kinetic model applied in a perfectly Bijoy Das: Methodology, Software, Formal analysis, Validation,

13
B. Das, et al. Fuel 280 (2020) 118582

Writing - original draft. Atmadeep Bhattacharya: Methodology, 2002;81:843–53. https://doi.org/10.1016/S0016-2361(01)00216-2.


Resources, Data curation, Writing - original draft. Amitava Datta: [19] Qin YH, Feng J, Li WY. Formation of tar and its characterization during air-steam
gasification of sawdust in a fluidized bed reactor. Fuel 2010;89:1344–7. https://doi.
Conceptualization, Writing - review & editing, Supervision. org/10.1016/j.fuel.2009.08.009.
[20] Gilbert P, Ryu C, Sharifi V, Swithenbank J. Tar reduction in pyrolysis vapours from
Declaration of Competing Interest biomass over a hot char bed. Bioresour Technol 2009;100:6045–51. https://doi.
org/10.1016/j.biortech.2009.06.041.
[21] de Jong W, Ünal Ö, Andries J, Hein KRG, Spliethoff H. Thermochemical conversion
The authors declare that they have no known competing financial of brown coal and biomass in a pressurised fluidised bed gasifier with hot gas fil-
tration using ceramic channel filters: Measurements and gasifier modelling. Appl
interests or personal relationships that could have appeared to influ- Energy 2003;74:425–37. https://doi.org/10.1016/S0306-2619(02)00197-6.
ence the work reported in this paper. [22] Salem AM, Paul MC. An integrated kinetic model for downdraft gasifier based on a
novel approach that optimises the reduction zone of gasifier. Biomass Bioenergy
2018;109:172–81. https://doi.org/10.1016/j.biombioe.2017.12.030.
Acknowledgments
[23] Barman NS, Ghosh S, De S. Gasification of biomass in a fixed bed downdraft gasifier
- A realistic model including tar. Bioresour Technol 2012;107:505–11. https://doi.
This research did not receive any specific grant from funding org/10.1016/j.biortech.2011.12.124.
agencies in the public, commercial, or not-for-profit sectors. [24] Li XT, Grace JR, Lim CJ, Watkinson AP, Chen HP, Kim JR. Biomass gasification in a
circulating fluidized bed. Biomass Bioenergy 2004;26:171–93. https://doi.org/10.
1016/S0961-9534(03)00084-9.
Appendix A. Supplementary data [25] Gambarotta A, Morini M, Zubani A. A non-stoichiometric equilibrium model for the
simulation of the biomass gasification process. Appl Energy 2018;227:119–27.
https://doi.org/10.1016/j.apenergy.2017.07.135.
Supplementary data to this article can be found online at https:// [26] Corella J, Caballero MA, Aznar MP, Brage C. Two advanced models for the kinetics
doi.org/10.1016/j.fuel.2020.118582. of the variation of the tar composition in its catalytic elimination in biomass gasi-
fication. Ind Eng Chem Res 2003;42:3001–11. https://doi.org/10.1021/ie020401i.
[27] Palma CF. Model for biomass gasification including tar formation and evolution.
References Energy Fuels 2013;27:2693–702. https://doi.org/10.1021/ef4004297.
[28] Ismail TM, Abd El-Salam M, Monteiro E, Rouboa A. Eulerian - Eulerian CFD model
[1] Jackson RW, Neto ABF, Erfanian E. Woody biomass processing: Potential economic on fluidized bed gasifier using coffee husks as fuel. Appl Therm Eng
impacts on rural regions. Energy Policy 2018;115:66–77. https://doi.org/10.1016/ 2016;106:1391–402. https://doi.org/10.1016/j.applthermaleng.2016.06.102.
j.enpol.2018.01.001. [29] Ku X, Li T, Løvås T. CFD-DEM simulation of biomass gasification with steam in a
[2] Kumar A, Cameron JB, Flynn PC. Pipeline transport and simultaneous sacchar- fluidized bed reactor. Chem Eng Sci 2015;122:270–83. https://doi.org/10.1016/j.
ification of corn stover. Bioresour Technol 2005;96:819–29. https://doi.org/10. ces.2014.08.045.
1016/j.biortech.2004.07.007. [30] Lu T, Law CK. Toward accommodating realistic fuel chemistry in large-scale com-
[3] Moon JH, Lee JW, Do LU. Economic analysis of biomass power generation schemes putations. Prog Energy Combust Sci 2009;35:192–215. https://doi.org/10.1016/j.
under renewable energy initiative with Renewable Portfolio Standards (RPS) in pecs.2008.10.002.
Korea. Bioresour Technol 2011;102:9550–7. https://doi.org/10.1016/j.biortech. [31] Gómez-Barea A, Leckner B. Modeling of biomass gasification in fluidized bed. Prog
2011.07.041. Energy Combust Sci 2010;36:444–509. https://doi.org/10.1016/j.pecs.2009.12.
[4] Sikarwar VS, Zhao M, Clough P, Yao J, Zhong X, Memon MZ, et al. An overview of 002.
advances in biomass gasification. Energy Environ Sci 2016;9:2339–977. https://doi. [32] Stark AK, Bates RB, Zhao Z, Ghoniem AF. Prediction and validation of major gas
org/10.1039/c6ee00935b. and tar species from a reactor network model of air-blown fluidized bed biomass
[5] Ahmed OY, Ries MJ, Northrop WF. Emissions factors from distributed, small-scale gasification. Energy Fuels 2015;29:2437–52. https://doi.org/10.1021/ef5027955.
biomass gasification power generation: Comparison to open burning and large-scale [33] Lee DH, Yang H, Yan R, Liang DT. Prediction of gaseous products from biomass
biomass power generation. Atmos Environ 2019;200:221–7. https://doi.org/10. pyrolysis through combined kinetic and thermodynamic simulations. Fuel
1016/j.atmosenv.2018.12.024. 2007;86:410–7. https://doi.org/10.1016/j.fuel.2006.07.020.
[6] Parraga J, Khalilpour KR, Vassallo A. Polygeneration with biomass-integrated ga- [34] Glarborg P, Kee RJ, Grcar JF, Miller JA. PSR: A FORTRAN program for modeling
sification combined cycle process: Review and prospective. Renew Sustain Energy well-stirred reactors. Sandia Report SAND86-8209. 1986.
Rev 2018;92:219–34. https://doi.org/10.1016/j.rser.2018.04.055. [35] Dupont C, Boissonnet G, Seiler JM, Gauthier P, Schweich D. Study about the kinetic
[7] Azimov U, Tomita E, Kawahara N, Harada Y. Effect of syngas composition on processes of biomass steam gasification. Fuel 2007;86:32–40. https://doi.org/10.
combustion and exhaust emission characteristics in a pilot-ignited dual-fuel engine 1016/j.fuel.2006.06.011.
operated in PREMIER combustion mode. Int J Hydrogen Energy 2011;36:11985–96. [36] Hafnera S, Rashidia A, Baldeaa G, Riedela U. A detailed chemical kinetic model of
https://doi.org/10.1016/j.ijhydene.2011.04.192. high-temperature ethylene glycol gasification. Combust Theory Model
[8] Kim YS, Lee JJ, Kim TS, Sohn JL. Effects of syngas type on the operation and per- 2011;15:517–35. https://doi.org/10.1080/13647830.2010.547602.
formance of a gas turbine in integrated gasification combined cycle. Energy Convers [37] Qiao L, Xu J, Sane A, Gore J. Multiphysics modeling of carbon gasification processes
Manag 2011;52:2262–71. https://doi.org/10.1016/j.enconman.2011.01.009. in a well-stirred reactor with detailed gas-phase chemistry. Combust Flame
[9] Oesch P, Leppämäki E, Ståhlberg P. Sampling and characterization of high-mole- 2012;159:1693–707. https://doi.org/10.1016/j.combustflame.2011.12.002.
cular-weight polyaromatic tar compounds formed in the pressurized fluidized-bed [38] Khanshan FS, West RH. Developing detailed kinetic models of syngas production
gasification of biomass. Fuel 1996;75:1406–12. https://doi.org/10.1016/0016- from bio-oil gasification using Reaction Mechanism Generator (RMG). Fuel
2361(96)00113-5. 2016;163:25–33. https://doi.org/10.1016/j.fuel.2015.09.031.
[10] Milne TA, Abatzaglou N, Evans RJ. Biomass Gasifier “ Tars ”: Their Nature , [39] Kim YD, Yang CW, Kim BJ, Kim KS, Lee JW, Moon JH, et al. Air-blown gasification
Formation , and Conversion, Report No. NREL/TP-570-25357. 1998. https://doi. of woody biomass in a bubbling fluidized bed gasifier. Appl Energy
org/10.2172/3726. 2013;112:414–20. https://doi.org/10.1016/j.apenergy.2013.03.072.
[11] Evans RJ, Milne TA. Molecular characterization of pyrolysis of biomass. 1. [40] Soria-Verdugo A, Von Berg L, Serrano D, Hochenauer C, Scharler R, Anca-Couce A.
Fundamentals. Energy Fuels 1987;1:123–37. Effect of bed material density on the performance of steam gasification of biomass
[12] Islam MW. A review of dolomite catalyst for biomass gasification tar removal. Fuel in bubbling fluidized beds. Fuel 2019;257:116118https://doi.org/10.1016/j.fuel.
2020;267:117095https://doi.org/10.1016/j.fuel.2020.117095. 2019.116118.
[13] Palma CF. Modelling of tar formation and evolution for biomass gasification: A [41] Zuber C, Husmann M, Schroettner H, Hochenauer C, Kienberger T. Investigation of
review. Appl Energy 2013;111:129–41. https://doi.org/10.1016/j.apenergy.2013. sulfidation and regeneration of a ZnO-adsorbent used in a biomass tar removal
04.082. process based on catalytic steam reforming. Fuel 2015;153:143–53. https://doi.
[14] Rios MLV, González AM, Lora EES, del Olmo OAA. Reduction of tar generated org/10.1016/j.fuel.2015.02.110.
during biomass gasification: A review. Biomass Bioenergy 2018;108:345–70. [42] Basu P. Gasification Theory and Modeling of Gasifiers. In: Biomass Gasification and
https://doi.org/10.1016/j.biombioe.2017.12.002. Pyrolysis: Practical Design, Academic Press; 2010, p. 117–65. https://doi.org/
[15] Valin S, Ravel S, Pons de Vincent P, Thiery S, Miller H. Fluidized bed air gasification https://doi.org/10.1016/B978-0-12-374988-8.00005-2.
of solid recovered fuel and woody biomass: Influence of experimental conditions on [43] Radmanesh R, Cnaouki J, Guy C. Biomass gasification in a bubbling fluidized bed
product gas and pollutant release. Fuel 2019;242:664–72. https://doi.org/10.1016/ reactor: Experiments and modeling. AIChE J 2006;52:4258–72. https://doi.org/10.
j.fuel.2019.01.094. 1002/aic.11020.
[16] Alauddin ZABZ, Lahijani P, Mohammadi M, Mohamed AR. Gasification of lig- [44] Roine A. Outotec HSC chemistry 6.0 user’s guide: chemical reaction and equili-
nocellulosic biomass in fluidized beds for renewable energy development: A review. brium software with extensive thermochemical database and flowsheet simulation.
Renew Sustain Energy Rev 2010;14:2852–62. https://doi.org/10.1016/j.rser.2010. HSC chemistry 6.0, Outotec Research Oy (Previously Outokumpu Research Oy),
07.026. Finland, ISBN: 952-9507-12-7. 2006.
[17] Kaushal P, Abedi J, Mahinpey N. A comprehensive mathematical model for biomass [45] Yan R, Yang H, Chin T, Liang DT, Chen H, Zheng C. Influence of temperature on the
gasification in a bubbling fluidized bed reactor. Fuel 2010;89:3650–61. https://doi. distribution of gaseous products from pyrolyzing palm oil wastes. Combust Flame
org/10.1016/j.fuel.2010.07.036. 2005;142:24–32. https://doi.org/10.1016/j.combustflame.2005.02.005.
[18] Morf P, Hasler P, Nussbaumer T. Mechanisms and kinetics of homogeneous sec- [46] Turns SR. Introduction to combustion (Vol. 287). McGraw-Hill Companies 1996.
ondary reactions of tar from continuous pyrolysis of wood chips. Fuel [47] Lutz AE, Kee RJ, Miller JA. SENKIN: A FORTRAN program for predicting

14
B. Das, et al. Fuel 280 (2020) 118582

homogeneous gas phase chemical kinetics with sensitivity analysis (No. SAND-87- gasification/pyrolysis: An overview. Renew Sustain Energy Rev 2008;12:397–416.
8248). Sandia National Labs., Livermore, CA (USA). 1988. https://doi.org/10.1016/j.rser.2006.07.015.
[48] Ranzi E, Cuoci A, Faravelli T, Frassoldati A, Migliavacca G, Pierucci S, et al. [57] Bridgwater AV. The technical and economic feasibility of biomass gasification for
Chemical kinetics of biomass pyrolysis. Energy Fuels 2008;22:4292–300. https:// power generation. Fuel 1995;74:631–53. https://doi.org/10.1016/0016-2361(95)
doi.org/10.1021/ef800551t. 00001-L.
[49] Sommariva S, Grana R, Maffei T, Pierucci S, Ranzi E. A kinetic approach to the [58] Dufour A, Masson E, Girods P, Rogaume Y, Zoulalian A. Evolution of aromatic tar
mathematical model of fixed bed gasifiers. Comput Chem Eng 2011;35:928–35. composition in relation to methane and ethylene from biomass pyrolysis-gasifica-
https://doi.org/10.1016/j.compchemeng.2011.01.036. tion. Energy Fuels 2011;25:4182–9. https://doi.org/10.1021/ef200846g.
[50] Campoy M, Gómez-Barea A, Villanueva AL, Ollero P. Air-steam gasification of [59] Colket MB. The pyrolysis of acetylene and vinylacetylene in a single-pulse shock
biomass in a fluidized bed under simulated autothermal and adiabatic conditions. tube. Symp Combust 1988;21:851–64. https://doi.org/10.1016/S0082-0784(88)
Ind Eng Chem Res 2008;47:5957–65. https://doi.org/10.1021/ie800220t. 80317-5.
[51] Ghassemi H, Shahsavan-Markadeh R. Effects of various operational parameters on [60] Cole JA, Bittner JD, Longwell JP, Howard JB. Formation mechanisms of aromatic
biomass gasification process; A modified equilibrium model. Energy Convers Manag compounds in aliphatic flames. Combust Flame 1984;56:51–70. https://doi.org/10.
2014;79:18–24. https://doi.org/10.1016/j.enconman.2013.12.007. 1016/0010-2180(84)90005-1.
[52] Lv PM, Xiong ZH, Chang J, Wu CZ, Chen Y, Zhu JX. An experimental study on [61] Horvat A, Kwapinska M, Xue G, Rabou LPLM, Pandey DS, Kwapinski W, et al. Tars
biomass air-steam gasification in a fluidized bed. Bioresour Technol from Fluidized Bed Gasification of Raw and Torrefied Miscanthus x giganteus.
2004;95:95–101. https://doi.org/10.1016/j.biortech.2004.02.003. Energy Fuels 2016;30:5693–704. https://doi.org/10.1021/acs.energyfuels.
[53] Liu Peng, Wang Yue, Zhou Zhengzhong, Yuan Haoran, Zheng Tao, Chen Yilu. Effect 6b00532.
of carbon structure on hydrogen release derived from different biomass pyrolysis. [62] Kinoshita CM, Wang Y, Zhou J. Tar formation under different biomass gasification
Fuel 2020;271:117638. https://doi.org/10.1016/j.fuel.2020.117638. conditions. J Anal Appl Pyrolysis 1994;29:169–81. https://doi.org/10.1016/0165-
[54] Senneca O, Cerciello F, Russo C, Wütscher A, Muhler M, Apicella B. Thermal 2370(94)00796-9.
treatment of lignin, cellulose and hemicellulose in nitrogen and carbon dioxide. [63] Beheshti SM, Ghassemi H, Shahsavan-Markadeh R. Process simulation of biomass
Fuel 2020;271:117656https://doi.org/10.1016/j.fuel.2020.117656. gasification in a bubbling fluidized bed reactor. Energy Convers Manag
[55] Stark AK, Ghoniem AF. Quantification of the influence of particle diameter on 2015;94:345–52. https://doi.org/10.1016/j.enconman.2015.01.060.
Polycyclic Aromatic Hydrocarbon (PAH) formation in fluidized bed biomass pyr- [64] Pio DT, Tarelho LAC, Matos MAA. Characteristics of the gas produced during bio-
olysis. Fuel 2017;206:276–88. https://doi.org/10.1016/j.fuel.2017.06.020. mass direct gasification in an autothermal pilot-scale bubbling fluidized bed re-
[56] Han J, Kim H. The reduction and control technology of tar during biomass actor. Energy 2017;120:915–28. https://doi.org/10.1016/j.energy.2016.11.145.

15

You might also like