1-s2.0-s0960852410013726-main.pdf

7
7/21/2019 1-s2.0-S0960852410013726-main.pdf http://slidepdf.com/reader/full/1-s20-s0960852410013726-mainpdf 1/7 Generation of hydrogen rich gas through fluidized bed gasification of biomass M.K. Karmakar * , A.B. Datta Thermal Engineering Group, Central Mechanical Engineering Research Institute (CSIR), Durgapur 713209, India a r t i c l e i n f o  Article history: Received 30 April 2010 Received in revised form 2 August 2010 Accepted 4 August 2010 Available online 6 August 2010 Keywords: Fluidized bed Biomass Steam gasification Equilibrium modeling Hydrogen a b s t r a c t The objective of this study was to investigate the process of generating hydrogen rich syngas through thermo chemical fluidized bed gasification of biomass. The experiments were performed in a laboratory scale externally heated biomass gasifier. Rice husk had been taken as a representative biomass and, steam had been used as the fluidizing and gasifying media. A thermodynamic equilibrium model was used to predict the gasification process. The work included the parametric study of process parameters such as reactor temperature and steam biomass ratio which generally influence the percentage of hydrogen con- tent in the product gas. Steam had been used here to generate nitrogen free product gas and also to increase the hydrogen concentration in syngas with a medium range heating value of around 12 MJ/Nm 3 .  2010 Elsevier Ltd. All rights reserved. 1. Introduction Energy is an essential factor in both livelihoods and industrial activities and, hydrogen has been identified as a clean energy source and a potential alternative fuel. When hydrogen is com- busted, it does not add any adverse effect to the environment. However, to make the sustainable energy security based on hydro- gen, renewable resource such as biomass instead of fossil fuels has to be utilized. Biomass gasification produces fuel gas or synthesis gas through the thermo chemical conversion of biomass, usually involving par- tial oxidation of feedstock in a reducing atmosphere in presence of air, oxygen and/or steam (Li et al., 2004). The biomass materials differ greatly in chemical, physical and morphological properties and, therefore, this necessitates developing different methods of gasification and consequently requires different reactor designs or even gasification technologies. There is an excellent overview concerning the characteristics of the various gasifiers, i.e. of up- draft and downdraft fixed bed gasifiers and bubbling, circulating fluidized and entrained beds (Bridgwater, 1995). Moving-bed gas- ifiers suffer from high tar yields in the product gas (Beenackers, 1999). The inability to maintain uniform radial temperature pro- files and to avoid local slagging problems makes the moving bed unsuitable for large installations (Babu, 1995). Fluidized beds now find wide application in biomass gasification ( Corella et al., 1998). Air-blown processes produce low calorific value gases with a typical Higher Heating Value (HHV) of 4–7 MJ/Nm 3 , while oxygen and steam blown processes result in gases with a HHV of 10– 18 MJ/Nm 3 (Schuster et al., 2001). Biomass gasification models are predominantly separated in two groups: equilibrium approach and kinetic approach (Li et al., 2001). Kinetic models take into account the chemical kinetics of the main reactions and the transfer phenomena among phases, estimating the composition of each species on any point of space and time of a system. These models are specific in general for each process, providing important considerations related to chemical mechanisms and ways to increase reaction rates and process per- formance. On the other hand, equilibrium models provide the greatest possible conversion of each species regardless the system size and the time needed to reach equilibrium. These models do not require details of system geometry neither estimate the neces- sary time to reach that equilibrium ( Rodrigues et al., 2009). Bilodeau et al. (1993) developed a mathematical model of bio- mass gasification in a fluidized bed reactor considering axial vari- ations of concentrations and temperature in bubble and emulsion phases. The mass balance involved instantaneous oxidation and equilibrium devolatilization of biomass, kinetics of solid–gas gasi- fication reactions as well as gaseous phase reactions with inter- phase mass transfer and gas convection. Constantineau et al. (2007) proposed to predict the performance of fluidized bed reac- tors operated wholly or partially in the slug flow regime. This mod- el simulated the transition from bubbling to slugging within a given reactor with operating variables such as superficial gas velocity, bed inventory and change of height. Corella and Sanz (2005) developed a one-dimensional model for an atmospheric circulating fluidized bed biomass gasifier under stationary state. 0960-8524/$ - see front matter   2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.biortech.2010.08.015 * Corresponding author. Tel.: +91 343 6452157; fax: +91 343 2546745. E-mail addresses:  [email protected],  [email protected] (M.K. Karmakar). Bioresource Technology 102 (2011) 1907–1913 Contents lists available at ScienceDirect Bioresource Technology journal homepage: www.elsevier.com/locate/biortech

Upload: enrique-franco-ariza

Post on 05-Mar-2016

212 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: 1-s2.0-S0960852410013726-main.pdf

7/21/2019 1-s2.0-S0960852410013726-main.pdf

http://slidepdf.com/reader/full/1-s20-s0960852410013726-mainpdf 1/7

Generation of hydrogen rich gas through fluidized bed gasification of biomass

M.K. Karmakar *, A.B. Datta

Thermal Engineering Group, Central Mechanical Engineering Research Institute (CSIR), Durgapur 713209, India

a r t i c l e i n f o

 Article history:

Received 30 April 2010

Received in revised form 2 August 2010Accepted 4 August 2010

Available online 6 August 2010

Keywords:

Fluidized bed

Biomass

Steam gasification

Equilibrium modeling

Hydrogen

a b s t r a c t

The objective of this study was to investigate the process of generating hydrogen rich syngas through

thermo chemical fluidized bed gasification of biomass. The experiments were performed in a laboratory

scale externally heated biomass gasifier. Rice husk had been taken as a representative biomass and, steamhad been used as the fluidizing and gasifying media. A thermodynamic equilibrium model was used to

predict the gasification process. The work included the parametric study of process parameters such as

reactor temperature and steam biomass ratio which generally influence the percentage of hydrogen con-

tent in the product gas. Steam had been used here to generate nitrogen free product gas and also to

increase the hydrogen concentration in syngas with a medium range heating value of around 12 MJ/Nm 3.

  2010 Elsevier Ltd. All rights reserved.

1. Introduction

Energy is an essential factor in both livelihoods and industrialactivities and, hydrogen has been identified as a clean energy

source and a potential alternative fuel. When hydrogen is com-

busted, it does not add any adverse effect to the environment.

However, to make the sustainable energy security based on hydro-

gen, renewable resource such as biomass instead of fossil fuels has

to be utilized.

Biomass gasification produces fuel gas or synthesis gas through

the thermo chemical conversion of biomass, usually involving par-

tial oxidation of feedstock in a reducing atmosphere in presence of 

air, oxygen and/or steam (Li et al., 2004). The biomass materials

differ greatly in chemical, physical and morphological properties

and, therefore, this necessitates developing different methods of 

gasification and consequently requires different reactor designs

or even gasification technologies. There is an excellent overview

concerning the characteristics of the various gasifiers, i.e. of up-

draft and downdraft fixed bed gasifiers and bubbling, circulating

fluidized and entrained beds (Bridgwater, 1995). Moving-bed gas-

ifiers suffer from high tar yields in the product gas (Beenackers,

1999). The inability to maintain uniform radial temperature pro-

files and to avoid local slagging problems makes the moving bed

unsuitable for large installations (Babu, 1995). Fluidized beds

now find wide application in biomass gasification (Corella et al.,

1998). Air-blown processes produce low calorific value gases with

a typical Higher Heating Value (HHV) of 4–7 MJ/Nm3, while oxygen

and steam blown processes result in gases with a HHV of 10–

18 MJ/Nm

3

(Schuster et al., 2001).Biomass gasification models are predominantly separated in

two groups: equilibrium approach and kinetic approach (Li et al.,

2001). Kinetic models take into account the chemical kinetics of 

the main reactions and the transfer phenomena among phases,

estimating the composition of each species on any point of space

and time of a system. These models are specific in general for each

process, providing important considerations related to chemical

mechanisms and ways to increase reaction rates and process per-

formance. On the other hand, equilibrium models provide the

greatest possible conversion of each species regardless the system

size and the time needed to reach equilibrium. These models do

not require details of system geometry neither estimate the neces-

sary time to reach that equilibrium (Rodrigues et al., 2009).

Bilodeau et al. (1993) developed a mathematical model of bio-

mass gasification in a fluidized bed reactor considering axial vari-

ations of concentrations and temperature in bubble and emulsion

phases. The mass balance involved instantaneous oxidation and

equilibrium devolatilization of biomass, kinetics of solid–gas gasi-

fication reactions as well as gaseous phase reactions with inter-

phase mass transfer and gas convection.   Constantineau et al.

(2007) proposed to predict the performance of fluidized bed reac-

tors operated wholly or partially in the slug flow regime. This mod-

el simulated the transition from bubbling to slugging within a

given reactor with operating variables such as superficial gas

velocity, bed inventory and change of height.   Corella and Sanz

(2005)   developed a one-dimensional model for an atmospheric

circulating fluidized bed biomass gasifier under stationary state.

0960-8524/$ - see front matter     2010 Elsevier Ltd. All rights reserved.doi:10.1016/j.biortech.2010.08.015

*   Corresponding author. Tel.: +91 343 6452157; fax: +91 343 2546745.

E-mail addresses:   [email protected],   [email protected]

(M.K. Karmakar).

Bioresource Technology 102 (2011) 1907–1913

Contents lists available at   ScienceDirect

Bioresource Technology

j o u r n a l h o m e p a g e :   w w w . e l s e v i e r . c o m / l o c a t e / b i o r t e c h

Page 2: 1-s2.0-S0960852410013726-main.pdf

7/21/2019 1-s2.0-S0960852410013726-main.pdf

http://slidepdf.com/reader/full/1-s20-s0960852410013726-mainpdf 2/7

The model was based on kinetic equations for the reaction network

solved together with mass and heat balances and, with several

hydrodynamic considerations. The reaction network used in the

model involved twelve different reactions.   Schuster et al. (2001)

also developed a model for steam gasification of biomass applying

thermodynamic equilibrium calculations. With this model, the

simulation of a decentralized combined heat and power station

based on a dual fluidized bed steam gasifier was carried out. Fuelcomposition, temperature and amount of gasification agent were

varied over a wide range and it was shown that the accuracy of 

an equilibrium model for the gas composition was sufficient for

thermodynamic considerations.

Kinetic models predict the progress and product composition at

different positions along a reactor, where as equilibrium model

predicts the maximum achievable yield of a desired product from

a reacting system (Li et al., 2004). Kinetic models always contain

parameters which make them hardly applicable to different plants

(Schuster et al., 2001). An accurate description of the chemical ki-

netic rate expression is a key issue. The choice of chemical kinetic

laws is difficult because there are as many kinetic laws as kinetic

studies. A large discrepancy can be observed between them and

it is highly hazardous to extrapolate literature results obtained un-

der different operating conditions (reactor, heating rate, tempera-

ture, biomass type) (Dupont et al., 2007). Although kinetic

models provide essential information on mechanisms and rates,

equilibrium models are valuable because they can predict thermo-

dynamic limits as a guide to process design, evaluation and

improvement. Equilibrium model also provides a useful design

aid in evaluating the limiting possible behavior of a complex react-

ing system which is difficult or unsafe to reproduce experimentally

or in commercial operation (Li et al., 2004).

The present work aimed to develop a fluidized bed biomass gas-

ifier using steam as the gasifying agent. A laboratory scale fluidized

bed gasifier had been developed for this purpose. The effect of 

steam biomass ratio and gasifying temperature on the product

gas composition had been carried out. A thermodynamic equilib-

rium model had been used to predict the gas composition withthe assumptions that the principal reactions were at thermody-

namic equilibrium conditions. The model considered three mass

balance equations of elements C, H, O and two equilibrium con-

stants relations, which were solved for predicting the gas

compositions.

2. Methods

 2.1. Materials

In this study, rice husk was taken as the raw material since it is

one of the important biomass resources in India. The proximate

and ultimate analyses of rice husk are presented in  Table 1. The

steam, the gasification agent, had been used as the fluidizing mediaas it prevented the dilution of product gases with nitrogen and also

increased the hydrogen content in the product gas.

The superheated steam was supplied to the gasification cham-

ber at the bottom in order to enable better fluidization and thereby

attaining better gasification.

As per Geldart’s classification of particle distribution, the silica

sands are classified as the group B particles (Geldart, 1973). These

silica sand particles were used as bed materials in the present

investigation. The sand samples were prepared using various mesh

sieves and the mean diameter was found as 0.334 mm. The sandparticles ensured proper fluidization in the gasifier bed and main-

tained the uniform temperature throughout the gasification zone.

During the experiment, the gasifier temperature was varied in

the range of 650–800  C. The experiments showed that rice husk

was difficult to be fluidized due to its non-granular and flaky nat-

ure. However, this fluidization behavior of rice husk was improved

due to presence of silica sand as bed materials.

 2.2. Experimental set up

A laboratory scale fluidized bed biomass gasifier had been set

up in the laboratory shed. The schematic diagram is shown in

Fig. 1. The gasifier vessel was made of stainless steel with inside

diameter of 50 mm and a height of 1200 mm. The distributor plate

was fitted with at the bottom and the product gas outlet pipe came

out from the top of gasifier. This gasifier was placed inside an elec-

tric furnace, which provided the heat for gasification reaction and

the temperature inside the gasifier was controlled using a thermo-

couple with a control panel system. There was an under-bed feed-

ing arrangement of biomass fuels into the reactor and a lock

hopper had been kept to dispose off the ash from the gasifier

periodically.

The under-bed feeding system of biomass was comprised of two

screw feeders fitted with a hopper. The upper feeder connected to a

variable speed drive system that controlled the fuel feed rate and

the lower high-speed screw feeder was used to feed the fuel inside

the gasifier vessel instantaneously. The lower screw feeder also

prevented the pyrolysis of biomass outside the gasifier.

Steam for gasification, obtained from a small boiler, was furthersuperheated in an electric furnace to the temperature range of 

200–250  C. The quantity of steam is very important for maintain-

ing fluidization conditions in the biomass gasifiers.

The gasifier had been operated at different process conditions to

evaluate the performance. The product gas was generated at vary-

ing reactor temperatures between 650 and 770  C at a fixed steam

biomass ratio of 1.32. In another set of tests, the steam biomass ra-

tios were varied in the range of 0.6–1.7 to produce the syngas

maintaining the gasifier temperature at 750  C.

The product gas from the gasifier was made dust-free and

cleaned before the sampling was done for analysis in a Gas Chro-

matograph. The hot product gas from the gasifier was passed

through a cyclone to remove the larger particles. After the cyclone,

the gas still contained dust particles and tar and hence, it was fur-ther de-dusted by passing it through a bag filter. A water cooler

and an ice trap were used in series for cooling of fuel gas and tar

capture. Finally, the product gas was passed through a silica gel

unit to remove the remaining moisture in the gas before taking

up the sampling process.

The sampling system was composed of gas sampling probes fit-

ted with septum. The stainless steel gas sampling probes were

12.5 mm in diameter and 50 mm in length and they were located

at the downstream of suction blower. A syringe of volume of 

10 ml was used to collect the gas sample through the gas sampling

probe.

The analysis was done in a Gas Chromatograph (Make–Chemito,

model – GC 1000). Two detectors had been used – one was Ther-

mal Conductivity Detector (TCD) and another was Flame IonizationDetector (FID). The standard gas mixtures had been used for cali-

 Table 1

Ultimate and proximate analysis of rice husk.

Ultimate analysis Proximate analysis

Components Percent Components Percent

Carbon 38.43 Volatile matter 55.54

Hydrogen 2.97 Fixed carbon 14.99

Sulfur 0.07 Moisture 9.95

Nitrogen 0.49 Ash 19.52

Oxygen 36.36

Ash 21.68

HHV = 15.68 MJ/kg

1908   M.K. Karmakar, A.B. Datta/ Bioresource Technology 102 (2011) 1907–1913

Page 3: 1-s2.0-S0960852410013726-main.pdf

7/21/2019 1-s2.0-S0960852410013726-main.pdf

http://slidepdf.com/reader/full/1-s20-s0960852410013726-mainpdf 3/7

brating the equipment. The GC showed the different picks for dif-

ferent constituents of product gas mixture.

An orifice meter was positioned on the pipe between the box

containing silica gel and the suction blower to measure the gas

yield. The pressure drop across this orifice plate was measured

using a micro manometer and this pressure drop was then used

to estimate the gas yield rate.

 2.3. Mathematical modeling 

At chemical equilibrium, a reacting system achieves its moststable composition when the entropy of the system is maximized

and its Gibbs free energy is minimized. Two approaches are de-

scribed for equilibrium modeling: stoichiometric and non-stoichi-

ometric (Smith and Missen, 1982). The stoichiometric approach

requires a clearly defined reaction mechanism incorporating all

chemical reactions and species involved. In a non-stoichiometric

formulation, on the other hand, no particular reaction mechanism

or species are involved in the numerical solution. The only input

required to specify is the feed elemental composition, which can

be readily obtained from ultimate analysis data. This method is

particularly suitable for problems with unclear reaction mecha-

nisms and feed streams like biomass whose precise chemical com-

positions are unknown.

The laboratory scale gasifier described above was operated un-der near-steady-state conditions. In the current equilibrium model

the reactor was implicitly considered to be zero-dimensional, i.e.

neither any spatial distribution of parameters is considered nor

were there any changes with time, because all forward and reverse

reactions had reached the chemical equilibrium state. Thus, the

molar inflow for any individual element involved in the chemical

reactions was written as the sum of moles of that element in the

various feed streams. In this study, tars had not been included in

the product stream.

To develop the model, the chemical formula of feedstock was

defined as CH xOY . The global gasification reaction could be written

as follows:

CH X OY  þ wH2O þ mH2O ¼  x1H2 þ x2CO þ x3CO2 þ x4H2O þ x5CH4

ð1Þ

where x  and  y  were the number of atoms of hydrogen and oxygen

for each atom of carbon in the feedstock, respectively,  w  was the

amount of moisture per kmol of feedstock and  m  was the amount

of steam supplied per kmol of feedstock. On the right-hand side,

 xi’s were the molar concentrations of species i that were also un-

known. The global steam gasification reaction in Eq. (1) represented

an overall reaction, but a number of competing intermediate reac-

tions took place during the process, of which the following were

predominant.

Steam gasification

C þ H2O ¼  CO þ H2   ðþ131:4 kJ=molÞ ð2Þ

Boudouard reaction

C þ CO2 ¼  2CO   ðþ172:6 kJ=molÞ ð3Þ

Methanation reaction

C þ 2H2 ¼  CH4   ð74:9 kJ=molÞ ð4Þ

Steam reforming reaction

CH4 þ H2O ¼  CO þ 3H2   ðþ206:2 kJ=molÞ ð5Þ

Water gas shift reaction

CO þ H2O ¼  CO2 þ H2   ð41:2 kJ=molÞ ð6Þ

According to Von Fredersdorff and Elliot (1963), the three reac-

tions namely, Boudouard, steam gasification and methanation

were in equilibrium and the water gas shift reaction was a combi-

nation of the Boudouard and steam gasification reactions.

To find the five unknown species i.e. x1, x2, x3, x4 and x5  of the

producer gas, five equations were required. Those equations were

generated using mass balance equations and equilibrium constants

relationships. Considering the global gasification reaction in Eq.

(1), the first three equations were formulated by balancing each

chemical element.

Carbon balance:

 x2 þ x3 þ x5  1 ¼  0   ð7Þ

Hydrogen balance:

2 x1 þ 2 x4 þ 4 x5  x 2w 2m ¼  0   ð8Þ

Fig. 1.   Schematic diagram of bubbling fluidized bed gasifier system.

M.K. Karmakar, A.B. Datta / Bioresource Technology 102 (2011) 1907–1913   1909

Page 4: 1-s2.0-S0960852410013726-main.pdf

7/21/2019 1-s2.0-S0960852410013726-main.pdf

http://slidepdf.com/reader/full/1-s20-s0960852410013726-mainpdf 4/7

Oxygen balance:

 x2 þ 2 x3 þ x4  w m Y  ¼  0:   ð9Þ

Ultimate analysis of rice husk (dry and ash free basis) yielded a

typical mass composition of 49% carbon, 46.4% oxygen, and 3.8%

hydrogen with the balance comprised of traces of nitrogen and sul-

fur. Considering the major elements, the fuel was represented on a

molar basis as CH0.93

O0.71

.

Chemical equilibrium is usually explained either by minimiza-

tion of Gibbs free energy or by using an equilibrium constant. To

minimize the Gibbs free energy, constrained optimization methods

are generally used which requires an understanding of complex

mathematical theories. The present thermodynamic equilibrium

model, a more simplistic one, was developed based on the equilib-

rium constant method and not on the Gibbs free energy principle.

In this study, the other two equations were obtained from the

equilibrium constants of the reactions described in Eqs. (4) and (6).

For the present model study, the thermodynamic equilibrium

was assumed for all chemical reactions in the gasification zone.

All gases were assumed to be ideal and all reactions form at 1

atmospheric pressure. Therefore, the equilibrium constants, which

were the functions of temperature for the methanation reaction

and the water gas shift reaction, are given below:The equilibrium constant  k1 for methanation reaction

k1 ¼  x5

ð x1Þ2

  ð10Þ

The equilibrium constant  k2 for water–gas shift reaction

k2 ¼ x3   x1

 x2   x4

ð11Þ

Heat balancing of the reactants and products of the global reac-

tion in Eq. (1) results in the equation shown below:

H 0f RH  þ wðH 0fH2O þ H vapÞ þ mH vap þ Q IN

¼ x1H 0fH2

þ x2H 0fCO þ x3H 

0fCO2

þ þ x4H 0fH 2O þ x5H 

0fCH4

þ DT ð x1C pH2 þ x2C pCO þ x3CpCO2

 þ  x4CpH2O þ x5CpCH4Þ ð12Þ

where   C p   is the specific heat of gas species. The heating value of 

biomass,   H 0fRH, was determined experimentally. The heat balance

Eq.  (12)  contained a term   Q IN  which stands for the external heat

addition required for endothermic reactions to occur. The heat of 

formation,  DH , of various gases could be sourced from the JANAF

Thermochemical Tables (Stull and Prophet, 1971). The dependence

of specific heat on temperature was given by various empirical

equations and according to Perry and Green (1997), the most sim-

plified version of specific heat, Cpam, at arithmetic mean tempera-

ture is expressed as

Cpam ¼  R½ A þ BT am þ C ð4T 2am  T 1T 2Þ=3 þ D=ðT 1T 2Þ ð13Þ

where T am = (T 1 +  T 2)/2, T 1 and  T 2 are the reference temperature andthe reactor temperature, respectively. The values of the heat capac-

ity constants A, B, C  and D are given by Perry and Green (1997) and R

is the universal gas constant in J/mol-K.

Following Zainal et al. (2001), the equilibrium constant k can be

written as

RTlnk ¼  DG0 ð14Þ

where  DG0 is the standard Gibbs function of formation.

dlnk

dT   ¼

 DH 0

RT 2  ð15Þ

Integrating Eq. (15) the heat of formation can be calculated as

lnk ¼Z   DH 0

RT 2  dT  þ I    ð16Þ

where I  is a constant of integration.

DH 0 were also calculated as given by Perry and Green (1997)

DH 0

R  ¼

  J 

Rþ D A:T  þ

DB:T 2

2  þ

DC :T 3

DD

T   ð17Þ

where D A,  DB,  DC ,  DD are the coefficients for determining specific

heats of gases and  J  is a constant.

Substituting Eq. (17) into Eq. (16), one gets

ln k ¼   J 

RT  þ D A: ln T  þ

DB:T 

2  þ

DC :T 2

6  þ

 DD

2T 2 þ I    ð18Þ

The dependence of  DG0 on temperature was analyzed as

DG0 ¼ J   RT ½D A: ln T  þ

DB:T 

2  þ

DC :T 2

6  þ

 DD

2T 2 þ I  ð19Þ

The values of standard Gibbs function of formation for various

gas compositions were obtained from Perry and Green (1997).

Both   J   and   I   were calculated, respectively, from Eqs. (17) and

(19) at temperature 298 K. Two equilibrium equations were re-

quired to determine the equilibrium constants  k1  and  k2. Equilib-

rium constant k1  for the reaction in Eq. (4) was solved as follows:

D ¼ CH4  C   2H2   ð20Þ

D A, DB, DC  and DD were obtained from the data on heat capac-

ity. The equations to determine the values of  D A,  DB,  DC  and  DD

are expressed as

D A ¼  ACH4   AC  2 AH2

  ð21Þ

DB ¼  BCH4  BC  2BH2

  ð22Þ

DC  ¼  C CH4  C C  2C H2

  ð23Þ

DD ¼  DCH4  DC  2DH2

  ð24Þ

Application of these equations revealed the equilibrium con-

stant k1.

Similarly, the equilibrium constant   k2   for the water gas shift

reaction in Eq. (6) could be solved by adopting the same procedure.

The molar concentrations  x1, x2, x3, x4  and  x5  were obtained by

solving Eqs. (7)–(11) simultaneously using Newton Raphson

method.

3. Results and discussion

The experimental tests were carried out to determine the effect

of steam biomass ratio and reactor temperature on gas composi-

tion and yield. The experimental values were compared with the

equilibrium model predictions.

 3.1. Model validation

The equilibrium model, described in Section 2.3, was needed for

validation by comparing the model results and the experimental

data. In the following, the present experimental data and data re-

ported by Rapanga et al. (2000)  were compared with the calcula-

tions from the currently developed model. The comparison is

shown in Table 2a. The error had been estimated using the statis-

tical parameter of root mean square (RMS) error, where

RMS ¼

 ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiPð X e  X  pÞ2

where, X e and  X p were the experimental data and the predicted va-

lue of product gas species from present model, respectively, and N

was the number of observations. The RMS error for the presentexperimental runs had been presented in  Table 2b. It was observed

1910   M.K. Karmakar, A.B. Datta/ Bioresource Technology 102 (2011) 1907–1913

Page 5: 1-s2.0-S0960852410013726-main.pdf

7/21/2019 1-s2.0-S0960852410013726-main.pdf

http://slidepdf.com/reader/full/1-s20-s0960852410013726-mainpdf 5/7

from these tables that the predicted results generally agreed well

with the experimental data.

 3.2. Effect of reactor temperature

Fig. 2 shows both theoretical and experimental composition of 

the product gas at different reactor temperatures. It was observed

that H2  and CO concentrations increased with rise in temperature

and the concentrations of CH4 and CO2  showed an opposite trend.

According to Le Chatelier’s principle, higher temperatures favour

the reactants in exothermic reactions and the products in endo-

thermic reactions. Therefore, the endothermic reaction in Eq.   (5)was strengthened with increasing temperature, which resulted in

an increase of H2  concentration and a decrease of CH4   concentra-

tion. It could be seen that a good agreement existed between the

predicted values and the experimental data, indicating that the

equilibrium model predicted reasonably well about the gas compo-

sitions for the fluidized bed gasifier.

 3.3. Effect of steam biomass ratio

The product gas composition varying with steam biomass ratio

is presented in Fig. 3. It is seen from the graphs that, with the in-

crease of steam biomass ratio, the experimental values of H2  and

CO2 concentrations showed a gradual increased trend, while those

of CO and CH4 decreased. The predicted results of H2  agreed fairly

well with the experimental data, but those of CO2 and CO had some

differences with the experimental values although the changes

showed similar trends. Also, the model values of CH 4   concentra-

tion, which increased with rise in steam flow rate, contradicted

the experimental trend. The possible explanation to this observa-

tion was that the equilibrium state might not have reached for

not having enough bed temperature in gasifier. The devolatiliza-

tion of biomass might have contributed higher content of methane

in product gas which did not react completely with the concentra-

tions of CO, CO2 and H2. Furthermore, the tar, actually produced in

the gasification process, was not considered in the present equilib-

rium model. Some of these facts might have caused the model val-

ues for CO and CO2   concentrations to differ from experimental

data. The gasification process involving thermal pyrolysis, homo-

geneous and heterogeneous reactions were complicated to deter-

mine the kinetics of the chemical reactions. A simple but

adequate kinetically-modified equilibrium model might have re-

sulted in better prediction of the gas composition.

The experimental data of bubbling fluidized bed gasifier againstvarious steam biomass (S /B) ratios at fixed temperature of 750  C

are shown in  Table 3. Steam being the major source of hydrogen

during steam gasification, an increase in steam biomass ratio re-

sulted in higher production of H2. Due to higher production of 

hydrogen, the product gas volume was also increased as indicated

 Table 2a

Comparison of present model result with the experimental data.

Gas composition % mol dry basis Temperature = 750 C,

steam biomass ratio = 1.0

Temperature = 770  C,

steam biomass ratio = 1.0

Present Model Present exp. data RMS error Present model Literature exp. data (Rapanga et al., 2000) RMS error

H2   50.76 48.88 2.66 55.0 52.2 2.517

CO 19.78 22.7 22.3 23.0

CO2   25.52 22.2 18.5 16.9

CH4   3.96 6.22 4.2 7.9

 Table 2b

RMS error for different experimental runs of the present work.

Expt. run No. Operating parameters RMS error

Gasifier temperature (C) Steam biomass ratio

1. 650 Const. at 1.32 1.44

2. 690 1.80

3. 730 2.04

4. 770 1.91

5. Const at 750 0.6 1.90

6. 1.0 2.65

7. 1.32 3.27

8. 1.70 3.62

640 660 680 700 720 740 760 780

0

10

20

30

40

50

60

   G  a  s  c  o  m  p  o  s   i   t   i  o  n   (   %    b

  y  m

  o   l  e   )

Temperature, deg C

 H2-Model value

 H2-Exp value

 CO-Model value

 CO-Exp value

 CO2-Model value

 CO2-Exp value

 CH4-Model value

 CH4-Exp value

Fig. 2.   Gas composition vs. gasification temperature with steam biomass ratio of 1.32.

M.K. Karmakar, A.B. Datta / Bioresource Technology 102 (2011) 1907–1913   1911

Page 6: 1-s2.0-S0960852410013726-main.pdf

7/21/2019 1-s2.0-S0960852410013726-main.pdf

http://slidepdf.com/reader/full/1-s20-s0960852410013726-mainpdf 6/7

in Table 3. The proportion of CO in the product gas became low

when it was produced by steam reforming of CH4. With the in-

crease in steam biomass ratio, the steam reforming of methane be-

came significant and as a result, the proportion of CO was

decreased. The higher heating value (HHV) of product gas de-

creased with rise in steam biomass ratio from 12.21 MJ/Nm3 for

S /B  ratio of 0.6–11.18 MJ/Nm3 for S/B ratio of 1.7 at temperature

of 750  C.

 3.4. Carbon conversion and gasification efficiency

The carbon conversion is defined as the fraction of carbon in thefeed converted to gaseous products. The results of carbon conver-

sion and gasification efficiency are presented in   Table 3. Higher

temperature and higher steam biomass ratio favoured more carbon

conversion. For all the runs in the present study, the overall carbon

conversion was within 84–90% closure. Analysis of tar and the frac-

tion of unconverted solid carbon had not been included in the

study and, these components might account for the rest.

Cold gas efficiency (CGE) was used to evaluate the gasification

performance. The cold gas efficiency is defined as the percentage

of the heating value of biomass converted into the heating value

of the product gas. The system ran on steam only and no oxygen

was supplied. Thus the endothermic heat was required to be sup-

plied externally for gasification reactions to happen. This heat was

considered to be added to the denominator to calculate gasificationefficiency.

CGE ¼  Gas yield ðNm3 kg

1fuelÞ HHV of productgas  ðMJNm3 Þ

HHV of fuel  ðMJkg1

Þ þ Heat addition for endothermic reaction  ðMJkg1

Þ

Calculated gasification efficiencies for different experimental

runs are given in   Table 3. Experimental cold gas efficiencies for

the present runs were in the range of 63–66%. The cold gas effi-

ciency, taking account of the external heat supply, measured the

gasification performance.

4. Conclusion

During the study of steam gasification of rice husk in a fluidizedbed gasifier, it was seen that hydrogen content in product gas in-

creased with rise in temperature and increased in steam biomass

ratio. Hydrogen content in product gas had reached as high as

53.08%. Carbon monoxide content increased with rise in tempera-

ture, but decreased with steam biomass ratio. The percentage of 

methane decreased with rise in temperature and steam biomass

ratio. Increase in steam biomass ratio generally gave rise to higher

gas yields. Higher temperature also yielded the higher quantity of 

gas. The proposed mathematical model closely predicted the

experimental results.

 Acknowledgements

The authors thankfully acknowledge Prof. Gautam Biswas,Director, Central Mechanical Engineering Research Institute

0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.80

10

20

30

40

50

60

   G  a  s  c  o  m  p  o  s   i   t   i  o  n   (   %

    b  y  m  o   l  e   )

Steam Biomass ratio

 H2-Model value

 H2-Exp value

 CO-Model value

 CO-Exp value

 CO2-Model value

 CO2-Exp value CH4-Model value

 CH4-Exp value

Fig. 3.   Gas composition vs. steam biomass ratio at reactor temperature of 750  C.

 Table 3

Carbon balance and gasification efficiency of experimental results of fluidized bed gasifier.

Exp. run no. 1 2 3 4 5 6 7 8

Steam biomass ratio Constant at 1.32 0.60 1.00 1.32 1.70

Gasifier temp. (C) 650 690 730 770 Constant at 750

Gas Composition (% of mole)

H2   47.25 50.5 52.2 53.08 47.81 48.88 51.17 51.89

CO 11.25 12.83 15.9 17.85 27.48 22.70 19.65 17.38

CO2   31.9 28.51 25.65 23.9 18.09 22.20 23.15 24.81

CH4   9.6 8.16 6.25 5.17 6.62 6.22 6.03 5.92

HHV of prod. gas (MJ/Nm3) 11.28 11.23 11.16 11.09 12.21 11.59 11.42 11.18

Gas yield (Nm3/kg of raw biomass) 1.03 1.12 1.16 1.21 1.05 1.09 1.16 1.21

Carbon conversion (%) 84.10 85.82 85.83 87.88 84.83 86.25 87.68 90.11

Cold gas efficiency (%) 62.99 64.78 65.58 66.06 65.75 65.96 66.10 66.15

1912   M.K. Karmakar, A.B. Datta/ Bioresource Technology 102 (2011) 1907–1913

Page 7: 1-s2.0-S0960852410013726-main.pdf

7/21/2019 1-s2.0-S0960852410013726-main.pdf

http://slidepdf.com/reader/full/1-s20-s0960852410013726-mainpdf 7/7

(CMERI), Durgapur, India and Dr. P.K. Chatterjee, Scientist G and

Head, Thermal Engineering Group, CMERI for their continuous sup-

port, enthusiasm and encouragement.

References

Babu, S.P., 1995. Thermal gasification of biomass technology development: end of task report for 1992 to 1994. Biomass and Bioenergy 9, 5–15.

Beenackers, A.M., 1999. Biomass gasification in moving beds, a review of Europeantechnologies. Renewable Energy 16, 1180–1186.

Bilodeau, J.F., Thérien, N., Proulx, P., Czernik, S., Chornet, E., 1993. A mathematicalmodel of fluidized bed biomass gasification. The Canadian Journal of ChemicalEngineering 71 (4), 549–557.

Bridgwater, A.V., 1995. The technical and economic feasibility of biomassgasification for power generation. Fuel 74 (5), 631–653.

Constantineau, J.P., Grace, J.R., Lim, C.J., Richards, G.G., 2007. Generalized bubbling–slugging fluidized bed reactor model. Chemical Engineering Science 62, 70–81.

Corella, J., Orio, A., Aznar, P., 1998. Biomass gasification with air in fluidized bed:reforming of the gas composition with commercial steam reforming catalysts.Industrial and Engineering Chemistry Research 37, 4617–4624.

Corella, J., Sanz, A., 2005. Modeling circulating fluidized bed biomass gasifiers, apseudo-rigorous model for stationary state. Fuel Processing Technology 86,1021–1053.

Dupont, C., Boissonnet, G., Seiler, J.M., Gauthier, P., Schweich, D., 2007. Study aboutthe kinetic processes of biomass steam gasification. Fuel 86, 32–40.

Geldart, D., 1973. Types of gas fluidization. Powder Technology 7, 285–292.Li, X., Grace, J.R., Watkinson, A.P., Lim, C.J., Ergudenler, A., 2001. Equilibrium

modeling of gasification: a free energy minimization approach and itsapplication to a circulating fluidized bed coal gasifier. Fuel 80, 195–207.

Li, X.T., Grace, J.R., Lim, C.J., Watkinson, A.P., Chen, H.P., Kim, J.R., 2004. Biomassgasification in a circulating fluidized bed. Biomass and Bioenergy 26, 171–193.

Perry, R.H., Green, D.W., 1997. Perry’s Chemical Engineers’ Handbook, seventh ed.McGraw-Hill, New York.

Rapanga, S., Jand, N., Kiennemann, A., Foscolo, P.U., 2000. Steam-gasification of biomass in a fluidized-bed of olivine particles. Biomass and Bioenergy 19, 87–

197.Rodrigues, R., Secchi, A.R., Marcílio, N.R., Godinho, M., 2009. Modeling of biomass

gasification applied to a combined gasifier-combustor unit: equilibrium andkinetic approaches. 10th international symposium on process systemsengineering. Computer Aided Chemical Engineering 27, 657–662.

Schuster, G., Loffler, G., Weigl, K., Hofbauer, H., 2001. Biomass steam gasification –an extensive parametric modeling study. Bioresource Technology 77, 71–79.

Smith, W.R., Missen, R.W., 1982. Chemical reaction equilibrium analysis: theory andalgorithms. Wiley Publisher, New York.

Stull, D.R., Prophet, H., 1971. JANAF Thermochemical Tables. Dow PublishingCompany, Midland, USA.

Von Fredersdorff, C.G., Elliot, M.A., 1963. Chemistry of Coal Utilization,Supplementary Volume. Wiley Publisher, New York.

Zainal, Z.A., Ali, R., Lean, C.H., Seetharamu, K.N., 2001. Prediction of performance of adowndraft gasifier using equilibrium modeling of different biomass materials.Energy Conversion and Management 42, 1499–1515.

M.K. Karmakar, A.B. Datta / Bioresource Technology 102 (2011) 1907–1913   1913