Sunteți pe pagina 1din 8

Available online at www.sciencedirect.

com

ScienceDirect
Procedia Technology 25 (2016) 982 – 989

Global Colloquium in Recent Advancement and Effectual Researches in Engineering, Science and
Technology (RAEREST 2016)

Stoichiometric Equilibrium Model based Assessment of Hydrogen


Generation through Biomass Gasification
Joel Georgea*, P. Aruna, C. Muraleedharana
a
Department of Mechanical Engineering, National Institute of Technology Calicut, Kozhikode-673601, India

Abstract

Hydrogen produced from renewable energy sources is clean and sustainable. Biomass gasification has a significant
role in the context of hydrogen generation from biomass. Assessment of the performance of biomass gasification
process regarding the product gas yield and composition can be performed using mathematical models. Among the
different mathematical models, thermodynamic equilibrium models are simple and useful tools for the first estimate
and preliminary comparison and assessment of gasification process. A stoichiometric thermodynamic equilibrium
model is developed here, and its performance is validated for steam gasification and air-steam gasification. The
model is then used to assess the feasibility of different biomass feedstock for gasification based on hydrogen yield
and lower heating value.
© 2016
© 2015TheTheAuthors.
Authors. Published
Published by by Elsevier
Elsevier Ltd.Ltd.
This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
Peer-review under responsibility of the organizing committee of RAEREST 2016.
Peer-review under responsibility of the organizing committee of RAEREST 2016
Keywords:Biomass; Stoichiometric thermodynamic equilibrium model; Steam gasification; Air-steam gasification.

1. Introduction

Sustainable energy generation has to become the key focus of the current world energy scenario relying on clean
and renewable energy resources. Hydrogen produced from carbon neutral biomass is significant in this perspective.
Biomass gasification is a thermochemical process by which hydrogen is extracted from biomass. In gasification
biomass is partially oxidised at a temperature around 800 °C to produce a gaseous mixture widely known as
producer gas. It consists mainly of hydrogen, carbon monoxide, carbon dioxide and methane along with char and
tar. The quality and composition of the product gas generated in gasification in a gasifier is influenced by the nature
of biomass, gasification media and key operating parameters like reactor temperature (T), equivalence ratio (ER),
and steam to biomass ratio (SBR). It is possible to evaluate the performance of gasification regarding product gas
*Corresponding author Tel. +91 9446179554
Email address: joelpull@gmail.com

2212-0173 © 2016 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
Peer-review under responsibility of the organizing committee of RAEREST 2016
doi:10.1016/j.protcy.2016.08.194
Joel George et al. / Procedia Technology 25 (2016) 982 – 989 983

yield and composition and make predictions concerning the quantity and quality of product gas by using
mathematical models [1].
Researchers have formulated different models like thermodynamic equilibrium [2, 3], kinetic [4, 5]
computational fluid dynamics [6, 7] and artificial neural network [8, 9]) to analyse gasification [10]. Arnavat et al.
[11], Ahmed et al. [12] and Baruah et al. [1] present a detailed review of different gasification models. The review
of the gasification models reveals that the thermodynamic equilibrium model (TEM) is a simple and useful tool for
the first estimate and preliminary comparison and judgement of gasification process. According to Li et al. [13],
TEM is valuable as it can predict the thermodynamic limits. TEM though independent of gasifier design
conveniently studies the influence of fuel and process parameters and makes a reasonable prediction of the
maximum achievable yield of a particular product useful for a designer.

2. Model Development

The concept of thermodynamic equilibrium model is based on the second law of thermodynamics as applied to
chemical reacting systems [14]. Accordingly all spontaneous processes occurring in a chemical system proceed in
the direction so as to increase overall entropy. When system composition reaches the situation where overall entropy
is maximum equilibrium state is attained. The analysis is then carried out through different approaches to determine
the equilibrium composition. A stoichiometric thermodynamic equilibrium model (STEM) based on specific
chemical reactions are used here for the estimation of product gas composition. Thermodynamic equilibrium models
do not require any knowledge of the mechanisms of transformation. Moreover, they are independent of the reactor
and not limited to a specified range of operating conditions. A STEM is formulated to assess the hydrogen yield
from different locally available biomasses at specified operating conditions to select the most appropriate one.
Biomass gasification being a complex process the formulation of the mathematical model requires certain
assumptions as follows:
1. The biomass is modelled considering the carbon, hydrogen and oxygen atoms only.
2. The gasifier is a steady state system with uniform pressure and temperature.
3. At equilibrium the reaction system achieves the most stable composition.
4. Gases H2, CO, CO2, CH4, H2 O and N2 considered in the reaction system behaves ideally.
5. Gasification reaction rate is fast enough and residence time is long enough for the equilibrium state.
6. No tar is supposed to leave the reaction system at the end of the process.
7. The fraction of char that bypasses the reaction zone remains unreacted.

2.1. Problem formulation and description

Let M kg of dry ash free biomass produce 1 Nm3 of product gas in a biomass gasification system maintained at
1 atm. The elemental composition of biomass based on ultimate analysis is expressed in weight percentages of
carbon (C), hydrogen (H), and oxygen (O). As biomass contains only negligible amount of nitrogen and sulphur
they are not considered. MO and AS respectively denote the moisture and ash percentage of the biomass based on
proximate analysis. Let VH2 ,VCO ,VCO2 ,VCH4 ,VH2O , and VN2 represent the volume fractions of hydrogen (H2),
carbon monoxide (CO), carbon dioxide (CO2), methane (CH4), water vapour (H2O) and nitrogen (N2), respectively
present in 1 Nm3 product gas. AR represents the mass of air in kg/kg of biomass while OA and NA are mass fractions
of oxygen and nitrogen respectively in the air supplied. S indicates the mass of super-heated steam provided/kg of
dry ash free biomass while W denotes the mass of moisture content of biomass feedstock/kg of dry ash free biomass.

Global gasification reaction equation is modified as follows as per the assumptions made.
(VH  VCO  VCO  VCH  VH O  VN )
C H O S W AR * O A AR * N A 2 2 4 2 2
M *(   )  M *( )  M *(  )o (1)
12 2 32 18 32 28 22.4

The carbon conversion efficiency (B) of gasification process will be less than 100% as a small fraction of char
bypasses the reaction zone. Experimental investigations [15, 16] on carbon conversion efficiency reveal that it varies
from 85 to 95% depending on the type of gasification process. Baruah et al. [1] observed that the carbon conversion
984 Joel George et al. / Procedia Technology 25 (2016) 982 – 989

efficiency increases with increase in reactor temperature up to a certain reactor temperature and then remained
constant. A char conversion model is formulated with these inputs after repeated trials. Modified global gasification
reaction equation contains seven unknowns which include the volume fractions of the constituent gases,
VH ,VCO ,VCO ,VCH ,VH O ,VN and the mass of dry ash free biomass (M) to produce 1 Nm3 product gas. Four linear
2 2 4 2 2

equations can be formed based on carbon, hydrogen, oxygen, and nitrogen molar mass balance between inflow and
outflow streams. A fifth linear equation can be formed based on the stated assumption that the volume fractions of
all constituents of the product gas add up to 1. The remaining two equations can be formed by assuming that all the
reactions occurring in the gasification zone are in thermodynamic equilibrium.

C
Carbon balancing; VCO  VCO  VCH 22.4 * B * ( )*M (2)
2 4 100

S H W
Hydrogen balancing: VH  2 * VCH  VH O 22.4 * (( )( )  ( )) * M (3)
2 4 2 18 200 18

O
Oxygen balancing: 0.5 * VCO  VCO  0.5 * VH O (0.623 * ( S  W )  0.701* (( )  ( AR * ER * 0.23))) * M (4)
2 2 100

N
Nitrogen balancing: VN (0.8 * ( )  0.8 * AR * ER * 0.75) * M (5)
2 100

As the product of gasification is assumed to be 1 Nm3, the volume fractions of all the constituents add up to 1.

VH  VCO  VCO  VCH  VH O  VN 1 (6)


2 2 4 2 2
Considering the prominent reactions occurring in the gasification zone:

Boudouard reaction: C  CO ƒ 2CO (7)

Water gas reaction: C  H 2O ƒ CO  H 2 (8)

Water gas shift reaction: CO  H 2O ƒ CO2  H 2 (9)

Methane reaction: C  2 H 2 ƒ CH 4 (10)

Combinations of equations (7) and (8) will give the water gas shift reaction equation (9). Hence equations (9) and
(10) are only considered to formulate the equations for equilibrium constants.
The equilibrium constant K1 of water gas shift reaction (9) is given by,

PCO * PH VCO * VH
K1 2 2 2 2 (11)
PCO * PH O VCO * VH O
2 2

VCO * VH O * K1 VCO * VH (12)


2 2 2

The equilibrium constant K2 of methane reaction (10) is given by,


Joel George et al. / Procedia Technology 25 (2016) 982 – 989 985

PCH 4 VCH 4
K2 (13)
( PH 2 ) 2 (VH 2 ) 2
2
(VH ) * K 2 VCH (14)
2 4
Equilibrium constants K1 and K2 can be expressed as functions of temperature [3] considering product gas as ideal.

5878 58200
K1 exp(  1.86ln T  0.27 u103 T   18) (15)
T T2

7082.842 7.467 u103 2.167 u106 2 0.702 u105


K2 exp(  6.567 ln T  T T   32.541) (16)
T 2 6 2T 2

Thus to determine seven unknowns there are five linear equations (2, 4, 6, 8 and 10) and two non-linear equations
(12 and 14). Volume fractions of the product gas constituents and the mass of dry ash free biomass consumed can be
obtained by simultaneously solving these equations using Newton – Raphson method in MATLAB platform which
uses the results of ultimate and proximate analysis of biomass as input data. Once the volume fractions are
determined the lower heating value of the product gas can also be calculated from the gas composition and is
expressed in volume basis [17] as shown in equation (17).

LHV 10.79*V  12.26*V  35.81*V (17)


H CO CH
2 4
2.2 Model validation and performance assessment

Stoichiometric equilibrium model though simple, deviations from the experimental results are likely to occur due
to the various assumptions considered during the formulation of the model. In reality, the reaction system might not
have reached equilibrium. To account these factors, modifications are made in the estimated values of equilibrium
constants. These changes will augment the predictions. The accuracy of the model is checked by comparing
predicted gas composition from the proposed model with experimental results. The error in prediction is estimated
with statistical parameter of root mean square (RMS) value [2, 18, 19, 20] shown in equation (18).

2
¦ (Xe - X p )
RMS (18)
N
Xe, Xp and N are experimental data, predicted value, and number of observations, respectively.

Assessment of STEM is carried out by comparing the experimental results (EXP) based on the investigations by
Loha et al. [18] for steam gasification and Turn et al. [21] for air-steam gasification respectively, with the
predictions made by STEM. The chemical characteristics of rice husk and saw dust used in the experiments are
presented in Table 1.

Table 1. Chemical characteristics of biomass


Ultimate Analysis (% Wt.) Proximate Analysis (%Wt.)

Biomass Carbon Hydrogen Oxygen Nitrogen Fixed Carbon Moisture Volatiles Ash Reference
Rice Husk 49.07 3.79 46.42 0.63 14.99 9.95 55.54 19.52 [18]
Sawdust 48.01 6.04 45.43 0.15 18.7 7.5 76.78 0.32 [21]

Loha et al. [18] performed steam gasification in a fluidised bed gasifier using rice husk as the biomass at an
average bed temperature of 1023 K with different steam to biomass ratio (SBR). A comparison between the gas
986 Joel George et al. / Procedia Technology 25 (2016) 982 – 989

yields as volume percentages for EXP and STEM for different SBR is represented using a cluster bar chart in Fig. 1.
Comparison of hydrogen yield for different cases is separately shown in Fig. 2. It is observed that the error in
prediction (RMSE) for different cases is less than 3.7 which is within the reasonable limits.

Gas Yield Vol. %) 60


50
40
30
20
10
0
EXP STEM EXP STEM EXP STEM EXP STEM
0.6 1 1.32 1.32
SBR
H2 CO CO2 CH4 RMSE

Fig. 1. Comparison of gas yield between the experimental results (EXP) and STEM for different SBR at 1023 K

60
Hydrogen Yield (Vol. %)

55
50
45
40
35
30
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
SBR
EXP STEM

Fig. 2. Comparison of hydrogen yield between the experimental results (EXP) and STEM for different SBR at 1023 K

50
Gas Yield (Vol %)

40
30
20
10
0
EXP STEM EXP STEM EXP STEM EXP STEM EXP STEM
1023 1073 1123 1173 1223
Temperature (K)

H2 CO CO2 CH4 RMS

Fig. 3. Comparison of gas yield between the experimental results (EXP) and STEM for different temperature at SBR=1.4 and ER =0.18
Joel George et al. / Procedia Technology 25 (2016) 982 – 989 987

Assessment of STEM is further carried out by comparing the experimental results (EXP) from the investigations
by Turn et al. [21] on biomass air-steam gasification with the predictions made by STEM. Gasification is performed
in a fluidised bed gasifier using sawdust as the biomass at different bed temperature while keeping steam to biomass
ratio (SBR) and equivalence ratio (ER) as 1.4 and 0.18 respectively throughout. A comparison between the gas
yields at various temperatures in volume percentages is represented in Fig. 3 and hydrogen yield for different cases
in Fig. 4. It is observed that the error in prediction (RMSE) for different cases is less than 3.5 except in the first
instance (RMSE above 5). The predictions made by STEM are within the reasonable limits.

60
Hydrogen Yield (Vol %)

50

40

30

20
1000 1050 1100 1150 1200 1250
Temperature (K)
EXP STEM

Fig. 4. Comparison of hydrogen yield between the experimental results (EXP) and STEM for different temperature at SBR=1.4 and ER =0.18

3. Assessment of feedstock’s feasibility with STEM

Hydrogen yield and heating value of product gas generated in gasification depend on the biomass feedstock. The
feasibility of different biomass feedstock for gasification can be assessed using STEM based on hydrogen yield and
lower heating value (LHV). Appropriate biomass for gasification can be selected from the locally available lot.
Chemical characteristics of seven different biomasses along with their predicted hydrogen yields and heating values
are presented in Table 2. The predictions are made at standard operating conditions of average bed temperature of
1073 K and SBR = 1 for steam gasification and for air-steam gasification, at ER=0.22

Table. 2. Chemical characteristics of locally available biomass and their predicted hydrogen yields and lower heating value

Air-steam
Composition (%) Steam Gasification Gasification
Gas LHV Gas LHV
Hydrogen Yield (MJ/ Hydrogen Yield (MJ/
Feed Stock C H O Mo As Yield (g) (Nm3) Nm3) Yield (g) (Nm3) Nm3)
Coffee Husk (BM1) 42.1 6.33 49 10.7 7.8 69.65 1.51 10.44 37.97 1.75 5.83
Saw dust (BM2) 46.5 5.82 47.5 7 1 74.46 1.64 10.48 36.98 1.86 5.88
Rubber Seed Shell (BM3) 47.15 8.34 42.3 13.1 4.4 85.4 1.78 11.66 52.18 2.21 6.34
Coconut shell (BM4) 45.6 5.61 48.2 8 4 72.45 1.61 10.31 35.57 1.8 5.8
Coir pith (BM5) 44.1 4.09 51.2 10 13 64.24 1.49 9.37 27.17 1.58 5.32
SCF (BM6) 36.6 4.56 57.4 13.1 15.1 51.47 1.21 8.72 24.7 1.3 4.91
Rice Husk (BM7) 34.35 5.22 57.7 12 18 50.17 1.16 8.95 26.57 1.28 5.04
988 Joel George et al. / Procedia Technology 25 (2016) 982 – 989

4. Results and discussion

BM7 BM7
BM6 BM6
Biomass

BM5 BM5

Biomass
BM4
BM4
BM3
BM3
BM2
BM2
BM1
BM1
0 10 20 30 40 50 60 70 80 90
0 2 4 6 8 10 12
Hydrogen Yield (g)
Lower Heating Value (MJ/Nm3}

Hydrogen Yield (g) Hydrogen Yield (g) LHV (MJ/Nm3 LHV (MJ/Nm3

Fig. 5. (a) Predicted hydrogen yields (b) Lower heating value from steam gasification and air-steam gasification from different biomasses

Figure 5 (a) illustrates the predictions on hydrogen yield from steam gasification and air-steam gasification for
various biomasses. Steam gasification will generate more hydrogen because of the water gas shift reaction occurring
in the presence of large quantities of super-heated steam [10]. However, the hydrogen generated depends upon the
H/C and O/C ratios. Biomass with higher H/C ratio and lower O/C will have higher hydrogen yield [10].
Predictions on the hydrogen yield from rubber seed shell (BM3) and rice husk (BM7) reveal that the former (having
higher H/C ratio and lower O/C) will generate more hydrogen than rice husk (with lower H/C ratio and higher O/C
ratio). Among the different biomasses, rubber seed shell (BM3) with higher H/C ratio will yield maximum hydrogen
(85.4 g). Other biomasses like sawdust (BM2), coconut shell (BM 4), and coffee husk (BM1) have the potential of
appropriate biomass for gasification. They can be recommended for high end use applications. However, coir pith
(BM5), rice husk (BM7) and shredded coconut frond (SCF) (BM6) have low hydrogen yield because of higher O/C
ratio. Being locally available in plenty they can be used for lower end use applications. Air-steam gasification
considered for hydrogen generation makes use of the exothermic heat generation within the process to sustain the
gasification reactions which are generally endothermic. This action will reduce the external heat requirement
necessary to maintain the gasification. The quantification of reduction in heat input and other economic concerns are
not the subject of interest in this work.
The predictions of heating value for different biomasses undergoing steam gasification and air-steam gasification
are presented in Fig. 5(b). Biomasses are characterised by high O/C ratio. It is observed from Fig. 5(b) that
biomasses with higher hydrogen yields have higher heating value. Rubber seed shell (BM3) having higher hydrogen
yield will have more heating value. It is predicted that steam gasification will generate producer gas having heating
value more than 10 MJ/Nm3 compared to air-steam gasification that produces product gas with heating value more
than 5 MJ/Nm3. Among the different biomasses, Rubber seed shell generates product gas with heating value 11.66
MJ/Nm3 (steam gasification) and 6.34 MJ/Nm3 (air-steam gasification). SCF with more O/C ratio will generate
product gas with lower heating value.

5. Conclusion

Appropriate selection of biomass is necessary to maximize the yield of the product gas in the gasification
process. Thermodynamic equilibrium models are simple mathematical tools to do a preliminary feasibility study in
the gasification process. The STEM model formulated for this purpose is tested, and its performance is validated
using the statistical parameter RMSE. This model is used to predict the hydrogen yield and heating value of the
locally available biomasses. It is observed that the feedstock with maximum H/C ratio and minimum O/C ratio have
the potential to generate hydrogen having yields above 85 g/kg of dry ash free biomass and heating value above
Joel George et al. / Procedia Technology 25 (2016) 982 – 989 989

11 MJ/Nm3 for steam gasification. However for air-steam gasification, yield and heating value are reduced by about
50% with the advantage of lower heat input for the gasification process.

Acknowledgments
Authors gratefully acknowledge the financial support provided by Ministry of New and Renewable Energy
through R&D project on ‘Investigation on bio-hydrogen production by thermo-chemical method in fluidised bed
gasifier under catalytic support and its utilisation’ (No. 103/181/2010-NT).

References
[1] Baruah D, Baruah DC. Modeling of biomass gasification: a review. Renewable and Sustainable Energy Reviews 2014; 39: 806–815.
[2] Jarungthammachote S, Dutta A. Thermodynamic equilibrium model and second law analysis of a downdraft waste gasifier. Energy, 2007;
32:1660-1669.
[3] Zainal ZA, Ali R, Lean CH, Seetharamu KN. Prediction of performance of a downdraft gasifier using equilibrium modeling for different
biomass materials. Energy Conversion Management 2001; 42:1499–1515.
[4] Inayat A, Ahmad MM, Yusup S, Mutalib MIA. Biomass steam gasification with in-situ CO2 capture for enriched hydrogen gas production:
a reaction kinetics modelling approach. Energies 2010; 3:1472–1484.
[5] Sreejith CC, Muraleedharan C, Arun P. Air–steam gasification of biomass in fluidized bed with CO 2 absorption: A kinetic model for
performance prediction. Fuel Processing Technology 2015; 130:197–207.
[6] Gao J, Zhao Y, Sun S, Che H, Zhao G, Wu J. Experiments and numerical simulation of sawdust gasification in an air cyclone gasifier.
Chemical Engineering Journal 2012; 213:97–103.
[7] Janajreh I, Al Shrah M. Numerical and experimental investigation of downdraft gasification of wood chips. . Energy Conversion
Management 2013; 65:783–792.
[8] Arnavat MP, Herna´ndez JA, Bruno JC, Coronas A. Artificial neural network models for biomass gasification in fluidized bed gasifiers.
Biomass and Bioenergy 2013; 49: 279–289.
[9] Sreejith CC, Muraleedharan C, Arun P. Performance prediction of fluidised bed gasification of biomass using experimental data-based
simulation models. Biomass Conversion and Biorefinery 2013; 3:1–22.
[10] Basu P. Biomass gasification and pyrolysis: practical design and theory. MA, U.S.: Elsevier; 2010.
[11] Arnavat MP, Bruno JC, Coronas A. Review and analysis of biomass gasification models. Renewable and Sustainable Energy Reviews
2010; 14:2841–2851.
[12] Ahmed TY, Murni M, Ahmad MM, Yusup S, Inayat A, Khan Z. Mathematical and computational approaches for design of biomass
gasification for hydrogen production: A review. Renewable and Sustainable Energy Reviews 2012; 16:2304–2315.
[13] Li X, Grace JR, Watkinson AP, Lim CJ, Ergudenler A. Equilibrium modeling of gasification: a free energy minimization approach and its
application to a circulating fluidized bed coal gasifier. Fuel 2001; 80:195-207.
[14] Mahishi M, Goswami D. Thermodynamic optimization of biomass gasifier for hydrogen production. International Journal of Hydrogen
Energy 2007; 32:3831–3840.
[15] Van der Meijden CM, Veringa HJ, Rabou LPLM. The production of synthetic natural gas (SNG): A comparison of three wood gasification
systems for energy balance and overall efficiency. Biomass and Bioenergy 2010; 34:302-311.
[16] Van der Drift A, Van Doorn J, Vermeulen JW. Ten residual biomass fuels for circulating fluidized-bed gasification. Biomass and
Bioenergy 2010; 20:45-56.
[17] Kaewluan S, Pipatmanomai S. Potential of synthesis gas production from rubber wood chip gasification in a bubbling fluidised bed gasifier.
Energy Conversion and Management 2011, 52:75–84.
[18] Loha C, Chatterjee PK, Chattopadhyay H. Performance of fluidized bed steam gasification of biomass: modeling and experiment. Energy
Conversion and Management 2011; 52: 1583–1588.
[19] Sreejith CC, Muraleedharan C, Arun P. Equilibrium modeling and regression analysis of biomass gasification. Journal of Renewable and
Sustainable Energy 2012; 4(062134):1-19.
[20] Rupesh S, Muraleedharan C, Arun P. Analysis of hydrogen generation through thermochemical gasification of coconut shell using
thermodynamic equilibrium model considering char and tar. International Scholarly Research Notices 2014; 2014:1–9
[21] Turn S, Kinoshita C, Ishimura D. An experimental investigation of hydrogen production. International Journal of Hydrogen Energy 1998;
23:641–648.

S-ar putea să vă placă și