Sie sind auf Seite 1von 11

international journal of hydrogen energy 35 (2010) 49704980

Available at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/he

Thermodynamic analysis of hydrogen production from biomass gasication


M.K. Cohce*, I. Dincer, M.A. Rosen
Faculty of Engineering and Applied Science, University of Ontario Institute of Technology, 2000 Simcoe Street North, Oshawa, Ontario, Canada L1H 7K4

article info
Article history: Received 19 June 2009 Received in revised form 26 August 2009 Accepted 31 August 2009 Available online 7 October 2009 Keywords: Biomass Gasication Hydrogen Thermodynamics Energy Exergy Efciency Oil palm shell SMR

abstract
An investigation is reported of the thermodynamic performance of the gasication process followed by the steam-methane reforming (SMR) and shift reactions for producing hydrogen from oil palm shell, one of the most common biomass resources. Energy and exergy efciencies are determined for each component in this system. A process simulation tool is used for assessing the indirectly heated Battelle Columbus Laboratory (BCL) gasier, which is included with the decomposition reactor to produce syngas for producing hydrogen. A simplied model is presented here for biomass gasication based on chemical equilibrium considerations, with the Gibbs free energy minimization approach. The gasier with the decomposition reactor is observed to be one of the most critical components of a biomass gasication system, and is modeled to control the produced syngas yield. Also various thermodynamic efciencies, namely energy, exergy and cold gas efciencies are evaluated which may be useful for the design, optimization and modication of hydrogen production and other related processes. Crown Copyright 2009 Published by Elsevier Ltd on behalf of Professor T. Nejat Veziroglu. All rights reserved.

1.

Introduction

Biomass, a relatively large energy source globally which includes wood, municipal solid wastes and agricultural residues, is being investigated in various countries as a potentially signicant renewable resource. Biomass is derived from solar energy. Biomass is relatively clean compared to other sources of energy, as it releases no net CO2 emissions when carefully managed since CO2 is xed by photosynthesis during biomass growth and is released during utilization. This form of energy can be converted to gaseous fuel through thermochemical

gasication [1]. Such a fuel can be used for various tasks, including producing hydrogen, which can be used cleanly and efciently as a fuel in combustion engines and fuel cells. Hydrogen is likely to be an important energy carrier in the future. Presently, it can be produced by the steam reforming of natural gas, coal gasication and water electrolysis among other processes. However these current processes are not sustainable because they use fossil fuels or electricity from non-renewable resources. Hydrogen production can be made more sustainable if it is produced from sustainable energy resources. In this regard, alternative thermochemical

* Corresponding author. E-mail addresses: mehmet.cohce@uoit.ca (M.K. Cohce), ibrahim.dincer@uoit.ca (I. Dincer), marc.rosen@uoit.ca (M.A. Rosen). 0360-3199/$ see front matter Crown Copyright 2009 Published by Elsevier Ltd on behalf of Professor T. Nejat Veziroglu. All rights reserved. doi:10.1016/j.ijhydene.2009.08.066

international journal of hydrogen energy 35 (2010) 49704980

4971

(pyrolysis and gasication) and biological (biophotolysis, water-gas shift reaction and fermentation) processes are practical and can be more sustainable than present processes [2,3]. Overviews of technologies for hydrogen production from biomass have been reported [48]. Many researchers are focusing their research on the gasier portion of this process, as gasication appears to be more favorable for hydrogen production than pyrolysis [9]. The gasier is the most important component in any biomass gasication system [10]. Gasication, which is characterized by partial oxidation, is utilized in numerous clean energy processes including hydrogen production via biomass gasication. Currently, 8085% of the worlds total hydrogen production is derived from natural gas via steam-methane reforming (SMR) [11]. Much research has been reported on the production of hydrogen by SMR [1215], with many studies concentrating on analysing the reforming reactor [16]. Gasier modelling and simulation, using programs such as Aspen Plus, have been ongoing [1719]. Some researchers predict biomass gasication in supercritical water to be a promising technology for hydrogen production for it allows the utilization of wet biomass [20,21]. Recent research has focused on determining energy and exergy efciencies [2225] and improving understanding and data availability with experimental studies. In this study, the gasier, the most signicant part of the system, is analyzed in detail. Exergy analysis is a tool for understanding and improving efciency [4], and is used throughout this investigation in addition to energy analysis. Extensive exergy analyses have been reported using devices, technologies and systems, in a variety of elds. Some examples include exergy assessments of power plants for transportation and power generation, chemical and metallurgical processing facilities and building systems [26,27]. Exergy-based assessments and comparisons have also been performed of such energy technologies as hydrogen production [2833], electricity generation with fuel cells and other technologies, for instance, coal-red and nuclear [34,35], and larger energy systems [36,37]. The aim of the present work is to investigate hydrogen production by thermochemical biomass gasication using energy and exergy methods, and to evaluate the potential of hydrogen production from biomass. A parametric analysis of factors inuencing the thermodynamic efciency of biomass gasication is carried out. Energy and exergy assessments are performed and the effects are determined on both system efciency and hydrogen yield of varying parameters of the fractioned syngas. The system considered has two parts, and both include syngas production as an input for producing hydrogen in the hydrogen plant. The manner is assessed by which hydrogen production is changed by fractioning the syngas streams and by importing methane from external sources. The focus of this research is on the gasier since it requires the most improvement in terms of energy and exergy utilization. The signicance of the present work is its focus on green energy sources and increasing the efciency of systems for hydrogen production. It is anticipated that the results will assist efforts to optimize and enhance the environmental performance of energy systems. The use of computer simulation yields a better understanding of the overall system, and is a particularly useful tool in support of design.

2.

Thermodynamic evaluation

A thermodynamic evaluation of a complex system requires consideration of its components and their characteristics, chemical reactions and thermal losses. Recently, biomass gasication in indirectly heated steam gasiers has received much attention for the conversion of biomass to combustible gas [38,39]. In our simulation, we consider the energy efciency of the gasication reaction as the total energy of the desired products divided by the total energy of the process inputs [28]. For this analysis, the products are taken to be a mixture of H2O, N2, H2, CO2, CH4, CO, NH3 and H2S. Char is assumed to consist of solid carbon (C) and tar is not taken into account in the simulation. Simulations are preformed with the Aspen Plus simulation software, which is utilized in a wide range of industrial applications [18]. A Fortran subroutine is applied to control process yields. In Aspen Plus, streams represent mass or energy ows. Energy streams may be dened as either work or heat streams, of which the latter also contain temperature information to avoid infeasible heat transfer. Mass streams are divided by Aspen Plus into three categories: mixed, solid, and non-conventional (for substances like biomass). Mixed streams contain mixtures of components, which can be in gaseous, liquid and solid phases. The solid phase component in this simulation is solid carbon (C). Thermodynamic properties are dened in the Aspen Plus libraries for chemical components. Components present in the mixed and solid stream classes may participate in phase and chemical equilibrium, and are automatically ashed by Aspen Plus at stream temperature and pressure. Non-conventional components are dened in Aspen Plus by supplying standard enthalpy of formation and the elementary composition (ultimate and proximate analyses) of the components may also be dened. Biomass is characterized in this manner here. Although Aspen Plus calculates enthalpies and entropies for conventional components, ambient temperature and pressure, which are required in evaluations of exergy, are not readily available in the result output. A property termed availability by Aspen Plus is calculated for conventional components, but this does not include chemical exergy. For these reasons, the EES program is used to calculate total exergy (physical and chemical) for each stream in this simulation. The following simplifying assumptions are made in the simulation:  Char only contains solid carbon and ash, and there is no tar yield.  The process occurs at steady state and isothermally, and residence time is not considered. Also catalysis is not used.  The ZNO-bed and the pressure swing adsorption (PSA) system are not included in the energy and exergy calculations.  All gases behave ideally.  Air is considered on a volume basis as 79% nitrogen and 21% oxygen.  A heat stream is used as a heat carrier in Aspen Plus instead of sand.

4972

international journal of hydrogen energy 35 (2010) 49704980

2.1.

Balances

exch

X xi exch RT0 ln xi i
i

(8)

Mass and enthalpy values are evaluated with Aspen Plus. For a general steady-state process, we can write mass and energy balances, respectively, as X
i

_ mi

X
o

_ mo

(1)

Here, xi is the mole fraction and exch the standard chemical i exergy of component i. Standard chemical exergy values used here are taken from model 2 in Szargut et al. [26]. The entropy balance for a steady-ow reacting system can be written as _ X X Qj X _ _ _ mi si mo so Sgen 0 _j T (9)

X
i

_ Ein

X
o

_ Eout

(2)

An overall exergy balance can be written for a steady-state process as follows: X _ Exi 
in

The exergy destroyed due to irreversibility can be expressed as follows: _ _ Exdest T0 Sgen (10)

X

_ Exj


out

X  _ Ex

dest

(3)

where X X _ Exi _ Exj 


in

_ _ _ _ _ Exair Exdrybio Exbiomoist Exst Exmeth _ _ _ _ _ Exprodg Exunconcarb Exst Exexh Wturb

(4) (5)


out

The physical exergy of biomass is zero when it is entering the system at temperature T0 and pressure P0. Thermodynamic properties are needed for the calculation of the chemical exergy of biomass. Since such properties for oil palm shell biomass are not available, a correlation factor for solid biomass (for O/C < 2) is used based on statistical correlations developed by Szargut et al. [29]: b 1:044 0:0160H=C 0:3493O=C1 0:0531H=C 0:0493N=C 1 0:4124O=C (11) The specic chemical exergy for biomass can then be determined as Exch biosolid bLHVbio (12)

These values are available in Table 1 Both physical and chemical exergy inlet and outlet values are determined for the Gasication, Combustion, SMR, HTS and LTS product gases, are used to assess exergy destructions. Some components possess only physical exergy. The specic ow exergy associated with a specied state is expressible by the sum of specic physical and specic chemical exergy: exprodg exph exch The physical exergy dened as exph h ho T0 s so (7) (6)

and the chemical exergy contribution can be calculated for an ideal gas mixture as follows:

Although the magnitude of the physical exergy of biomass is small, it is calculated in this simulation after the drying process. The calculation is performed using the heat capacity of dry biomass which is described by Gronli and Melaaen [37]: Cpbio 1:5 103 T (13)

Table 1 Exergy results for the overall hydrogen plant for Case 1. Exergy ow rate (kJ/h)
Inputs Wet biomass Water Air Methane (CH4) Purchased electricity Total Outputs and destructions Hydrogen Electricity Exhaust Water Total output Exergy destruction Total output and destruction 2,397,523 8,469 3,432 527,500 0 2,936,924

where Cp(bio) is the heat capacity and T is the temperature of the biomass. The change in specic entropy in Eq. (7) can be written for biomass as Ds Z
T T0

Percentage of total exergy inlet


81.0 3.0 1.0 16.0 0.0 100.0

Cp;i dT T

(14)

Equation (14) can be used to calculate the physical exergy with Eq. (7) at the specied temperature. The heat capacity of solid carbon is determined using Aspen Plus property data and substituted into Eq. (7) to nd entropy values, which are used for the thermal exergy calculation in Eq. (7).

653,052 46,800 303,573 49,850 1,053,265 1,883,658 2,936,923

22.3 1.5 10.3 1.7 35.9 64.1 100.0

2.2. Energy, cold gas and exergy efciencies for BCL gasication
Tables 2 and 3 show the properties and gas composition of the main streams and the overall system performance based on the simulation results. The latter can be expressed with the cold gas efciency. This measure is the ratio of the chemical energy of the produced gas to that of the biomass feed energy

international journal of hydrogen energy 35 (2010) 49704980

4973

Table 2 Properties and composition of the main streams in the H2 plant. Quantity COMP Outlet
T ( C) P (bar) Flow rate (kg/h)


Stream SMR outlet


850 28 121.23

HTS Outlet
518 27 121.23

LTS outlet
241 26.5 120.90

PSA inlet
43 25.2 70.695

Off-gas
40 1 69.70

43.5 31 52.06

Dry gas composition (% vol) 0.0 H 2O 46.0 H2 CO 30.0 14.0 CO2 10.0 CH4

49.0 28.0 15.0 6.0 0.01

38.0 40.0 4.00 17.0 1.00

34.0 43.0 1.0.0 21.0 1.0

0.0 65.60 0.40 32.15 1.85

0.0 26.58 0.30 72.32 0.8

content. The cold gas energy efciency of the BCL hydrogen plant is determined as hcg _ mprodg LHVprodg _ mdrybio LHVdrybio (15)

_ _ where Exdest i and Exin i respectively are the exergy destruction rate and the exergy input rate for component i. The steambiomass ratio (STBR) can be expressed as STBR _ mst _ mdrybio (20)

Hydrogen is the desired product in this simulation, so the production efciency can be described as the system efciency in general. Energy efciency h and exergy efciency j values are often evaluated for steady-state processes, and can be written here as follows: hsys _ _ Eprodg Wturb _ air Edrybio Ebiomoist Est Emeth _ _ _ _ E _ _ Exprodg Wturb _ _ _ _ _ Exair Exdrybio Exbiomoist Exst Exmeth (16)

Also, the ratio of exergy destruction xdest for a component can be evaluated by dividing its exergy destruction by the total exergy provided to the system. Here, we can write _ Exdest i xdest P _ Exi in (21)

_ where Exdest i is the exergy destruction for each component P_ and Exi in is the exergy ow of all input material streams. (17)

jsys

_ _ where Eprodg is the rate of product energy output, Exprodg is the _ rate of product exergy output and Wturb is the turbine work rate. Also the energy efciency of component i may be written as follows: hi 1 _ Eout i _ Ein
i

3.

Case study

(18)

_ _ where Eout i and Ein i are the energy output and input rates for component i. Similarly, the exergy efciency for component i may be written as ji 1 _ Exdest i _ Exin i (19)

Table 3 Simulation results for cases 1 and 2. Quantity


Biomass ow rate (wet) (kg/h) Biomass ow rate (dry) (kg/h) Steam input to gasier (kg/h) Syngas fraction to SMR-COMB CH4 input to SMR-COMB (kmol/h) Hydrogen production rate (kg/h) Steam-biomass ratio (STBR) Cold gas efciency, hcg System energy efciency, hsys System exergy efciency, jsys

Case 1
166.67 88.40 33.17 0.196 0.0 5.53 0.38 0.30 0.24 0.22

Case 2
166.67 88.40 33.17 0.0 0.475 6.91 0.38 0.40 0.27 0.25

Heating and drying are endothermic processes that require a source of heat. Heat can be supplied by an external source via indirect heating. In gasication, indirect heating allows us to have indirectly heated gasication. More often, a small amount of air or oxygen, typically not more than 25% of the stoichiometric requirement for complete combustion of the fuel, is input for the purpose of partial oxidation, which releases sufcient heat for drying and pyrolysis as well as for the subsequent endothermic chemical reactions. These include the carbonoxygen, boudovard, carbonwater and hydrogenation reactions [35]. The system simulated in this paper includes the gasication plant and the hydrogen plant. Related energy efciency and economic analyses have been performed [38]. In this investigation, the system analyzed by those authors is modied and energy and exergy efciencies are evaluated for varied conditions (e.g., water streams and the amount of entered biomass). In this study, we focus on the gasier. The R-GIBBS block reactor uses single-phase chemical equilibrium, or simultaneous phase and chemical equilibrium, by minimizing the Gibbs free energy, subject to atom balance constraints. This block reactor is useful when the temperature and pressure are known and the reaction stoichiometry is unknown. The latter reactor and the decomposed (RYIELD) reactor combined have been used to model the BCL low-pressure indirectly heated

4974

international journal of hydrogen energy 35 (2010) 49704980

Table 4 Proximate and ultimate analyses and other data for oil palm shell [41].
Proximate analysis (wt% dry basis) Volatile matter Fixed carbon Ash Ultimate analysis (wt% dry basis) C H O N S Moisture content (wt %) Average particle size (mm) Molecular formula Lower heating value (MJ/kg) 73.74 18.37 2.21

53.78 7.20 36.30 0.00 0.51 5.73 0.250.75 CH1.61O0.51 22.14

gasier. The feedstock used for this analysis is oil palm shell, delivered at 50 wt% moisture; the ultimate and the proximate analyses for the feed used in this study are given in Table 4. The plant capacity is designed to be 2000 dry tonne/day (83.3 t/ h), and the lower heating value (LHV) of the dry biomass is 22.14 MJ/kg. In this paper we divide the system into two parts. The rst is the gasication plant in Fig. 1, which produces syngas for the second part of the system, which is the hydrogen plant. In this assessment, we consider two cases that have been designed for producing hydrogen via gasication. In the rst case, 19.6% of the produced syngas and the total amount of off-gas are fed to the steam-methane reformer combustion (SMR-COMB) in order to supply heat for the SMR in the hydrogen plant in Fig. 1. Also, in the rst case energy and exergy analyses for the entire system are carried out. In the second case, instead of feeding the SMR-COMB with fractioned syngas, it is fed by just enough methane gas (CH4) from external sources to enhance the hydrogen yield at the end of the simulation. As a case study, biomass and ue gas are mixed in the dry reactor in order to evaporate the water, and dry the wood from 50% to 5.7% moisture content. After the drying process is completed in the stoichiometric reactors (RSTOIC) in Fig. 1, the gas passes through the decomposition (RYIELD) reactor. Normally the BCL device is an indirectly heated gasier consisting of two main reactors [38]: the gasier and the combustor. However, in this simulation the BCL unit contains three main reactors which are the decomposer, the gasier and the combustor. First the biomass is decomposed in the RYIELD reactor (this reactor is used in the Aspen plus simulation); this reactor simulates the decomposition of the feed at low temperature (394 K, 1 atm). In this step, biomass is converted into its constituent components including carbon (C), hydrogen (H2), oxygen (O2), sulphur (S), nitrogen (N2) and ash, by specifying the yield distribution according to the biomass ultimate analysis. These components enter the Gibbs reactor (at 1 atm and 1162 K) to produce syngas using steam (at 923 K and 1 atm), as seen in Table 5. The heat of combustion of the actual indirectly heated gasier system is transferred to the gasier by recirculating hot inert material, usually sand.

In this simulation, however, it is a designed heat stream using just enough heat to supply the gasier heat demand and the gas cleaning section (see Fig. 1). At the same time combustion occurs in a third reactor, which is fed with methane gas (CH4) from an external supply and char generated by the gasier. There are two combustors which operate at different temperatures; the rst combustor runs at a temperature of 1255 K, while the second runs at 1355 K with 15% excess air. Fig. 1 shows that the syngas enters the scrubber which is designed for syngas cleaning. During this process some of the toxic gas is cleaned and water in the syngas is condensed. After entering the scrubber the syngas passes to the separator (SP1), from which part goes to the SMR-COMB. Note that in this study, the rst case has the fraction of the product stream at 19.6% but the fraction of the product stream is 0% for the second case. After, the syngas passes through the vestage compressor system, which has polytropic efciency of 79% for each compressor stage and a mechanical efciency of 95%. The syngas is cooled, the preferred method being air cooling as it avoids excess pressure losses. After the compression and cooling processes, the syngas pressure increases from 1 to 31 bar while the temperature increases by 43 K. Before reaching the ZnO-Bed, the syngas is heated to 653 K because the ZnO-Bed cannot function at a lower temperature [40]. After sulphur cleaning in the ZnO-Bed, the syngas undergoes three main reactions: steam-methane reforming (for which the main reaction is CH4 H2O $ CO 3H2), hightemperature shift (for which the main reaction is CO H2O $ CO2 H2), and low-temperature shift. The watergas shift reaction is usually performed in two stages in commercial processes: a high-temperature shift (HTS) in the range of 643693 K and a low-temperature shift (LTS) in the range of 473523 K [38]. The sulphur-free syngas mixes with the steam from the superheater in mixer 1 to drive SMR. The reforming condition is xed at 1123 K and 28 bar because methane conversion decreases at high pressure. After the syngas enters the water heat boiler (WHB3), it cools to 677 K before entering the high-temperature shift reactor (HTS). While reducing carbon monoxide in the shift reactor, the hydrogen yield increases by almost 7.5%. As shown in Table 2, after the HTS, the syngas passes through a heat exchanger (HE6) and superheater, where its temperature reduces to 473 K. The nal treatment for the syngas before pressure swing adsorption (PSA) is the low-temperature shift reactor (LTS), where the carbon monoxide is converted and hydrogen content increased. The outlet of the LTS has the highest hydrogen ow rate (6.51 kg/h here). The PSA unit puries the syngas by separating the hydrogen from the other components in the shifted gas stream, mainly CO2 and unreacted CO, CH4 and other hydrocarbons. Based on studies and data from industrial gas producers, the shifted gas stream must contain at least 70 mol% hydrogen before it can be economically puried in the PSA unit. For the present analysis, the concentration of hydrogen in the shifted stream prior to the PSA unit is between 60 and 65 mol%. Therefore, part of the PSA unit hydrogen product stream is recycled back into the PSA feed. For a 70 mol% hydrogen PSA feed, a hydrogen

international journal of hydrogen energy 35 (2010) 49704980

4975

wet biomass

SYNGAS

HOPPER DRY-REAC 3 4 2

FLUEGAS DRY-FLSH

GASIFIER
RYIELD

DRYBIO

CYCLONE

Q 8 BIOGAS QCOMB WHB 7 COMB1 CHAR RGIBBS

CH4 SP 1 52 13

AIR

CH4 addition
STEAM

WHB2 SYN-COMP 14 SCRUBBER 15 ZNO-BED 16 HEATER1 32 SMR-COMB Q2 22 WHB3 38 39 54 HE2 HTS 40 74 18 34 23 SP2 SP3 48 17 55 HE1 MIX2 19 21 SMR 51 STEAM2 50 9 MIX5 49 TURBINE ELECT W

35 Q1 HE3 24 47 46

HE6 53

41

25

HE4 37

SH

26

42 HE5 MIX3

EC2

36

27 TOSTACK 33

LTS

PSA MIX4 73 31 COOLER 30 EC 1 29 44 72 SP4 H2-REC L6 43 45 EC3 28

P2 H2 WATER1 WATER2

P1

Fig. 1 Aspen Plus simulation ow diagram of a process for hydrogen production by biomass indirectly heated gasication.

4976

international journal of hydrogen energy 35 (2010) 49704980

Table 5 Conditions at the gasier outlet. Quantity


Gasier outlet temperature ( C) Combustor 1 outlet temperature ( C) Gasier outlet composition (kg/h) H2O H2 CO CH4 CO2 NH3 H2S N2 C (solid) Ash

HE1
SMR

Value
890 982
CH4 SMR-COMB 32 Q2

21

38.23 3.52 31.87 5.93 22.98 0.08 0.08 0.70 18.02 0.77

22

40

54

Split syngas and off gas from HE3

AIR from HE2

recovery rate of 85% is typical with a product purity of 99.9% by volume [40]. For the SMR-COMB unit, air enters at 298 K and 1 atm and passes through two heat exchangers (HE2 and HE4). The temperature of the air rises to 1060 K, while concurrently the off-gas from the PSA and the fractioned syngas passes through heat exchanger 5 and heat exchanger 3 at 720 K and 1 atm, and enters the SMR-COMB to supply heat for SMR (Case 1). In addition heat exchanger 3 and WHB3 produce the steam for the turbine which produces electricity. The effects of fractioned syngas on the hydrogen production yield are shown in Fig. 2. That gure shows that the steam-methane reformer combustion (SMR-COMB) is supplied with methane gas at a rate of 0.76 kg/h CH4. The equivalent thermal contribution of this gas is the same as the 19.6% of the fractioned syngas in Case 1.

Fig. 2 Aspen Plus simulation ow diagram of a process for steam-methane reforming with heat supplied by combustion of methane gas (CH4) from external source.

4.

Results and discussion

We now report the results of the energy and exergy analyses, including the energy and exergy efciencies and exergy destructions for each component. The results are reported for Case 1 in Table 1. It is demonstrated that the inlet and outlet exergy ows for the hydrogen plant are mainly attributable to the energy and exergy inlet with the biomass and the methane gas. The produced electricity also contributes to the system products and efciency, on energy and exergy bases. Further, the exergy losses are observed to be due to emissions and internal consumptions associated with chemical reactions, particularly those related to combustion and gasication. Note that inlet exergy values are evaluated for fuels on an LHV basis. Some key results from the simulations for the two cases are presented and compared in Table 3. In this simulation at 1162 K and 1 atm, for the low-pressure indirectly heated gasier (R-GIBBS) with the decomposition reactor (RYIELD), the maximum hydrogen production efciency for the rst case is found to be 24% while exergy efciency is 22%, as determined using Eqs. (16) and (17). These values are relatively low because of losses for the gasier, combustor and SMR. For

the second case, performance improvements are noted, as the energy efciency is found to be 27% and the exergy efciency 25%. The efciency increase occurs because importing methane gas from external sources increases the hydrogen production rate from 5.53 kg/h to 6.91 kg/h after the PSA unit. The energy and exergy efciency values for Cases 1 and 2 differ because of the improvement applied to the system in Case 2. These differences are not attributable to uncertainties associated with computer simulation. The improvements in Case 2, as previously noted, include non-fractioned syngas which passes sequentially through the SMR, HTS and LTS units in order to increase the hydrogen yield in the reforming and shift reactions. In Case 1 which used fractioned syngas, 19.6% of the gaseous product cannot pass the SMR, HTS and LTS processes so less hydrogen is produced, which affects the overall system energy and exergy efciencies. Also, the cold gas efciencies differ because the second case has a higher productivity rate than the rst. The products from this process are hydrogen and electricity, but other energy streams also exit. There are two sources of ue gas: the char combustor and the second combustor (SMR-COMB). Together, their energy contents account for about 4% of the energy in the dried biomass. The simulated hydrogen production ow rates for Cases 1 and 2 are shown in Fig. 3. It is seen there that the hydrogen production is affected by fractioned and unfractioned syngas through the component split 1 (point 13 in Fig. 1). It is observed for the rst case that the hydrogen production rate is lower than for the second case, in which the SMR-COMB unit is fed with fractioned syngas. At the end of the simulation, therefore, the fractioned syngas is prevented from passing through the SMR, HTS and LTS processes, which are where the hydrogen yield is increased. In the second case, the unfractioned syngas passes through of the SMR, HTS and LTS units. The energy and exergy efciencies of the main components involving chemical reactions are shown in Fig. 4. The energy

international journal of hydrogen energy 35 (2010) 49704980

4977

Fig. 3 Simulated hydrogen production rates for Cases 1 and 2.

efciency of the gasier is approximately 72% while the corresponding exergy efciency is 66% based on Eqs. (18) and (19). Normally the energy efciency at these conditions may be expected to be around 80%; the value here is lower since the system assessed has unconverted solid carbon as char and catalysis is not used to promote the gasier reactions. In addition, the fuels are over-oxidized in the gasier in order to attain the required gasication temperature [41], and this process may reduce the gasier efciency. The gasier exergy efciency is lower than the energy efciency, mainly due to chemical reactions and oxidization. Both combustion reactors operate with high energy and low exergy efciencies. The latter are associated primarily with internal irreversibilities. For the steam-methane reformer, the energy efciency is found to be 83% and the exergy efciency 77%. These values are consistent with those reported in the literature [34,42]. Note that in the HTS and LTS units, the shift reactions occur but there is no combustion. Therefore internal exergy

destructions are very low, leading to high exergy efciencies for these devices. It is observed that signicant heat is transferred to water to produce steam in heat exchangers, boilers and economizers. The results suggest that the low-pressure indirectly heated gasier requires improvements in terms of energy recovery. For this system, with its feed rate of 4000 kg per day of wet biomass to the gasication process, the hydrogen production energy efciency is 24% (see Table 3). It is determined with the simulation that 132.7 kg hydrogen can be produced from 4000 kg biomass with an energy rate of 4.5 MW. For the second case, also with a feed rate of 4000 kg per day of wet biomass to the gasication process, the hydrogen production energy efciency is improved to 27% (see Table 3). Also, it can be found that 165 kg hydrogen is produced from 4000 kg biomass with an equivalent thermal input 5.6 MW. It can be seen in Fig. 5 that the reactors with the highest exergy destruction rates based on Eq. (21) are the gasier

Fig. 4 Energy and exergy efciencies for main system components for Case 1.

4978

international journal of hydrogen energy 35 (2010) 49704980

Fig. 5 Exergy destruction ratios for main components for Case 1.

(R-GIBBS), in which 34% of the total exergy inlet is destroyed. This observation implies that the gasier is an important component for efciency system improvement, especially since the biomass can be gas, solid and liquid. The combustion reactors are also responsible for large exergy destructions, mainly due to irreversibilities associated with the combustion reactions. These exergy losses mainly relate to chemical exergy destructions, and for both combustion units 1 and 2 are around 11%. It is also interesting that the dry reactor has a high exergy destruction rate, which is approximately 4.5% due to the 50% moisture content of the inlet biomass. Thus, the heat demand is high for this process, resulting in high exergy destruction rates. Exergy destruction ratios (on a percentage basis) for Case 1 are shown in Fig. 5 for the system components and in Fig. 6 for components with lower exergy losses. The exergy destruction ratio is the ratio of the exergy destruction for a component to the overall system exergy destruction. The exergy destruction ratio in Fig. 5 shows that the HE2 and WHB units are each responsible for about 1% of the exergy destruction rate, and the SMR, HTS and LTS reactors, which do not include combustion, also have very low exergy destruction rates. There are two cases of note in this assessment. The steammethane reforming (SMR) heat demand is supplied by

fractioned syngas in one case, and by external methane gas (CH4) supply in the other. Assessments are required to understand better how hydrogen production is affected by changing this parameter, with the aim of investigating the feasibility of producing hydrogen from biomass and better understanding the potential of biomass as a renewable energy source. As pointed out earlier, detailed energy and exergy assessments were performed for just one system (Case 1). For Case 2, the improvement of the system performance was gauged by altering one parameter and observing the impact on hydrogen production and system overall energy and exergy efciencies. The energy and exergy efciencies and exergy destruction ratios for the main system components for Case 2 are similar to those for Case 1, so performance gures are not presented for Case 2. Also, the exergy destruction ratios for auxiliary components for Cases 1 and 2 are similar. However, improvements are observed for Cases 1 and 2 in hydrogen production rate and overall system energy and exergy efciencies. On a broader scale, the results of this study support the contention by many that biomass may contribute to a future hydrogen economy. Although biomass has the advantage of being renewable if managed properly, it has challenges; large quantities of biomass need to be grown and transported to

Fig. 6 Exergy destruction ratio for auxiliary components for Case 1.

international journal of hydrogen energy 35 (2010) 49704980

4979

produce a small amount of hydrogen. Transportation concerns may be alleviated by using pyrolysis of biomass to produce bio-oil, as opposed to direct gasication. Based on the costs and availability of hydrogen production processes, it is likely that hydrogen will be produced by steam-methane reforming or coal gasication during a transition to a hydrogen economy. Future advances in water-splitting processes may allow them to replace fossil fuel processes as cleaner, long-term energy solutions. Many predictions of how a hydrogen economy will unfold have been published. For instance, a roadmap was created that provides an overview of a possible evolution of hydrogen production technologies in the future [43]. The timing of each step in this evolution towards a hydrogen economy depends on how quickly technology advances and other factors.

STBR T T0 _ W x

steambiomass ratio temperature, K reference-environment temperature, K work rate, kJ/h exergy ratio

Greek symbols j exergy efciency, % h energy efciency, % b correlation factor, % Subscripts bio biomass biomoist biomass moisture cg cold gas dest destroyed drybio dry biomass en energy gen generated i,j index for components in input meth methane gas (CH4) out output st steam sys system turb turbine prodg produced gas unconcarb unconverted carbon Supercripts ch chemical ph physical Acronyms COMB combustion COMP compressor COOL cooling EC economizer HE heat exchanger HTS high-temperature shift LTS low-temperature shift PSA pressure swing adsorption WHB waste heat boiler

5.

Conclusions

The energy and exergy analyses performed of biomass-based hydrogen production have yielded energy and exergy efciencies and an understanding of the impact on performance of several parameters. The feasibility of producing hydrogen from biomass and a better understanding the potential of biomass as a renewable energy source have been attained by considering two methods: 1) the heat required for steammethane reforming is supplied by fractioned syngas, and 2) the SMR-COMB reactor is provided with externally supplied methane gas. Oil palm shell is the biomass considered. For the direct gasication reaction, a BCL-type low-temperature indirectly heated steam gasier is examined. The thermodynamic assessments for the two cases demonstrate that the processes have low efciencies. The simulation conrms for the system that the second case considered, which indicates performance improvements, has higher energy and exergy efciencies than the rst case.

Acknowledgments
The authors acknowledge the support provided by the Ontario Research Excellence Fund and the Natural Sciences and Engineering Research Council of Canada.

references

Nomenclature
[1] Kruse A, Gawlik A. Biomass conversion in water at 330 410  C and 3050 MPa: identication of key compounds for indication different chemical reaction pathways. Industrial Engineering and Chemistry Research 2003;42(2):26779. [2] Ni M, Leung DYC, Leung MKH, Sumathy K. An overview of hydrogen production from biomass. Fuel Processing Technology 2006;87:46172. [3] Ntaikou I, Gavala HN, Kornaros M, Lyberatos G. Hydrogen production from sugars and sweet sorghum biomass using Ruminococcus albus. International Journal of Hydrogen Energy 2008;33:115363. [4] Dincer I, Rosen MA. Exergy: energy, environment and sustainable development. Oxford, UK: Elsevier; 2007. [5] Kelly-Yong TL, Lee KT, Mohamed AR, Bhatia S. Potential of hydrogen from oil palm biomass as a source of renewable energy worldwide. Energy Policy 2007;35:5692701.

Cp _ E _ Ex ex h LHV mi mo Po Q R S

specic heat, kJ/kg K energy ow rate, kJ/h exergy ow rate, kJ/h specic exergy, kJ/kg specic enthalpy, kJ/kg lower heating value, MJ/kg inlet mass, kg outlet mass, kg reference-environment pressure, kPa heat, kJ universal gas constant, kJ/kmol K entropy, kJ/K

4980

international journal of hydrogen energy 35 (2010) 49704980

[6] Midilli A, Dincer I. Key strategies of hydrogen energy systems for sustainability. International Journal of Hydrogen Energy 2007;32(5):51124. [7] Milne TA, Elam CC, Evans RJ. Hydrogen from biomass: state of the art and research challenges. Report No IEA/H2/TR-02/ 001. Golden, CO, USA: National Renewable Energy Laboratory; 2002. [8] Luyben WL. Chemical reactor design and control. Hoboken, New Jersey: Wiley; 2007. [9] Mahishi MR, Goswami DY. Thermodynamic optimization of biomass gasier for hydrogen production. International Journal of Hydrogen Energy 2007;32:383140. [10] Mahishi MR, Sadrameli MS, Vijayaraghavan S, Goswami DY. A novel approach to enhance the hydrogen yield of biomass gasication using CO2 sorbent. Journal of Engineering for Gas Turbines and Power 2008;130. paper 0115011. [11] Rosen MA. Thermodynamic investigation of hydrogen production by steammethane reforming. International Journal of Hydrogen Energy 1999;16:20717. [12] Simpson AP, Lutz AE. Exergy analysis of hydrogen production via steam methane reforming. International Journal of Hydrogen Energy 2007;32:481120. [13] Toonssen R, Woudstra N, Verkooijen AHM. Exergy analysis of hydrogen production plants based on biomass gasication. International Journal of Hydrogen Energy 2008; 33:407482. [14] Lutz AE, Bradshaw RW, Keller JO, Witmer DE. Thermodynamic analysis of hydrogen production by steam reforming. International Journal of Hydrogen Energy 2003;28:15967. [15] Handbook of heterogeneous catalysis. Weinheim, PA: VCH Verlagsgesellschaft Mbh; 1997. [16] Kotas TJ. The exergy method of thermal plant analysis. Malabar, Florida: Kreiger; 1995. [17] Robinson PJ, Luyben WL. Simple dynamic gasier model that runs in Aspen dynamics. Industrial Engineering and Chemistry Research 2008;47:778492. [18] Nikoo MB, Mahinpey N. Simulation of biomass gasication in uidized bed reactor using ASPEN PLUS. Biomass and Bioenergy 2008;32:124554. [19] Lu Y, Guo L, Zhang X, Yan Q. Thermodynamic modeling and analysis of biomass gasication for hydrogen production in supercritical water. Chemical Engineering Journal 2007;131: 23344. [20] Lei Y, Feng X, Min S. Parameters optimization of hydrogen production from glucose gasied in supercritical water by equivalent cumulative exergy analysis. Applied Thermal Engineering 2007;27:232431. [21] Feng W, Van der Kooi HJ, Arons JS. Biomass conversions in subcritical and supercritical water: driving force, phase equilibria, and thermodynamic analysis. Chemical Engineering and Processing 2004;43:145967. [22] Penninger JML, Rep M. Reforming of aqueous wood pyrolysis condensate in supercritical water. International Journal of Hydrogen Energy 2006;31:1597606. [23] Penninger JML, Maass GJJ, Rep M. Compressed hydrogen-rich fuel gas (CHFG) from wet biomass by reforming in super critical water. International Journal of Hydrogen Energy 2007; 32:14726. [24] Antal Jr MJ, Allen SG, Schulman D, Xu X, Divilio RJ. Biomass gasication in supercritical water. Industrial Engineering and Chemistry Research 2000;39:404453.

[25] Simbeck DR. Hydrogen costs with CO2 capture. Presented at 7th international conference on greenhouse gas control technologies (GHGT-7). Vancouver, British Columbia, Canada; September 610, 2004. [26] Szargut J, Morris DR, Steward FR. Exergy analysis of thermal, chemical and metallurgical processes. New York: Hemisphere; 1988. [27] Edgerton RH. Available energy and environmental economics. Toronto: D.C. Heath; 1982. [28] Ptasinski KJ, Prins MJ, Pierik A. Exergetic evaluation of biomass gasication. Energy 2007;32:56874. [29] Szargut J. Exergy method: technical and ecological applications. Boston: WIT Press; 2005. [30] Berthiaume R, Bouchard C, Rosen MA. Exergetic evaluation of the renewability of a biofuel. Exergy International Journal 2001;1:25668. [31] Turpeinen E, Raudaskoski R, Pongracz E, Keiski RL. Thermodynamic analysis of conversion of alternative hydrocarbon-based feedstocks to hydrogen. International Journal of Hydrogen Energy 2003;33: 663543. [32] Moran MJ. Availability analysis: a guide to efcient energy use. New York: American Society of Mechanical Engineers; 1989. [33] Rosen MA, Scott DS. Comparative efciency assessments for a range of hydrogen production processes. International Journal of Hydrogen Energy 1998;23:6539. [34] Rosen MA, Scott DS. A thermodynamic investigation of the potential for cogeneration for fuel cells. International Journal of Hydrogen Energy 1988;13:77582. [35] Rosen MA. Thermodynamic comparison of coal-red and nuclear electrical generating stations. Transactions of the Canadian Society for Mechanical Engineering 2000;24(1B): 27383. [36] Higman C, Burgt VM. Gasication. Amsterdam: Elsevier; 2003. [37] Lover KA. Biomass gasication integration in recuperative gas turbine cycles and recuperative fuel cell integrated gas turbine cycles. MSc thesis, Department of Energy and Process Engineering, Norwegian University of Science and Technology, Norway; 2007. [38] Corradetti A, Desideri U. Should biomass be used for power generation or hydrogen production? ASME Journal of Engineering for Gas Turbines and Power 2007;129: 62934. [39] Li XT, Grace JR, Lim CJ, Watkinson AP, Chen HP, Kim JR. Biomass gasication in a circulating uidized bed. Biomass Bioenergy 2004;26:17193. [40] Spath P, Aden A, Eggeman T, Ringer M, Wallace B, Jechura J. Biomass to hydrogen production detailed design and economics utilizing the BCL indirectly heated gasier. Report No. TP-51037408. National Renewable Energy Laboratory; 2005. [41] Babu SP. Biomass gasication for hydrogen production process description and research needs. Des Plaines, Illinois: Gas Technology Institute; 2002. [42] Rosen MA. Evaluation of energy utilization efciency in Canada using energy and exergy analyses. Energy International Journal 1992;17(4):33950. [43] Miller GQ, Penner DW. Possibilities and challenges for a transition to a hydrogen economy. Chemical Engineering Transactions 2004;4:518.

Das könnte Ihnen auch gefallen