You are on page 1of 14

Chemical Engineering Science 142 (2016) 201–214

Contents lists available at ScienceDirect

Chemical Engineering Science


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

Thermodynamic analysis of a combined heat and power system


with CO2 utilization based on co-gasification of biomass and coal
Po-Chih Kuo, Wei Wu n
Department of Chemical Engineering, National Cheng Kung University, Tainan 70101, Taiwan

H I G H L I G H T S

 A CHP system based on co-gasification of biomass and coal is investigated.


 Identifying the CBP through the specific optimization algorithm.
 The performance of the CHP system can be effectively improved under CO2 addition.
 Co-gasification of torrefied biomass and coal blends is able to suppress the CO2 specific emissions.

art ic l e i nf o a b s t r a c t

Article history: In this article, a co-gasification system blending coal and biomass is examined in the Aspen Plus
Received 12 June 2015 environment in terms of energy conversion efficiency (ECE) and exergy efficiency (EE). Although the
Received in revised form percentage of raw wood (RW), torrefied wood (TW) and coal in steam co-gasification is one of the most
31 October 2015
important parameters that affect the gasification process, the addition of CO2 could effectively improve
Accepted 23 November 2015
Available online 11 December 2015
ECE and EE while the steam-to-carbon ratio (S/C) is adjusted at the carbon boundary point (CBP). The
optimum operating conditions such as S/C and CO2 supply ratio, are determined by solving a series of
Keywords: constrained optimization algorithms for maximizing ECE and EE. A combined heat and power (CHP)
Co-gasification system using the maximum waste heat recovery and Rankine cycle is illustrated to assess performance in
Torrefied biomass
terms of power generation and CO2 utilization. The results show that the total power generation by
CO2 utilization
feeding the TW-based fuel blend of 40 wt% TW and 60 wt% coal is increased by 8.43%, as compared to
Carbon boundary point
Combined heat and power that of the RW-based fuel blend of 40 wt% RW and 60 wt% coal. Compared with 100 wt% coal fuel, the
Optimization TW-based fuel can significantly reduce CO2 specific emission by 38.23%.
& 2015 Elsevier Ltd. All rights reserved.

1. Introduction important role in energy and electricity production (Goransson et


al., 2011; Woolcock and Brown, 2013). However, coal gasification
Gasification is a promising thermo-chemical conversion pro- combined-cycle power plants release a large amount of green-
cess for producing syngas (i.e. H2 þCO) (Goransson et al., 2011; house gas (GHG) emissions, mainly carbon dioxide (CO2), nitro-
Woolcock and Brown, 2013). Syngas has been widely applied to gen- (N) and sulfur- (S) based gases, into the atmosphere, thus
the synthesis of chemicals, such as methanol, dimethyl ether causing environmental pollution (Taba et al., 2012). It is thus
(DME), and methyl tert-butyl ether (MTBE), as well as some liquid necessary to reduce the GHG emissions associated with this
transportation fuels via the Fischer–Tropsch synthesis method technology.
Biomass is one of the most important renewable and sustain-
based on various H2/CO ratios (Goransson et al., 2011; Tijmensen
able energy sources to produce syngas and electricity, and thus a
et al., 2002; Woolcock and Brown, 2013). Moreover, both heat and
promising alternative to the use of fossil fuels (Abdollahi et al.,
power can be generated in an integrated gasification combined
2010; Taba et al., 2012). Biomass is an abundant resource that is
cycle (IGCC) power plant (Sadhukhan et al., 2010).
environmental friendly because of its low nitrogen and sulfur
Typically, syngas is converted from carbonaceous fuels, such as
contents. It is also seen as a carbon-neutral fuel, which produces
coal, biomass, plastic and waste, of which coal currently plays an essentially zero net CO2 emissions to the environment (Abdollahi
et al., 2010; Taba et al., 2013). However, when raw biomass is
n
Corresponding author. Tel.: þ 886 6 2757575x62689; fax: þ886 6 2344496. developed for bioenergy purposes the utilization efficiency is
E-mail address: weiwu@mail.ncku.edu.tw (W. Wu). relatively low, due to the disadvantageous intrinsic properties of

http://dx.doi.org/10.1016/j.ces.2015.11.030
0009-2509/& 2015 Elsevier Ltd. All rights reserved.
202 P.-C. Kuo, W. Wu / Chemical Engineering Science 142 (2016) 201–214

this material compared to those of coal. For example, the major power generation is an important and potential clean coal tech-
properties of raw biomass include (Chen and Kuo, 2011; Sarvar- nology (CCT) that is able to reduce emissions of CO2, the primary
amini et al., 2013; Taba et al., 2013; Thanapal et al., 2014): (1) high greenhouse gas that is responsible global warming. A review of the
moisture content, (2) low calorific value, (3) low energy density, past literature suggests that most of the related studies focus on
(4) high hydrogen-to-carbon (H/C) ratio and oxygen-to-carbon (O/ the type of oxygen-blown (Emun et al., 2010; Zhang et al., 2013),
C) ratio, (5) hygroscopic nature, and (6) a low bulk density, making air-blown (Klimantos et al., 2009; Giuffrida et al., 2013), and a
both transportation and storage costs higher. These properties mixture of steam/oxygen (Lee et al., 2014; Asif et al., 2015) IGCC
mean that raw biomass gasification has a lower conversion effi- systems, while relatively little research has been carried out using
ciency than that seen with coal gasification. In recent years various a mixture of CO2/O2 (Oki et al., 2011; Prabu, 2015). If CO2 is used as
studies have reported that using the combination of coal and a gasifying agent it can not only be converted to carbon monoxide
biomass as a feedstock, in a process known as co-gasification, is as via the Boudouard reaction in gasification, but also reduce the
promising technology for the production of syngas. This process level of CO2 emission into the atmosphere (Prabu, 2015). Fur-
has the advantages of being both economical and environmentally thermore, a number of studies have examined the use of steam
friendly (García et al., 2013; Taba et al., 2012, 2013), and a number and CO2 as a gasifying agent. Castaldi and Dooher (2007) adopted
of researchers have examining blending various types of biomass the Aspen Plus simulator to investigate coal steam gasification
with coal. For instance, Saw and Pang (2013) examined the use of with a CO2 recycle stream to the reformer. They found that the
lignite and wood as a feedstock at various blending mass ratios for addition of CO2 could lead to decreased energy consumption in the
co-gasification. They reported that the H2/CO ratio can be reformer, and thus this is a suitable technique for reducing the
enhanced during co-gasification, and that a H2/CO ratio of 2 (for energy requirements of such systems. Butterman and Castaldi
Fischer–Tropsch synthesis) can be obtained by blending 60% wood (2007) also studied biomass steam gasification by introducing CO2
with 40% lignite. Aigner et al. (2011) reported an increase in via thermogravimetric analysis-gas chromatography (TGA-GC),
hydrogen production and a decrease in CO production with a ris- and found that the existence of CO2 significantly promotes the
ing coal ratio. Furthermore, the concentrations of impurities such production of CO when the operating temperature was beyond
as NH3 and H2S in the gas product have also been found to 700 °C, while the production of H2 is reduced. Prabowo et al.
decrease during the co-gasification process. Howaniec and Smo- (2014) compared the gasification performance under the atmo-
linski (2013) investigated the influence of synergetic effects on the spheres of steam, CO2, and steam–CO2 mixtures. They found that
steam co-gasification process. They reported that the reactivity of the gasification thermal efficiency increased along with the CO2
fuel blends increased at a gasification temperature of 900 °C, due mixing ratio at high gasification temperatures. Jayaraman and
to the alkali content of the biomass, and especially the presence of Gokalp (2015) carried out char gasification to investigate the
potassium, which acts as a catalyst to aid in the process of char kinetic behaviors of gasification under a steam–CO2 environment
gasification. in the temperature range of 850–950 °C. They found that a higher
Compared to traditional approaches to the production of raw temperature reduced the time required for 50% char conversion.
biomass energy, the application of torrefied biomass is attracting Despite the potential advantages of using steam–CO2 mixtures
increasing attention in recent years, as it offers several advantages. as a gasifying agent, as detailed above, the process of co-
Torrefied biomass is produced by thermally pretreating the raw gasification under a steam–CO2 environment for a class of power
biomass. This process is known as torrefaction, which is a mild generation system has not yet been investigated. The present
pyrolysis process carried out in the temperature range of 200– study thus aims to design a combined heat and power (CHP)
300 °C under an inert or nitrogen atmosphere. Torrefaction system based on the steam co-gasification of biomass and coal
enhances the physical and chemical properties of raw biomass. with the addition of CO2. Particular emphasis is placed on the
Torrefied biomass has the following improved properties com- comparison of co-gasification performance between raw/torrefied
pared with the untreated material (Chen and Kuo, 2011; Sarvar- biomass and coal blends. Two important parameters of the steam-
amini et al., 2013; Thanapal et al., 2014; Kuo et al., 2014): (1) lower to-carbon ratio (S/C) and CO2 supply ratio (yCO2 ) are adjusted to
moisture content (hydrophobicity), (2) higher energy density, evaluate their influences on system performance. Finally, the
(3) greater ignitability, reactivity and grindability, (5) a lower optimum operating conditions are obtained by solving a series of
oxygen-to-carbon (O/C) ratio, and (6) lower costs of transportation constrained optimization algorithms
and storage. Torrefied biomass is thus seen as a more valuable and
economic fuel than raw biomass. In the recent literature Deng
et al. (2009) reported that a combination of torrefaction and co- 2. Process simulation
gasification is a promising process for the production of syngas.
Chen et al. (2011) gasified torrefied sawdust in a bench-scale Coal and two types of biomass, namely raw wood (RW) and
laminar entrained-flow gasifier, and found that the cold gas effi- torrefied wood (TW), are selected as the feedstock for gasification
ciency increased compared to that of raw sawdust, especially for and co-gasification in this study (Park et al., 2012). Their proper-
sawdust torrefied at 250 °C. Weiland et al. (2014) observed that ties, such as proximate analysis, elemental analysis, and higher
carbon conversion was improved by using torrefied wood residue heating values (HHV), are presented in Table 1. The simulation
as a feedstock in an entrained flow gasifier. Furthermore, and more model is established using a commercial heat and mass balance
recently, Dudyński et al. (2015) used torrefied woody biomass for analysis package, Aspen Plus V8.4. The following main assump-
industrial-scale gasification experiments, and found that syngas tions are made in this study for developing the process model:
with a higher calorific value was obtained by using torrefied pel- (1) the process is steady-state; (2) the feedstock is at normal
lets as a feedstock. conditions (i.e. 25 °C and 1 atm); (3) the solid and gaseous phases
These earlier studies show that torrefaction in association with are in a state of thermodynamic equilibrium; (4) the following
gasification is an effective and promising method to enhance the gaseous species are considered in the whole system: H2O, H2, O2,
performance of gasification. Although numerous works have N2, CO, CO2, CH4, H2S, COS, SO2,HCN, NO, NO2 and NH3; and
investigated the gasification characteristics of torrefied biomass, (5) char only contains carbon and ash, and tar formation is
little information is available on the co-utilization of coal and neglected (Kannan et al., 2013; Shen et al., 2008; Song et al., 2013).
torrefied biomass during a steam co-gasification process. On the In the simulation, the global thermodynamic property method
other hand, the integrated gasification combined cycle (IGCC) used in this model is the Peng-Robinson equation of state with the
P.-C. Kuo, W. Wu / Chemical Engineering Science 142 (2016) 201–214 203

Boston-Mathias alpha function (PR-BM). The fuels are defined as and ash. The product yield distributions in RYield are calculated by
an unconventional component, based on the experimental results a calculator block which is controlled by the FORTRAN statement
of elemental analysis and proximate analysis. Due to the non- in accordance with the component characteristics of the feedstock
conventional stream for the fuels, the HCOALGEN model, including (Table 1). The heat of the reaction from the decomposition is
a number of empirical correlations for heat of combustion, heat of simulated by a heat stream (Qdecomp).
formation, and heat capacity, is used in the simulation. The density For the gasification model, an RGibbs reactor is used to model
of nonconventional fuels is established by the DCOALIGT model reactions according to chemical equilibrium. In this block, the
(Ramzan et al., 2011). The simplified block diagram of the whole chemical and phase equilibrium calculations are carried out by
system is shown in Fig. 1, which consists of a gasification unit, gas minimizing the Gibbs free energy. The mathematical model of the
cooling unit, gas cleaning unit, gas burning unit, and power RGibbs reactor is described in a previous study (Kuo et al., 2014),
generation unit. and this reactor is operated at 900 °C (Howaniec and Smolinski,
2013; Jayaraman and Gokalp, 2015; Prabowo et al., 2014). The heat
2.1. Co-gasification system of the reaction from the endothermic gasification reactions is
simulated by a heat stream (Qgasifier).The detailed operating con-
Fig. 2a shows a schematic of the co-gasification process, which ditions of the co-gasification process are presented in Table 2, and
is composed of a number of blocks. The total mass flow rate of fuel the major chemical reactions that take place during the gasifica-
into the system is kept constant at 100 kg/h throughout the tion are listed below (Prabowo et al., 2014; Taba et al., 2012,, 2013):
simulations. When the “fuel” (stream 1) is fed into the system, the Water gas reaction
first stage is the drying of the fuel. An RStoic reactor is used to
C þ H2 O-CO þH2 ; ΔH0 ¼ þ 131:4kJmol  1 ð2Þ
simulate the drying process and the heat produced by the reaction
is simulated by a heat stream (Qdrying). The process of drying is Water gas shift reaction
simulated by the following chemical reaction (Chen et al., 2013):
CO þ H2 O2CO2 þ H2 ; ΔH0 ¼ 42kJmol  1 ð3Þ
Fuel-0:0555H2 O þ fueldry ð1Þ
Boudouard reaction
After fuel drying, a yield reactor is used to simulate the
decomposition of fuel. In the yield reactor, the unconventional C þ CO2 -2CO; ΔH0 ¼ þ172:6kJmol  1 ð4Þ
components of the fuel (stream 2) are decomposed into conven- Methanation reaction
tional constituents (stream 3), including N2, O2, H2, S, H2O, carbon
C þ 2H2 2CH4 ; ΔH0 ¼  75kJmol  1 ð5Þ
Table 1
Proximate and elemental analyses of the feedstock used in the simulation (Park et CH4 þ H2 O2CO þ 3H2 ; ΔH0 ¼  206kJmol  1 ð6Þ
al., 2012).
Regarding the gasifying mediums (stream 4), steam and CO2,
Feedstocks Raw wood Torrefied wood Coal are mixed and preheated by a heater (H1) before being fed into the
RGibbs reactor. Notably, a heater (H1) with a heat stream (Qgasifying
Proximate analysis (wt%)
Moisture 3.81 3.19 6.67 agents) is used to adjust the inlet temperature from 25 °C (stream 4)
Volatile matter (VM) 88.72 81.19 27.25 to 200 °C (stream 5). The outlet stream (stream 6) resulting from
Fixed carbon (FC) 7.42 15.51 54.5 the RGibbs reactor is divided into two streams, product gas
Ash 0.05 0.11 11.58
(stream 7) and char (stream 8) through an SSPLIT block. The
Elemental analysis (wt%)
C 46.73 52.22 74.12
product gas is cooled down by a cooler (C1), which reduces the
H 6.46 5.18 4.22 temperature to 150 °C (Emun et al., 2010). The steam and acid
N 0.41 0.55 1.91 gases produced from the co-gasification process are then removed
Oa 46.35 41.94 6.93 from the product gas in a separator block, where a heat stream
S 0 0 0.41
Higher heating value (MJ/kg) 18.51 20.56 26.82
(Qsep) is simulated to the energy demand of the separator. The
clean gas (stream 9) is then obtained and combusted with excess
a
By difference. air (stream 11) in order to provide complete combustion in the

Gasifying agents
Fuels (Steam/CO2)
(RW, TW, Coal)
Co-gasification Gas cooling and cleaning
Raw gas
Clean gas
Heat Recovery Steam
Power Generator (HRSG)
Gas Burning
and Rankine cycle
Flue gas

Fig. 1. Block diagrams for the CHP system.


204 P.-C. Kuo, W. Wu / Chemical Engineering Science 142 (2016) 201–214

Fig. 2. Process flow chart of (a) the co-gasification system, and (b) power generation system.

burner to form flue gas (stream 12), which can be used in the pump in which it is pressurized from 1 to 240 atm, resulting in
power generation system. To validate the present model, simula- energy consumption (WPump). The high pressure feedwater
tions have been performed for gasification of rubber wood in a (stream 14) is then fed into shell side of the multi-stream heat
fixed bed gasifier operated at atmospheric pressure (1 atm) exchangers (feedwater regenerative heaters) and this water gets
(Mahishi and Goswami, 2007) and the gasification temperature is heated by the exhaust steam entering in the tube side, so that the
900 °C. The validation has been examined in a previous study (Kuo steam is heated to 250 °C (stream 17) and sent to the economizer.
et al., 2014), where the predicted results are in good agreement The steam (stream 17) is first heated by the outlet stream of the
with the experimental values. hot flue gas to the temperature of 550 °C after leaving the boiler
(stream 19). To achieve the supercritical state, the temperature of
2.2. Power generation system the steam is then raised from 550 to 600 °C through superheater 1,
and it subsequently feeds into a high pressure turbine (HPT)
The schematic process of power generation is shown in Fig. 2 (stream 20) to expand and produce electricity (WHPT). The pres-
(b) where the power generation system major consists of a steam sure of the steam exhausted from HPT is 49 atm (stream 21). The
generator and Rankine cycle (Chen and Wu, 2014). After burning Fsplitter (F1) is then specified as 0.7 and used to split the steam
the product gas from the co-gasification process, the high tem- stream; that is, a 0.195 kg/s of steam stream (stream 22) flows to
perature flue gas is obtained and sent to the heat recovery steam superheater 2 to reheat the temperature to 620 °C (stream 23),
generator (HRSG) and steam turbine (ST) process. A series of heat while a 0.0837 kg/s of steam stream (stream 28) is sent to the
exchangers is used, including a radiant water-wall evaporator, feedwater regenerative heaters to reheat the feedwater. Similarly,
radiant superheater, and economizer, in which steam is generated after reheating to 620 °C (stream 23) in superheater 2, the steam is
in conjunction with the heat transfer from hot product gas and also delivered into an intermediate-pressure turbine (IPT) and
then sent to three steam turbines, namely a high pressure (HP), exhausted to a low-pressure turbine (LPT) (stream 25) to produce
intermediate pressure (IP), and low pressure (LP) one, for power electricity (WIPT and WLPT). After removing the heat from the flue
generation. The detailed operating conditions of the steam cycle gas, the temperature of exhausted gas drops to 50 °C (stream 38)
are shown in Table 2. First of all, 0.279 kg/s of the feedwater (Sorgenfrei and Tsatsaronis, 2014), and the exhaust steam (stream
(stream 13) with a temperature of 45 °C goes through a water 37) is then cooled down to 45 °C at 0.1 atm as saturated water
P.-C. Kuo, W. Wu / Chemical Engineering Science 142 (2016) 201–214 205

Table 2 The ECE is defined as the ratio of energy output of the co-
Operating parameters of co-gasification for power generation systems. gasification system to the energy input, whereas NEE is the ratio
of net power generation to energy input (Mahishi and Goswami,
Parameter Value
2007; Speidel and Worner, 2015). The energy input is calculated as
Feedstock CH1.68O0.752N0.008 (raw wood) the sum of the energy content of the fuel, heat of reaction, and
CH1.04O0.486N0.006 (torrefied wood) energy requirement of the preheating gasifying agents and
CH0.683O0.070N0.022S0.002 (Coal)
separator. The energy output is the lower heating value (LHV, kJ/
Fuel flow rate (kg/h) 100
Blending ratio (wt%) 20–80 Nm3) of the product gas for the co-gasification system (Lv et al.,
Gasifier Temperature (°C) 900 2004; Kuo et al., 2014) or net power output (Ẇnet , kW) for the
Pressure (atm) 1 power generation system. These are defined as follows:
Burner Air flow rate (kg/h) 1000
Sensitivity analysis Steam to carbon ratio (S/C) 0.1–2 F productgas U GP U LHVproductgas
CO2 content in gasifying 0.01–2
ECEð%Þ ¼ ̇
 100% ð10Þ
agents F fuel U LHVfuel þ Q H
Heat recovery steam generator Feedwater (kg/s) 0.279
(HRSG) and Rankine cycle HP steam turbine inlet 240 Q ̇H ¼ Q ḋ rying þ Q ̇decomp þ Q ̇gasifier þ Q ġ asifying agents þ Q ṡ ep ð11Þ
pressure (atm)
HP steam turbine inlet 600  
temperature (°C) LHVproduct gas ¼ 30:0xCO þ 25:7xH2 þ 85:4xCH4  4:2 ð12Þ
Fsplitter (F1) 0.7
Steam mass flow rate (kg/s) 0.195 (stream Ẇnet
NEEð%Þ ¼  100% ð13Þ
22) F fuel U LHVfuel þ Q ̇H
0.084 (stream
28)
Ẇnet ¼ ẆHPT þ ẆIPT þ ẆLPT  ẆPump ð14Þ
IP steam turbine inlet 49
pressure (atm)
where F productgas and F fuel are the mass flow rates of the product
IP steam turbine inlet tem- 350
perature (°C)
gas and fuel (kg/s), respectively; GP is the volume of product gas
Fsplitter (F2) 0.9 from the gasification per unit weight of fuel (Nm3/kg-fuel); Q ̇H is
Steam mass flow rate (kg/s) 0.186 (stream the heat required for the co-gasification process (kW); x stands for
25) the mole fraction of gas species in the product gas (dry basis);
0.009 (stream
Ẇnet is the net power of the whole system (kW); ẆPump is the
32)
LP steam turbine inlet 5 energy requirement of the water pump (kW), and ẆHPT , ẆIPT , and
pressure (atm) ẆLPT are the energy outputs (kW) from individual steam turbines,
Fsplitter (F3) 0.9 corresponding to the high-pressure turbine, intermediate-
Steam mass flow rate (kg/s) 0.176 (stream
pressure turbine, and low-pressure turbine, respectively.
27)
0.009 (stream
35) 2.5. Exergy analysis
Condenser pressure (atm) 0.1

In addition to the energy analysis, the exergy efficiency (EE) of


the co-gasification system is also evaluated. For a steady state
(stream 13) by the cooling water in an external water-cooled
condenser (C2). process, the overall exergy balance between the inlet and outlet
streams can be written as (Kaska, 2014; Prins et al., 2003; Zhang et
2.3. Process parameters al., 2012, 2013):
Ẋ Ẋ
Exi ¼ Exj ð15Þ
The steam-to-carbon (mass flow rate) ratio (S/C ratio) and CO2
in out
supply ratio (yCO2 ) are two significant parameters in this process.
They are expressed as Ẋ ̇ ̇ ̇
Exi ¼ Exfuel þ Exgasifying agents þ Exheat ð16Þ
F steam
S=C ¼ ð7Þ in
yc F fuel
Ẋ ̇ ̇ ̇ ̇
F CO2 Exj ¼ Exproduct gas þ Exchar þ Exw þExloss ð17Þ
yCO2 ¼ ð8Þ
F steam þ F CO2 out

Ṗ Ṗ
where yc is the carbon content in the fed fuel, and yCO2 is defined where Exi and Exj are the exergy rates of the input and output
by the weight ratio of the CO2 contained in the total gasifying in out
streams, respectively. Ėxfuel is the input exergy rate of fuel (kW),
agents from both steam and CO2. That is, when yCO2 is equal to 0, it
means that steam is the only gasifying agent fed to the gasifier. Ėxgasifyingagents is the exergy rate of the gasifying agents (kW), Ėxheat
In addition, four biomass blending ratios (BR) of raw/torrefied is the exergy rate of heat required by the gasification system (kW),
biomass and the coal are considered for the co-gasification pro- Ėxproduct gas is the output exergy rate of the product gas (kW), Ė
xchar is the exergy rate of char (kW), Ėxw is the exergy rate of work
cess, and these are 20, 40, 60, and 80 wt%, which are expressed as
(kW), and Ėxloss is exergy loss rate from the system (kW).
F biomass Therefore, the total exergy of the material streams includes
BRðwt%Þ ¼  100% ð9Þ
F biomass þ F coal physical exergy and chemical exergy, and these are defined in the
following equations:
2.4. Energy analysis
Ėxtotal ¼ Ėxph þ Ėxch ð18Þ

The energy conversion efficiency (ECE) and net energy effi- where Ėxtotal represents the total exergy of the material streams
ciency (NEE) are two important indexes with regard to the per- (kW), Ėxph is the physical exergy of the material streams (kW), and
formance of the system, and thus they are evaluated in this study. Ėxch is the chemical exergy of the material streams (kW).
206 P.-C. Kuo, W. Wu / Chemical Engineering Science 142 (2016) 201–214

Fig. 3. Distributions of (a) H2 flow rate, (b) CO flow rate, (c) CO2 flow rate in the product gas, and (d) char flow rate from steam gasification of three fuels.

In the foregoing equation, the Ėxph for each species in the For coal, the specific chemical exergy can be expressed by
product gas can be described by (Zhang et al., 2013)
̇  
Ėxph ¼ ðh  h0 Þ  T 0 ðs  s0 Þ ð19Þ Excoal ¼ Q L  1:0438 þ 0:0013H=C þ0:1083O=C þ 0:0549N=C
 
þ 6:710S O=C r 0:666 ð22Þ
where h and s are the specific enthalpy and entropy of the gas
species at a given state, while h0 and s0 are the specific enthalpy where Q L is the heating value of the coal (kJ/kg), and C, H, O, N, S
and entropy of the gas species at the environment state T 0 ¼ 25 °C are the mass fractions of carbon, hydrogen, oxygen, nitrogen, and
and P 0 ¼ 1 atm, respectively. sulfur, respectively.
The Ėxch for each species in the gas mixture can be evaluated as However, for biomass, the specific chemical exergy can be
follows: obtained by (Prins et al., 2003; Zhang et al., 2012)

̇ X  n
 Ėxbiomass ¼ βF biomass LHVbiomass ð23Þ
Exch ¼ ni Exch;i þ RT 0 ln P i ð20Þ
i
ni 1:044 þ 0:0160H=C 0:3493O=Cð1þ 0:0531H=CÞ þ 0:0493N=C
β¼ ðO=C r2Þ
1  0:4124O=C
where ni is the mole flow rate of species i in the product gas ð24Þ
(kmol/s), Exch;i is the standard chemical exergy of species i in the
where LHVbiomass is the heating value of the biomass (kJ/kg), and C,
product gas, and R is the universal gas constant (kJ kmol  1 K  1).
H, O, N are the mass fractions of carbon, hydrogen, oxygen, and
The exergy of heat streams is defined as follows (Kaska, 2014):
nitrogen, respectively.
X T0

As a result, the exergy efficiency (EE) of the product gas in the co-
Ėxheat ¼ 1 QH ð21Þ
T gasification system is defined by (Prins et al., 2003; Zhang et al., 2012)
Ėxproduct gas
where T is the operating temperature for the system, and Q H is the EEð%Þ ¼  100% ð25Þ
Ėxfuel þ Ėxgasifying agents þ Ėxheat
heat requirement of the system (kW).
P.-C. Kuo, W. Wu / Chemical Engineering Science 142 (2016) 201–214 207

Moreover, the overall exergy efficiency (OEE) of the whole


system is defined as the ratio of net power output of the power
generation system to the total exergy input, as follows (Kaska,
2014):
Ẇnet
OEEð%Þ ¼  100% ð26Þ
Ėxfuel þ Ėxgasifying agents þ Ėxheat

3. Results and discussion

In the following discussion, the steam gasification character-


istics (i.e. yCO2 ¼ 0) of RW and TW are first studied, and coal is also
examined for comparison purposes. Subsequently, the effects of
the CO2 supply ratio (yCO2 ) on the steam gasification characteristics
are examined, and the performance from various operating BRs is
also investigated. Finally, a power generation system is examined
to observe the influences that the fuel has on its electricity pro-
duction and the specific CO2 emissions.

3.1. Gasification characteristics and optimization

The effects of S/C ratios on product characteristics during


gasification using three fuels are shown in Fig. 3. It can be seen
that the S/C ratio has a significant impact on H2 flow rate (Fig. 3a)
and CO flow rate (Fig. 3b), no matter what fuels are examined.
When increasing the S/C ratio from 0.1 to 2.0, the increases in the
H2 flow rate are 1.96, 3.36, and 6.00 kmol/h, corresponding to RW,
TW, and coal, respectively. This is attributed to the fact that an
increase in the S/C ratio facilitates H2 production, because of the
water gas reaction (Eq. (2)) and water gas shift reaction ((Eq. (3)).
Notably, the amount of H2 formation from TW steam gasifica-
tion is higher than that with RW when the S/C ratio is greater than
0.5. This trend can be explained by the proprieties of the fuels, in
which carbon is the major element, at 46.73 wt% for RW,
52.22 wt% for TW, and 74.12 wt% for coal. Park et al. (2012) indi-
cated that the atomic H/C and O/C ratios were close to those in coal
when RW undergoes torrefaction, implying that torrefaction could
make the fuel characteristics closer to those of coal. Therefore,
since the carbon content in TW is higher than that in RW, this
implies that more steam needs to be supplied when the TW is
gasified, resulting in more H2 being formed. In contrast, CO for-
Fig. 4. Distributions of (a) syngas yield, and (b) lower heating value from the steam
mation shows an increasing trend at first, but this is then followed gasification of three fuels.
by a significant decreasing trend. This is because of a series of
reactions, including both water gas reaction (Eq. (2)) and water gas first increase substantially and then become almost constant with
shift reaction ((Eq. (3)). At first the process is dominated by the further increases in the S/C ratio, meaning that the addition of
water gas reaction (Eq. (2)) when a low S/C ratio is used. However, more steam is almost independent of the syngas yield. Specifically,
once the S/C ratio is further increased, the water gas shift reaction the syngas yield is improved in the range of 6.27–20.76% for the
((Eq. (3)) plays a more significant role because the carbon content steam gasification of TW, as compared to RW. The maximum
in the fuel is completely reacted with the steam at a certain S/C
syngas yields are 1.90 Nm3/kg for RW, 2.30 Nm3/kg for TW, and
ratio (Fig. 3d), at which point the amount of CO is gradually con-
3.22 Nm3/kg of coal. Fig. 4b shows that the LHV of the product gas
sumed by the steam and more H2 and CO2 are produced (Fig. 3a
reaches the optimal values and then goes down significantly. This
and d). Notably, the CO2 production from RW and TW is higher
is because the LHV is contributed by the heating values of the
than that from coal due to the higher oxygen content in biomass
product gas, as expressed in Eq. (12), and thus the first increase in
(Table 1). Fig. 3d shows that more unconverted char residue is
observed from TW and coal, implying that more steam needs to be LHV is due to the increase in syngas yield. However, the excess
supplied in order to achieve a high level of carbon conversion. On steam supplied to the gasification system then lowers the LHV.
the other hand, it is notable that there is a carbon boundary point This trend is attributed to the lower CO production in the product
(CBP) where the all carbon in the fuel is totally reacted by the gas (Fig. 3b) while the syngas formation remains constant, as
gasifying agents (Prins et al., 2003; Renganathan et al., 2012), observed in Fig. 4a.
regardless of what the fuel is. A specific optimization algorithm is The effects of the S/C ratio on the energy conversion efficiency
used to obtain the CBP location for all three fuels at certain (ECE) and exergy efficiency (EE), as expressed in Eqs. (10) and (25),
operating conditions, as explained below. are presented in Fig. 5a and b, respectively. The distributions of
Regarding the syngas yield and LHV of the product gas, the both ECE and EE show an initially growing trend until they reach a
distributions at various S/C ratios are shown in Fig. 4a and b. It is maximum, and then both decrease with an increase in the S/C
apparent that the distributions of syngas yield for the three fuels ratio. It can be seen that there are optimal values for ECE and EE
208 P.-C. Kuo, W. Wu / Chemical Engineering Science 142 (2016) 201–214

for all three fuels. The maximum LHV of the product gas should be optimization algorithms, as follows,
determined in order to find the maximum efficiency of the steam 
 F out
 syngas
gasification process (Fig. 4b), which implies that the maximum max J i ¼ F out
syngas i ¼ ; i ¼ I; II; III ð27Þ
uij  F in
syngas formation should be first optimized (Fig. 2). According to i

the operating parameters in Table 2, two operating variables are subject to


selected (F in
i ; F steam ) to meet this aim by solving the constrained aj r uij r bj j ¼ 1; 2 ð28Þ

Δ F steam
0:1 r S=C ¼ r2 ð29Þ
yc F in
i

where F out
syngas j i represents the objective (Ji) of different fuels, I; II; III
are three kinds of fuel, corresponding to RW, TW, and coal, and
h iT
uij ¼ F steam ; F in
i represents the steady-state operating conditions.
j
To determine the optimal steady-state operating conditions, the
lower and upper bounds of uij , ai and bi , are given in Table 3. As a
result, the optimization of the maximum syngas formation of the
steam gasification process subject to constraint equations is
described by FORTRAN code and carried out in the Aspen Plus
environment based on the sequential quadratic programming
(SQP) method.
The optimal solutions are achieved through the specific opti-
mization algorithm and the optimal values of RW, TW, and coal are
obtained at S/C ratios of 0.4, 0.8, and 1.35, respectively. Notably, as
mentioned earlier, these locations with respect to the S/C ratio are
so-called carbon boundary points (CBPs), where no more carbon is
observed with a further increase in the S/C ratio (Fig. 2d).
Similarly, the maximum ECE and EE are obtained by solving the
following optimization algorithms

max J i ¼ ECEi ; i ¼ I; II; III ð30Þ
uij


max J i ¼ EEi ; i ¼ I; II; III ð31Þ
uij

subject to
aj r uij r bj j ¼ 1; 2 ð32Þ

Δ F steam
0:1 r S=C ¼ r2 ð33Þ
yc F in
i

Notably, it is found that these optimal values are also obtained


at the CBP. Similar results are also reported in the study of Prins
et al. (2003), which concluded that the maximum values of energy
and exergy efficiencies were observed at the CBP for steam
gasification.
In summary, Fig. 6 shows a comparison between gasification of
both RW and TW under the conditions of CBP, where the

Fig. 5. Distributions of (a) energy conversion efficiency, and (b) exergy efficiency
from the steam gasification of three fuels.

Table 3
Bounds of manipulated variables.

ui;j aj bj

 
ui;1 ¼ F steam kg=h 0 150
in  
10 100
ui;2 ¼ yc  F i kg=h
 
ui;3 ¼ F CO2 kg=h 0 10
 
uIV;1 ¼ F steam kg=h 0 150
 
uIV;2 ¼ yc  F in 0 100
coal kg=h
 
in
uIV;3 ¼ yc  F biomassl kg=h 0 100
 
uV;5 ¼ F steam kg=h 0 130
 
uV;6 ¼ F CO2 kg=h 0 20 Fig. 6. Profiles of syngas flow rate, energy conversion efficiency, and exergy effi-
ciency from the steam gasification of three fuels at the carbon boundary point.
P.-C. Kuo, W. Wu / Chemical Engineering Science 142 (2016) 201–214 209

Fig. 7. Distributions of the flow rates of H2, CO, and CO2 in the product gas from the
steam–CO2 gasification of (a) raw wood, and (b) torrefied wood at various CO2
supply ratios.

maximum syngas formations of RW, TW, and coal are 7.70, 9.33,
and 13.10 kmol/h (Fig. 3), respectively, revealing that the process
of syngas formation is enhanced by a factor of 21.17% when the TW
is gasified. Furthermore, both the ECE and EE can be improved by
up to 4.04% and 5.43%, respectively. It is thus verified that steam
gasification using TW to replace RW as a fuel can not only facilitate
syngas production, but also improve the energy and exergy
efficiencies.

3.2. Gasification with CO2 utilization and optimization

As mentioned previously, the optimal solutions with respect to


the S/C ratio are achieved by identifying the CBPs for the three
fuels. Next, the gasification characteristics under the impact of CO2
addition are investigated when the S/C ratio is fixed at CBP. A
mixture of steam and CO2 as the gasifying agent (Eq. (8)) is
examined, and the effects of the CO2 supply ratio on syngas for- Fig. 8. Distributions of the (a) lower heating value, (b) energy conversion efficiency,
mation under the condition of CBP are plotted in Fig. 7a for RW and (c) exergy efficiency from the steam–CO2 gasification at various CO2 supply
ratios.
and Fig. 7b for TW, respectively. In the case of RW, the trend of the
syngas flow rate changes slightly, meaning that both the H2 and
CO flow rates are insensitive to the CO2 supply ratio. In contrast to goes down. In addition, the effects of the CO2 supply ratio on CO2
RW, the H2 flow rate from TW is sensitive to the CO2 supply ratio; emissions show that a decrease in CO2 emissions with an increase
in particular, the CO flow rate reaches an optimal value and then in the CO2 supply ratio. These observed characteristics are
210 P.-C. Kuo, W. Wu / Chemical Engineering Science 142 (2016) 201–214

attributed to the fact that the addition of CO2 enhances CO pro-


14
duction because the Boudouard reaction (Eq. (4)) becomes more
dominant. Similar results have been observed in Butterman and
20%
Castaldi (2007) and Prabowo et al. (2014). 40%
12
The distributions of LHV with CO2 utilization are shown in 60%

Syngas flow rate (kmol/h)


80%
Fig. 8a. The addition of CO2 lowers the H2 flow rate in the product
gas and increases the CO flow rate to an optimal value (Fig. 7). The
10
former decreases the LHV of the product gas but the latter
increases it. Meanwhile, by examining the LHV equation ((Eq.
(12)), the heating value of CO is higher than that of H2. When the
8
two factors are considered together, the LHV of the product gas is
thus improved by increasing the CO2 supply ratio, especially for
TW. This is the reason why more CO production promotes the LHV
6
of the product gas during steam–CO2 gasification. The values of
LHV are enhanced by approximately 0.39% and 0.88%, corre-
sponding to RW and TW, respectively. The effects of CO2 addition
4
on ECE and EE are shown in Fig. 8b and c, respectively. Upon 0.5 1 1.5 2
inspection of the above observations in Fig. 8a, it is evident that S/C
the ECE and EE can be improved by the addition of CO2, and it can
be seen that optimal values can be determined for both RW and 14 90
TW. To obtain the maximum ECE and EE of the steam–CO2 gasi- Syngas
fication process, three operating variables are selected ECE
13 EE
(F in
i ; F steam ; F CO2 ) to solve the constrained optimization algorithms,
88

Syngas flow rate (kmol/h)


as follows:
12
max J i ¼ ECEj i ; i ¼ I; II ð34Þ

Efficiency (%)
uij 86

 11
max J i ¼ EEi ; i ¼ I; II ð35Þ
uij
84
10
subject to
aj r uij r bj j ¼ 1; 2; 3 ð36Þ 82
9

Δ F steam
0:1 rS=C ¼ r2 ð37Þ
yc F in
i 8 80
0 20 40 60 80 100
Δ F CO2 BR (%)
0:01 r yCO2 ¼ r 0:2 ð38Þ
F steam þ F CO2
Fig. 9. Distributions of the (a) syngas flow rate versus S/C ratio, and (b) optimal
h iT values of energy conversion efficiency and exergy efficiency from the steam co-
where uij ¼ F CO2 ; F steam ; F in
i represents the adjustable variables. gasification of blending torrefied wood with coal at various BRs.
j
Similarly, the upper and lower bounds of uij are shown in Table 3.
Therefore, the maximum values of ECE (Fig. 8b) and EE (Fig. 8c) values of efficiencies under the condition of various BRs, the fol-
from the steam–CO2 gasification of RW and TW are located at the lowing constrained optimization algorithms are solved by adjust-
same place of yCO2 ¼ 0:065, where their values are 83.35 and ing three operating variables (F in in
coal ; F biomass ; F steam ),
87.74% for ECE, while 77.07 and 82.60% for EE, respectively. It is 
max J i ¼ ECEi ; i ¼ IV ð39Þ
thus verified that steam gasification with CO2 utilization further uij
promotes the system performance, especially for the case of TW,

revealing that 0.64% and 0.18% increase in ECE and EE are max J i ¼ EEi ; i ¼ IV ð40Þ
achieved, respectively. uij

subject to
3.3. Co-gasification characteristics and optimization
aj r uij r bj j ¼ 1; 2; 3 ð41Þ
From the above observations, it is evident that the gasification
Δ F steam
performance is effectively enhanced by using TW as a fuel in a 0:1 r S=C ¼ r2 ð42Þ
steam–CO2 environment. On account of the advantages of TW, the yc F in
i

influence of the blending ratio (BR), as defined in Eq. (9), on the


co-gasification of TW and coal is examined. Fig. 9a presents the F in in in
i ¼ F biomass þ F coal ¼ 100 ð43Þ
profiles of the syngas formation for four BRs with respect to the S/
in
C ratio. These show that an increase in BR results in decreasing Δ F biomass
0 r BR ¼ r1 ð44Þ
syngas formation, because there is less carbon content in the fuel F in
coal
to be converted. On the other hand, at a constant BR the syngas h iT
formation increases linearly along with the S/C ratio, and then it where IV is the four kinds of BRs, anduij ¼ F in in
coal ; F biomass ; F steam
j
remains almost constant, no matter what the BR is. Similar to the represents the steady-state operating condition, and the lower and
observations in Fig. 4a, Fig. 9a suggests that an optimal S/C ratio upper bounds of uIV;j , ai and bi , are shown in Table 3.
corresponding to the CBP for different BRs exists during the steam The optimization results for different BRs of syngas formation,
co-gasification process. Consequently, in order to obtain maximum ECE, and EE are shown in Fig. 9b, and it can be seen that the syngas
P.-C. Kuo, W. Wu / Chemical Engineering Science 142 (2016) 201–214 211

Fig.11. Distributions of lower heating value, energy conversion efficiency, and


Fig. 10. Distributions of power output versus BR. exergy efficiency from the steam–CO2 gasification of blending torrefied wood with
coal at various CO2 supply ratios.

formation is lower when the BR is higher. For example, the


following constrained optimization algorithms are solved,
amount of syngas formed at BRs of 20, 40, 60, and 80 wt% under

the conditions of CBP are 11.91, 11.60, 10.55, and 10.13 kmol/h, max J i ¼ ECEi ; i ¼ V ð45Þ
uij
respectively. Meanwhile, the value of ECE at BR ¼20% is the lowest
among four BRs, because the input energy is raised when blending 
with 80 wt% coal. It is also worth noting that there is a decrease in max J i ¼ EEi ; i ¼ V ð46Þ
uij
EE when the BR is larger than 40%, due to the significant reduction
in syngas formation. It is thus concluded that BR should be con- subject to
trolled at 40% from the thermodynamic point of view.
aj ruij r bj j ¼ 5; 6 ð47Þ

3.4. Power generation in


Δ F biomass
BR ¼ ¼ 0:4 ð48Þ
F in
With regard to the power generation system, the distributions coal

of power output for different BRs at CBP are examined in Fig. 10.
Δ F steam
When coal (i.e. BR ¼0 wt%) and biomass (i.e. BR ¼100 wt%) are 0:1 r S=C ¼ r2 ð49Þ
used without blending in power generation system, the power yc F in
i

output for RW, TW, and coal are 278.70, 337.53, and 475.76 kW,
Δ F CO2
respectively. This suggests that using TW as an alternative fuel to 0:01 r yCO2 ¼ r 0:2 ð50Þ
F steam þ F CO2
RW can significantly increase the power output by a factor of
 T
21.11%. For the cases of different BRs, the results indicate that the where uij ¼ F steam ; F CO2 j represents the adjustable variables. The
power output of TW blending is higher than that of RW blending, upper and lower bounds of uV;j are presented in Table 3. The
no matter which BR is examined. From the viewpoint both of optimal values of ECE and EE are 86.24 and 84.81, respectively,
power generation and GHG emissions, if the TW-based fuel at when the CO2 supply ratio is operated at 0.08, and these produce
BR ¼20 wt% is used in the co-gasification power generation system 1.04 and 0.34% increases in ECE and in EE, respectively. Notably,
then the power output is only improved by a small extent com- when comparing the increases in both of ECE and EE, the impact of
pared with RW-based fuel, and this BR causes more GHG emis- CO2 utilization on co-gasification of TW blended with coal is more
sions due to the 80 wt% coal contained in the fuel. However, once significant than those seen with gasification of TW alone (Fig. 8b).
the BR is 40 wt%, the power output from the TW-based fuel is
enhanced by a factor of 8.43%. As a consequence, the condition of
BR ¼40 wt% for the TW-based fuel is the recommended operating 4. Net system efficiency and CO2 emissions
conditions for the co-gasification power generation system.
The net system efficiency (NEE) and overall exergy efficiency
3.5. Co-gasification with CO2 utilization and optimization (OEE) of the whole system are two key indicators for evaluating
the performance of the whole system, and these are defined in
The above observations show that the blending of 40 wt% TW Eqs. (13) and (26), respectively. In addition to these two indicators,
with 60 wt% coal is a better feedstock for a CHP system. For this the specific CO2 emissions when the biomass is assumed to be CO2
reason, the performance of co-gasification with CO2 utilization and neutral are also evaluated. The specific CO2 emissions are
the TW-based fuel at BR ¼40 wt% is carried out, with the results expressed by
shown in Fig. 11. With CO2 utilization, the LHV of the product gas,
F out
CO2
ECE, and EE are also significantly improved. Most importantly, the ECO2 ¼ ð51Þ
optimal conditions of the CO2 supply ratio should be determined W net
in this context. Similarly, the specific optimization algorithm is where ECO2 is the specific CO2 emissions (kg/kW h), and F out
CO2 is the
performed using the two operating variables (F steam ; F CO2 ), and the mass flow rate of CO2 emitted to the environment (kg/h).
212 P.-C. Kuo, W. Wu / Chemical Engineering Science 142 (2016) 201–214

Therefore, according to the optimum operating conditions, as and 0.99%, the CO2 specific emissions are reduced significantly by
mentioned earlier, the results for the blending of 40 wt% TW with 38.23%, meaning that CO2 specific emissions can be effectively
60 wt% coal and 100 wt% coal are compared with each other, as suppressed. Summarizing the observations in Figs. 11 and 12, it is
shown in Fig. 12. Although the values of NEE and OEE of the CHP verified that CO2 addition could effectively improve energy effi-
system using coal blended with TW as a fuel fall by about 3.55% ciency for co-gasification plants. From the viewpoint of environ-
mental pollution and the greenhouse effect, reusing CO2 into the
production of useful fuels is an attractive process. Moreover, co-
gasification of biomass and coal is also able to reduce CO2 emis-
sions, and can also reduce NOx and SOx levels in flue gases. For
these reasons, this technology is worth investigating thoroughly.

Table 4
Optimization of co-gasification power generation system from blending of 40% TW
with 60% coal.

Parameter Initial Optimized value Increasing factora


value (%)

S/C 1.2 1.22 –


CO2 supply ratio (yCO2) 0.08 0.103 –
Power generation (kW) 415.25 418.57 0.80
Net energy efficiency 38.51 40.98 6.41
(NEE) (%)
Overall exergy efficiency 37.97 38.27 0.79
Fig. 12. Profiles of net energy efficiency, overall exergy efficiency, and CO2 specific (OEE) (%)
emissions from the steam–CO2 gasification of blending torrefied wood with coal
under optimal operating conditions. a
Increasing factor ¼ OptimzedInitial
valueInitial value
value
 100.

Fig. 13. Process with heat integration under optimal operating conditions: (a) co-gasification system, and (b) power generation system.
P.-C. Kuo, W. Wu / Chemical Engineering Science 142 (2016) 201–214 213

4.1. Heat integration and optimization s specific entropy (kJ kmol  1 K  1)


S/C steam-to-carbon mass flow rate ratio
To further enhance the energy and exergy efficiencies from co- T temperature (°C)
gasification of the fuel blended with 40 wt% TW and 60 wt% coal, a TW torrefied biomass
heat-integrated design which achieves maximum waste heat W work (kW)
recovery is added to the system. Fig. 13 shows the optimized x mole fraction
process conditions with a heat-integrated design, in which the hot yCO2 CO2 supply ratio
and cold utilities are removed in the co-gasification system. yc mass fraction of carbon in the fuel (wt%)
However, the external heat demand of 300.70 kW is still required cp molar specific heat capacity at constant pressure
due to the endothermic reaction from the water gas reaction (Eq. (kJ kmol  1 K  1)
(2)) and Boudouard reaction (Eq. (4)). The optimized results are ΔH 0 Heat of reaction (at 298 K and 1 atm)
presented in Table 4, where the NEE and OEE are amplified by a
factor of 6.41% and 0.79%, respectively, based on the heat inte- Greek letters
gration design for the optimal process.
β correlation factor

5. Conclusions Superscripts

This study designed a combined heat and power (CHP) system, 0 standard reference state
and examined the performance of gasification and co-gasification
under a steam–CO2 atmosphere using two types of biomass (RW Subscript
and TW). The main results of this work are as follows:
ch chemical
1. Based on the steam gasification of three fuels, the optimal ph physical
values of ECE and EE occur at the ratios S/C of 0.4, 0.8, and 1.35 i, j species i, j
corresponding to RW, TW, and coal, respectively, where these
points are exactly located at the CBP.
2. The optimal values of ECE and EE can be effectively improved by
CO2 utilization in steam gasification, especially for TW gasifi- Acknowledgments
cation. Under the optimal CO2 supply ratio of 0.065, increases of
0.64% and 0.18% of ECE and EE are achieved, respectively. The authors would like to thank the Ministry of Science and
3. 40 wt% TWþ60 wt% coal is the recommended fuel for the CHP
Technology, Taiwan, for its partial financial support of this
system. With regard to process optimization, the simulation
research under Grant MOST 103-2221-E-006-251.
shows that CO2 specific emissions at the optimal operating
conditions are reduced by 38.23%, compared to those found
with coal. Furthermore, based on the heat integration design for
the optimal process, the NEE and OEE are increased by a factor References
of 6.41% and 0.79%, respectively.
Abdollahi, M., Yu, J., Hwang, H.T., Liu, P.K.T., Ciora, R., Sahimi, M., Tsotsis, T.T., 2010.
Process intensification in hydrogen production from biomass-derived syngas.
Ind. Eng. Chem. Res. 49, 10986–10993.
Nomenclature Aigner, I., Pfeifer, C., Hofbauer, H., 2011. Co-gasification of coal and wood in a dual
fluidized bed gasifier. Fuel 90, 2404–2412.
Asif, M., Bak, C.U., Saleem, M.W., Kim, W.S., 2015. Performance evaluation of inte-
BR biomass ratio (%) grated gasification combined cycle (IGCC) utilizing a blended solution of
ammonia and 2-amino-2-methyl-1-propanol (AMP) for CO2 capture. Fuel 160,
CBP carbon boundary point
513–524.
CHP combined heat and power Butterman, H.C., Castaldi, M.J., 2007. Influence of CO2 injection on biomass gasifi-
ECO2 the specific CO2 emissions (kg/kW h) cation. Ind. Eng. Chem. Res. 46, 8875–8886.
ECE energy conversion efficiency (%) Castaldi, M.J., Dooher, J.P., 2007. Investigation into a catalytically controlled reaction
gasifier (CCRG) for coal to hydrogen. Int. J. Hydrog. Energy 32, 4170–4179.
EE exergy efficiency (%) Chen, H.T., Wu, W., 2014. Efficiency enhancement of pressurized oxy-coal power
_
Ex exergy rates (kW) plant with heat integration. Int. J. Energy Res. 39, 256–264.
F mass flow rate (kg/s) Chen, Q., Zhou, J.S., Liu, B.J., Mei, Q.F., Luo, Z.Y., 2011. Influence of torrefaction
pretreatment on biomass gasification technology. Chin. Sci. Bull. 56,
GP the volume of product gas from the gasification of per 1449–1456.
unit weight of fuel (Nm3 kg-fuel  1) Chen, W.H., Kuo, P.C., 2011. Torrefaction and co-torrefaction characterization of
h specific enthalpy (kJ/kmol) hemicellulose, cellulose and lignin as well as torrefaction of some basic con-
stituents in biomass. Energy 36, 803–811.
HHV higher heating value (MJ kg-fuel  1) Chen, X., Zheng, D., Guo, J., Liu, J., Ji, P., 2013. Energy analysis for low-rank coal based
HPT high-pressure turbine process system to co-produce semicoke, syngas and light oil. Energy 52,
HRSG heat recovery steam generator 279–288.
Deng, J., Wang, G.J., Kuang, J.H., Zhang, Y.L., Luo, Y.H., 2009. Pretreatment of agri-
IPT intermediate-pressure turbine cultural residues for co-gasification via torrefaction. J. Anal. Appl. Pyrolsis 86,
LHVproductgas lower heating value of product gas (kJ Nm  3) 331–337.
LPT low-pressure turbine Dudyński, M., van Dyk, J.C., Kwiatkowski, K., Sosnowska, M., 2015. Biomass gasifi-
cation: influence of torrefaction on syngas production and tar formation. Fuel
NEE net energy efficiency (%)
Process. Technol. 131, 203–212.
OEE overall exergy efficiency (%) Emun, F., Gadalla, M., Majozi, T., Boer, D., 2010. Integrated gasification combined
QH heat requirement of the system (kW) cycle (IGCC) process simulation and optimization. Comput. Chem. Eng. 34,
R universal gas constant (kJ kmol  1 K  1) 331–338.
García, G., Arauzo, J., Gonzalo, A., Sánchez, J.L., Ábrego, J., 2013. Influence of feed-
RW raw wood stock composition in fluidised bed co-gasification of mixtures of lignite, bitu-
n number of moles minous coal and sewage sludge. Chem. Eng. J. 222, 345–352.
214 P.-C. Kuo, W. Wu / Chemical Engineering Science 142 (2016) 201–214

Giuffrida, A., Romano, M.C., Lozza, G., 2013. Efficiency enhancement in IGCC power Renganathan, T., Yadav, M.V., Pushpavanam, S., Voolapalli, R.K., Cho, Y.S., 2012. CO2
plants with air-blown gasification and hot gas clean-up. Energy 53, 221–229. utilization for gasification of carbonaceous feedstocks: a thermodynamic ana-
Goransson, K., Soderlind, U., He, J., Zhang, W., 2011. Review of syngas production via lysis. Chem. Eng. Sci. 83, 159–170.
biomass DFBGs. Renew. Sustain. Energy Rev. 15, 482–492. Sadhukhan, J., Zhao, Y., Shah, N., Brandon, N.P., 2010. Performance analysis of
Howaniec, N., Smolinski, A., 2013. Steam co-gasification of coal and biomass – integrated biomass gasification fuel cell (BGFC) and biomass gasification com-
synergy in reactivity of fuel blends chars. Int. J. Hydrog. Energy 38, bined cycle (BGCC) systems. Chem. Eng. Sci. 65, 1942–1954.
16152–16160. Sarvaramini, A., Assima, G.P., Larachi, F., 2013. Dry torrefaction of biomass – tor-
Jayaraman, K., Gokalp, I., 2015. Effect of char generation method on steam, CO2 and refied products and torrefaction kinetics using the distributed activation energy
blended mixture gasification of high ash Turkish coals. Fuel 153, 320–327. model. Chem. Eng. J. 229, 498–507.
Kannan, P., Shoaibi, A.A., Srinivasakannan, C., 2013. Energy recovery from co- Saw, W.L., Pang, S., 2013. Co-gasification of blended lignite and wood pellets in a
gasification of waste polyethylene and polyethylene terephthalate blends. 100 kW dualfluidised bed steam gasifier: the influence of lignite ratio on pro-
Comput. Fluids 88, 38–42. ducer gas composition and tar content. Fuel 112, 117–124.
Kaska, O., 2014. Energy and exergy analysis of an organic Rankine for power gen- Shen, L., Gao, Y., Xiao, J., 2008. Simulation of hydrogen production from biomass
eration from waste heat recovery in steel industry. Energy Convers. Manag. 77, gasification in interconnected fluidized beds. Biomass Bioenergy 32, 120–127.
108–117. Song, G., Chen, L., Xiao, J., Shen, L., 2013. Exergy evaluation of biomass steam
Klimantos, P., Koukouzas, N., Katsiadakis, A., Kakaras, E., 2009. Air-blown biomass gasification via interconnected fluidized beds. Int. J. Energy Res. 37, 1743–1751.
gasification combined cycles (BGCC): System analysis and economic assess- Sorgenfrei, M., Tsatsaronis, G., 2014. Design and evaluation of an IGCC power plant
ment. Energy 34, 708–714. using iron-based syngas chemical-looping (SCL) combustion. Appl. Energy 113,
Kuo, P.C., Wu, W., Chen, W.H., 2014. Gasification performances of raw and torrefied 1958–1964.
biomass in a downdraft fixed bed gasifier using thermodynamic analysis. Fuel Speidel, M., Worner, G.K.A., 2015. A new process concept for highly efficient con-
117, 1231–1241. version of sewage sludge by combined fermentation and gasification and
Lee, J.C., Lee, H.H., Joo, Y.J., Lee, C.H., Oh, M., 2014. Process simulation and ther- power generation in a hybrid system consisting of a SOFC and a gas turbine.
modynamic analysis of an IGCC (integrated gasification combined cycle) plant Energy Convers. Manag. 98, 259–267.
with an entrained coal gasifier. Energy 64, 58–68. Taba, L.E., Irfan, M.F., Daud, W.A.M.W., Chakrabarti, M.H., 2012. The effect of tem-
Lv, P.M., Xiong, Z.H., Chang, J., Wu, C.Z., Chen, Y., Zhu, J.X., 2004. An experimental perature on various parameters in coal, biomass and CO-gasification: a review.
study on biomass air-steam gasification in a fluidized bed. Bioresour. Technol. Renew. Sustain. Energy Rev. 16, 5584–5596.
95, 95–101. Taba, L.E., Irfan, M.F., Daud, W.M.A.W., Chakrabarti, M.H., 2013. Fuel blending effects
Mahishi, M.R., Goswami, D.Y., 2007. Thermodynamic optimization of biomass on the co-gasification of coal and biomass – a review. Biomass Bioenergy 57,
gasifier for hydrogen production. Int. J. Hydrog. Energy 32, 3831–3840. 249–263.
Oki, Y., Inumaru, J., Hara, S., Kobayashi, M., 2011. Development of oxy-fuel IGCC Thanapal, S.S., Chen, W., Annamalai, K., Carlin, N., Ansley, R.J., Ranjan, D., 2014.
system with CO2 recirculation for CO2 capture. Energy Procedia 4, 1066–1073. Carbon dioxide torrefaction of woody biomass. Energy Fuels 28, 1147–1157.
Park, S.W., Jang, C.H., Baek, K.R., Yang, J.K., 2012. Torrefaction and low-temperature Tijmensen, M.J.A., Faaij, A.P.C., Hamelinck, C.N., van Hardeveld, M.R.M., 2002.
carbonization of woody biomass: evaluation of fuel characteristics of the pro- Exploration of the possibilities for production of Fischer Tropsch liquids and
ducts. Energy 45, 676–685. power via biomass gasification. Biomass Bioenergy 23, 129–152.
Prabowo, B., Umeki, K., Yan, M., Nakamura, M.R., Castaldi, M.J., Yoshikawa, K., 2014. Weiland, F., Nordwaeger, M., Olofsson, I., Wiinikka, H., Nordin, A., 2014. Entrained
CO2–steam mixture for direct and indirect gasification of rice straw in a flow gasification of torrefied wood residues. Fuel Process. Technol. 125, 51–58.
downdraft gasifier: laboratory-scale experiments and performance prediction. Woolcock, P.J., Brown, R.C., 2013. A review of cleaning technologies for biomass-
Appl. Energy 113, 670–679. derived syngas. Biomass Bioenergy 52, 54–84.
Prabu, V., 2015. Integration of in-situ CO2-oxy coal gasification with advanced Zhang, G., Yang, Y., Jin, H., Xu, G., Zhang, K., 2013. Proposed combined-cycle power
power generating systems performing in a chemical looping approach of clean system based on oxygen-blown coal partial gasification. Appl. Energy 102,
combustion. Appl. Energy 140, 1–13. 735–745.
Prins, M.J., Ptasinski, K.J., Janssen, F.J.J.G., 2003. Thermodynamics of gas-char reac- Zhang, Y., Li, B., Li, H., Zhang, B., 2012. Exergy analysis of biomass utilization via
tions: first and second law analysis. Chem. Eng. Sci. 58, 1003–1011. steam gasification and partial oxidation. Thermochim. Acta 538, 21–28.
Ramzan, N., Ashraf, A., Naveed, S., Malik, A., 2011. Simulation of hybrid biomass
gasification using Aspen plus: a comparative performance analysis for food,
municipal solid and poultry waste. Biomass Bioenergy 35, 3962–3969.

You might also like