Vous êtes sur la page 1sur 12

Fuel 86 (2007) 244–255

www.fuelfirst.com

Some process fundamentals of biomass gasification in dual fluidized bed


Takahiro Murakami, Guangwen Xu *, Toshiyuki Suda, Yoshiaki Matsuzawa,
Hidehisa Tani, Toshiro Fujimori
Ishikawajima-Harima Heavy Industries Co., Ltd., Sin-nakahara-cho 1, Isogo-ku, Yokohama 235-8501, Japan

Received 18 December 2005; received in revised form 28 April 2006; accepted 24 May 2006
Available online 23 June 2006

Abstract

The dual fluidised bed gasification technology is prospective because it produces high caloric product gas free of N2 dilution
even when air is used to generate the gasification-required endothermic heat via in situ combustion. This study is devoted to providing
the necessary process fundamentals for development of a bubbling fluidized bed (BFB) biomass gasifier coupled to a pneumatic trans-
ported riser (PTR) char combustor. In a steam-blown fluidized bed of silica sand, gasification of 1.0 g biomass, a kind of dried coffee
grounds containing about 10 wt.% water, in batch format clarified first the characteristics of fuel pyrolysis (at 1073 K) under the
conditions simulating that prevailing in the gasifier intended to develop. The result shown that via pyrolysis more than 60% of fuel
carbon and up to 75% of fuel mass could be converted into product gas, while the simultaneously formed char was about 22% of fuel
mass. With all of these data as the known input, a process simulation using the software package ASPEN then revealed that the
considered dual bed gasification plant, i.e. a BFB gasifier + a PTR combustor, is able to sustain its independent heat and mass balances
to allow cold gas efficiencies higher than 75%, given that the fuel has suitable water contents and the heat carried with the product gas
from the gasifier and with the flue gas from the char combustor is efficiently recovered inside the plant. In a dual fluidized bed pilot
gasification facility simulating the gasification plant for development, the article finally demonstrated experimentally that the necessary
reaction time for fuel, i.e. the explicit residence time of fuel particles inside the BFB gasifier computed according to a plug granular flow
assumption, can be lower than 160 s. The results shown that varying the residence time from 160 to 1200 s only slightly increased the
gasification efficiency, but the reaction time available in the PTR, say, about 3 s in our case, was too short to assure the finish even
of fuel pyrolysis.
 2006 Elsevier Ltd. All rights reserved.

Keywords: Biomass; Coffee grounds; Dual fluidized bed gasification

1. Introduction tant not only to regulation of the nation’s CO2 emission


but also to her energy diversification.
In face of the global warming and climate change prob- Since 2003, The New Energy and Industrial Technology
lems, the utilization of biomass as an alternative of fossil Development Organization (NEDO), Japan, had spon-
fuels has been on its boom for many years. To Japan, bio- sored a three-year technical program to develop an
mass is an indispensable supplementary guard of the advanced upgrading and pyrolytic gasification system for
nation’s energy safe supply because the nation nearly com- high-moisture biomass fuels, such as various wet wastes
pletely relies on imported oils, coals and gases. Statistics from beverage and food industries. The technical system
anticipated that the usable biomass per year in Japan might consists of two parts, a reforming drier and a pyrolytic gas-
amount to 26 million kiloliter crude oils of heat equivalence ification plant. The drier works to reduce the water content
[1]. The efficient use of this energy resource is surely impor- of the wet biomass fuel to some reasonable values and to
upgrade the fuel through reforming some of its fats to
*
Corresponding author. Tel.: +81 45 759 2867; fax: +81 45 759 2210. improve the fuel’s thermal properties, such as O/C ratio,
E-mail address: gwxu@home.ipe.ac.cn (G. Xu). heating value and so on [2]. For the research program

0016-2361/$ - see front matter  2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.fuel.2006.05.025
T. Murakami et al. / Fuel 86 (2007) 244–255 245

the considered model biomass is coffee grounds containing unreacted char to air leads to char combustion and thereby
water higher than 65 wt.%. This model fuel not only typi- to heat up the heat carrier particles again. The heated par-
fies the high-moisture bio-wastes like tea grounds, soy ticles are in turn separated from the flue gas in the PTR
sauce lees, vinegar lees and bagasse which are concentrated cyclone and recirculated into the BFB gasifier to serve con-
already, but has also a total production of up to 30 thou- tinuously the fuel pyrolysis and gasification reactions by
sand tons per year in Japan. Instead of disposal or combus- providing them the needed endothermic heat.
tion for heat, many beverage works are thus interested with Therefore, the distinctive advantage of the gasification
conversion of such fuels into middle-caloric gases to lower plant is its separation of combustion exhaust from gasifica-
the works’ natural gas consumptions. tion gas product. This, similar to all the other dual bed
The devised pyrolytic gasification plant is schematically gasification processes (such as FICFB [3]), allows the
illustrated in Fig. 1. It is a bubbling fluidized bed (BFB) product gas to have high caloric values and low intake of
gasifier coupled to a pneumatic transported riser (PTR) inert gases (usually N2 and CO2). Furthermore, the quoted
char combustor. As shown in the plot, the lower end of gasification plant has a more compact design than the other
the PTR combustor is immersed into the particle bed (or commonly encountered dual bed gasification assembles. In
particle bulk) of the BFB gasifier, and the riser itself is the other cases the char combustor and fuel gasifier
seated around the central vertical line of the BFB gasifier. generally stand separately [3,4], which incurs surely more
Air nozzles extended to the downside entrance of the heat loss and requires a bigger space to construct the
PTR supply air to the combustor. This airflow not only plant.
provides the necessary oxidant for char combustion but Controlling both particle circulation and fuel mass and
catches also the particles from the BFB to let them move heat partitions between the BFB gasifier and PTR char
into the PTR. Steam, fed to the bottom of the BFB, is combustor is critical to the devised gasification plant and
the usually employed gasification reagent. The gasifier its implicated technology. While the circulation rate of heat
needs thus an external heat resource, which is the high- carrier particles determines if it is possible to carry the
temperature heat carrier particles, commonly sand at about required endothermic heat from the PTR combustor to
1173 K, circulated from the PTR. BFB gasifier, the partitions of fuel’s mass and heat specifies
The working procedure of the gasification plant is as fol- if suitable exothermic heat can be generated in the combus-
lows. Biomass fuel is fed to the BFB gasifier, wherein it is tor and how high the available gasification efficiency can
pyrolyzed and gasified through interactions with steam be. Too less char from the BFB to PTR means insufficient
and the high-temperature heat carrier particles (i.e. sand) heat supply to the fuel pyrolysis/gasification reactions so
coming from the PTR combustor. The gaseous product that the temperatures of the whole system will become
of fuel pyrolysis/gasification (including tars) independently gradually lower and unstable with operation. On the con-
exhausts from the gasifier, whereas the unreacted char trary, if the char moved to the PTR combustor is too
remains in the gasifier to move with the heat carrier parti- many, there must be an excessive heat production inside
cles. The amount of unreacted char is a function of gasifi- the combustor to cause its temperature out of control.
cation conditions, such as temperature, fuel particle Thus, the available gas production (or gasification) effi-
residence time in the BFB, etc. With progress of fuel pyro- ciency via the plant is subject to the quoted char partition
lysis and gasification the temperature of the heat carrier between the gasifier and combustor, although the efficiency
particles has to decrease, but the particles, with char varies as well with fuel’s water content.
blended, could quickly reach the jetted air to allow them The present article is about the proposed gasification
to be conveyed into the PTR. There, the expose of the plant and intends to provide a few relative process funda-
mentals. In a batch fluidized bed gasifier it clarified first
the available C and H conversions, product partition char-
acteristics and gas composition at the end of fuel pyrolysis
Combustion
exhaust gas
in a steam atmosphere. Using these data as the known
input, a process simulation with ASPEN was in turn con-
ducted to demonstrate the possibly available efficiencies
Pneumatic Fine powder in the devised gasification plant for fuels, here coffee
transported bed (ash) Gasification
char combustor
grounds, containing water of up to 45 wt.%. This resulted
gas
in clarification of the suitable fuel water contents that
assure the desired gasification efficiency, such as over
Biomass Fine powder 75%. Finally, a few gasification tests in a 5 kg/h pilot dual
Bubbling (ash) fluidised bed gasification facility simulating the gasification
fluidized bed plant illustrated in Fig. 1 were carried out to identify the
gasifier
dependence of fuel conversion on reaction time of fuel,
Air i.e. on the fuel particle residence time inside the gasifier,
Steam
so as to provide the first basis for designs of the gasification
Fig. 1. Conception of the devised pyrolytic gasification plant. reactor.
246 T. Murakami et al. / Fuel 86 (2007) 244–255

2. Methodology Table 1
Properties of coffee grounds
2.1. Batch fluidized bed gasifier No. 1 No. 2
Proximate (wet-wt.%)
Fig. 2 shows a schematic diagram of the used lab-scale Moisture 10.5 9.3
bubbling fluidized bed gasifier. As detailed in Xu et al. VM 71.8 69.4
FC 16.7 19.3
[5], the bed was made of quartz glass, and was 80 mm in Ash 1.0 2.0
i.d. and 1200 mm high. A golden furnace heated the bed,
Ultimate (db-wt.%)
with temperatures inside the bed measured at 250, 400,
C 52.97 54.9
700 and 1100 mm above the bed distributor (a sintered por- H 6.51 6.12
ous plate). Via batch gasification test in this gasifier we N 2.80 3.07
measured the available C and H conversions and the corre- S 0.05 0.01
sponding product distribution at the end of fuel pyrolysis O 36.62 33.62
HHV (kcal/kg-db) 5260 5682
in a steam atmosphere. As for the biomass gasification
Bulk density (kg/m3) Around 350
plant conceptualised in Fig. 1, the fuel pyrolysis may dom- Particle size <2.0 mm
inate the available gasification efficiency. The facts that
biomass fuel is abundant in volatile matters and the kinetic
rate of fuel pyrolysis is much quicker than that of char
steam gasification should be the causes. Meanwhile, the hydrocarbons of up to C3H8) and the total moles of
product distribution of biomass pyrolysis is closely depen- released C and H.
dent on measurement methods, such as heating device, The tested biomass fuel was dry coffee grounds obtained
because different methods may cause different heating rates from drying/upgrading high-moisture coffee grounds with
to fuel particles. Choosing the bubbling fluidized bed gas- water content over 65 wt.% according to the technology
ifier illustrated in Fig. 2 expected to make the measured of slurry dewatering in kerosene [2]. This article was
fuel pyrolysis characteristics representative of what actually involved with two types of dry coffee grounds (received
occurs in the gasifier conceptualised in Fig. 1. from another company at different time). Table 1 sum-
The test procedure was as follows. After the gasifier was marizes their properties determined with proximate and
set to the desired conditions, one gram of fuel (fuel no. 2 in ultimate analyses. We can see that both the fuels were
Table 2) was dropped into the fluidized bed of silica sand similarly rich in oxygen and contained water of about
particles to start the gasification reactions. The sand had 10 wt.% and volatiles of about 70 wt.%. Nonetheless, slight
a Sauter mean diameter of 190 lm, and the loaded amount differences existed in compositions, which might be due to
in the bed was 3.2 kg. This led to a fluidized particle bed of some different drying/upgrading conditions applied to
500–550 mm high under a steam feed (473 K) of 5.6 g/min them. The fuel no. 2 was used for the batch gasification test
and a N2 tracer gas flux of 4.0 LN/min. The gas from the specified above.
reactor, which was first dedusted, filtered and dried at the
reactor exit, was sampled with gas bags at a time interval 2.2. Process simulation
of 15 s. The sampled gas was then analysed in a micro
gas chromatograph for its molar composition. Referring The software package used was ASPEN plus. In order
further to the N2 tracer flux, the measurement determined to simulate the chemical steps occurring in the gasification
also the moles of various gaseous species (H2, CO, CO2 and plant sketched in Fig. 1, a process model highlighted in
Fig. 3 was constructed. The model consists of two basic
modules, the fuel pyrolytic gasification module and char
coffee grounds combustion module. The gasification module is further
T
T made of a fuel pyrolyzer, a char gasifier and a gas/tar refor-
P filter T : thermocouple mer. In actual gasification plant all these reactors should be
P : pressure sensor integrated together, which have thus the same output tem-
cyclone perature in Fig. 3. Unreacted char from the char gasifier is
completely sent to the char combustor of the combustion
T module, whereinto air preheated with hot product gas is
IDF
steam generator supplied to provide oxygen for combustion. Between the
two modules heat carrier particles (silica sand) are circu-
bag filter
lated. Independent exhausts are available to both the refor-
Gas mer and char combustor. Thus, the model simulated
P T analyzer
pump N
exactly the process features implicated in Fig. 1.
2
This article considered only the temperatures of 1173 K
Fig. 2. A schematic diagram of the lab-scale bubbling fluidized bed and 1073 K for the combustion and gasification modules,
gasifier. respectively. On this basis Fig. 3 shows also how the heat
T. Murakami et al. / Fuel 86 (2007) 244–255 247

Gasification air water Gasification


Tm 1073 K
Product gas Pyrolyzer
573K
Ta Fuel → Char + H2 + CO + + Tars, ΔH 0298 > 0,
373 K Reformer ( ΔH 0298 : Estimated from reaction’s enthalpy balance)
water+steam Water → Steam ΔH 0298 > 0

Ts1 1073 K cyclone 1 Gasifier


heat Char + H2O → CO + H2 ΔH 0298 = 131 kJ/mol
Gasifier Char + CO2 → 2CO ΔH 0298 = 172 kJ/mol
circulated
773K 1073 K 1073 K
sand Reformer
Pyrolyzer CO + H2O → CO2 + H2 ΔH 0298 = −41 kJ/mol
fuel
CH4 + H2 O → CO + 3H2 ΔH 0298 = 206 kJ/mol
Tar + H2O → aCO + bH2 ΔH 0298 > 0, Estimated.
Combustion air, Ta
Ts1 auxil. fuel
Combustor
flue gas Combustion
423K 1173 K 1173 K
cyclone 2 Combustor
373K
Char + 1/2O2 → CO ΔH 0298 = −111 kJ/mol
water+steam
CO + 1/2O2 → CO2 ΔH 0298 = −284 kJ/mol
Fuel + aO2 → bH2O + cCO2 ΔH 0298 , Based on HHV
Fig. 3. A process model simulating the gasification plant sketched in
Fig. 1.
Fig. 4. Considered reactions in process simulation.

carried with hot flue gas (1173 K) and hot product gas
(1073 K) is recovered in the modelling. It was supposed occur in two successive steps, first into CO and then into
that the virgin product gas from the reformer undergoes CO2 via oxidizing the formed CO.
a heat exchange first with water supplied for the gasifier. At this stage, we calculated merely the heat and mass
This reduces the gas’ temperature to an intermediate value balance of the whole system so as to demonstrate the pos-
of Tm (773 K) and also converts the water into a water- sibly available cold gas efficiency and C and H conversions
steam mixture of 373 K (under 1.0 atm). The remaining for fuels with various different water contents. The cold gas
usable heat with the product gas, i.e. the enthalpy between efficiency refers to a ratio of HHVs between the produced
Tm and its exhaust temperature of 573 K, is then consumed gas and treated fuel. For the calculation it was assumed
in heating the air fed to the char combustor by raising the that the fuel fed into the pyrolyzer is completely pyrolyzed
airflow’s temperature to Ta. In another heat exchanger, the to form gas, tars and char according to the product
heat from the flue gas converts further the aforementioned percentage identified in the preceding batch fluidised bed
water-steam mixture into a steam stream of Ts1. The flue gasification test (the Section 2.1). Thus, increasing the fuel
gas itself approaches in turn its final exhaust temperature water content has to decrease the available amounts of dry
of 423 K. In case Ts1 is lower than 773 K, further heating fuel and the formed char. This would lead the char com-
the steam stream to 773 K, the desired temperature of bustion to be unable to provide enough exothermal heat
gasification reagent, was conducted. In the depicted heat to maintain the system’s temperature. Once this occurs,
network the temperatures Ta and Tm are two dependent an auxiliary fuel feed (with the same water content as that
variables. Another dependent variable is the amount of of the fuel fed to the gasifier) is directed to the char com-
circulated silica sand. Because the temperature difference bustor so that the burning of this fuel can compensate
between the combustion and gasification modules is fixed for the insufficient heat.
at 100 K, the circulated particle amount has to vary with Consequently, the computation was iteratively per-
the amount and composition of fuel feed. Herein, however, formed to determine the percents of gasified char (the left
only the fuel’s water content is varied (based on coffee being combusted) or the necessarily required auxiliary fuel
grounds no. 2 in Table 1), while the fuel amount (wet-base) feed for char combustor. This in turn determines the circu-
remains in a constant. lated heat carrier particle amount and the temperatures Tm
Fig. 4 summarizes the considered chemical reactions, and Ta. It was suggested that 99% of the tars from pyrolysis
with as well their accompanying enthalpy variations, for is reformed, whereas reformation of CmHn was not consid-
each processor of the process model. In the char gasifier, ered. In addition, no kinetic limitation was applied to all
both steam and CO2 gasifications were considered. The ref- reactions (such as char combustion and gasification). The
ormation of tars and CH4 (representative of CmHn) is sug- needed enthalpy changes for reactions of fuel pyrolysis
gested to occur only in the reformer. In the reformer water and tar reformation were estimated as the enthalpy differ-
gas shift (WGS) reaction was supposed to be in its equilib- ence between their reaction products and reactants. In
rium state, noting that the gasification module has temper- these estimations the standard enthalpy of tars was roughly
atures over 973 K. The char combustion was thought to determined according to the tars’ element composition
248 T. Murakami et al. / Fuel 86 (2007) 244–255

obtained from element analysis (will be shown in the flow controller from a port located slightly above the fluid-
Section 3). ized-particle bed surface. With respective cyclones and gas
exhaustion lines, the pressures inside the fuel gasifier and
2.3. Pilot dual fluidised bed gasification facility char combustor could be independently controlled through
adjusting the performance of their IDFs (induction draft
Fig. 5 shows a sketch of the pilot dual fluidized bed gas- fans). In order to prevent intermixing of gases between
ification facility adopted to clarify the necessarily entailed the riser and gasifier, the gasifier had a specially designed
reaction time for the fuel and to demonstrate the chemical bed structure. This structure, called reactor siphon [7],
possibility of the devised gasification process illustrated in allowed the reactions of fuel pyrolysis and char gasification
Fig. 1. In the experimental facility, the pyrolytic gasifier to proceed inside the BFB, while it enabled as well the con-
and char combustor were separately installed. This makes trol of particle flow rate through the bed. The flue gas from
the whole facility be a modified circulating fluidized bed the riser combustor was directly cooled down and vented
(CFB) so that its operation can take advantage of the expe- into atmosphere after it passed through a bagfilter. Com-
riences from CFB combustion. On the other hand, the pared to this, the product gas from the BFB gasifier was first
acquired result in this facility relative to chemical reactions burned out in a combustion tube packed with ceramic balls
is surely applicable to the compact configuration shown in and heated electrically to about 823 K. Then, the gas was
Fig. 1 because for both the fuel pyrolysis/gasification pro- subject to the same ventilation system as adopted for the
ceeds in a BFB and char combustion in a PTR. However, flue gas from the PTR combustor.
the particle circulation control should be different from Herein, the presented test results were under given tem-
each other, raising thus another independent research of peratures of about 1073 K for gasifier and about 1103 K
ours on it in a cold model fluidised bed configured exactly for char combustor. The tested fuel was fuel no. 1 specified
according to Fig. 1 [6]. in Table 1. The fuel feed rate and steam-to-fuel mass ratio
As indicated in Fig. 5, the gasifier was a rectangular bed, were both fixed at about 3.6 kg/h and 1.0 kg/kg (S/C =
with a cross section of 80 · 370 mm2 and a height of 1.22 mol/mol), respectively. Thus, the only varied para-
1800 mm. The riser was a circular column of 52.7 mm in meter was the residence time of fuel particles inside the
i.d. and 6400 mm in height. Both of these reactors were elec- gasifier, which was calculated from
trically heated (at least to 973 K), while additional feed of
Particle amount held in the gasifier
propane into them to facilitate temperature rise was also ð1Þ
possible. The biomass fuel, as specified in Table 1, was sup- Particles circulated in unit time
plied via a table feeder and the fed fuel was in turn carried under an assumption that the particles pass through the
into the gasifier with a well-metered argon stream in a mass gasifier in a flow of plug type. For this calculation, the

Exhaust gas Gas


cooler Bag filter

IDF
Gas analyzer

Gs measure
Combustor
50A × 6400mmH
Gasification gas Combustion Gas
Bag filter
tube cooler
filter

Tar trap Gas analyzer


IDF
Biomass

Gasifier
370 × 80 × 1800mmH
37

FDF

Steam
generator

Fig. 5. A schematic diagram of the employed pilot dual fluidized bed gasification facility.
T. Murakami et al. / Fuel 86 (2007) 244–255 249

Table 2 15
C and H conversions, product distribution and composition at the end of Fuel : 1g

Concentration [vol.%]
fuel pyrolysis in a steam-blown fluidized bed of 1073 K 12 Temp. : 1073K
Conversion to gas (%) C: 63.0, H: 88.7 CO
Product distribution (wt.%)a Gas: 75, Char: 22.1, Tars: 2.9 9 H2
Molar composition (vol.%) 19.3H2 + 38.2CO + 9.5CO2 + 17.5CH4
(free of tracer gas) + 9.9C2H4 + 5.2C2H6 + 0.4C3H6 6
Char composition (wt.%)b 63.82C + 0.0H + 11.57N + 14.33O
+ 0.003S + 0.36Cl + 9.90 ash 3
Tar composition (wt.%)c 64.40C + 6.23H + 11.10N + 15.10O a
+ 3.01S + 0.11Cl 0
a
The fraction of tars was from pilot gasification test.
b
Derived from element mass balance. 100
c
From measuring the tar sample taken in pilot gasification test. b

Element conversion [%]


80

60
required particle circulation rate at the tested temperature
of 1103 K was experimentally measured by using a partic- 40 Fuel : 1g
ularly designed heat-resistant valve [8]. Temp. : 1073K
The product gas was sampled at the gas exit of the gas- 20 C
ifier cyclone to trap the tars with the gas (for tar element H
0
analysis) and to measure the gas’ molar composition. The 0 30 60 90 120
sample gas flow, induced with a suction pump, was first Time [sec]
condensed in a water condenser to trap most of the steam Fig. 6. Time series of (a) molar gas composition from a batch gasification
present in the gas. Then, the gas passed through three test in the steam-blown bubbling fluidised bed gasifier and (b) the
water bubblers kept in an ice-water bath to further trap corresponding integrated C and H conversions estimated from the molar
the tars with the gas. A wet gas meter was set behind the composition.
last water bubbler to measure the sample gas flow rate that
was controlled at about 2.0 nL/min. By using a fabric filter
to capture further the tars escaping from the water bub- the corresponding accumulative C and H conversions
blers (BTX, phenols and poly-aromatics exclusively) it (Fig. 6b) calculated according to the gas composition.
was demonstrated that the tar trapping efficiency in our Surely, in this batch test the produced gas, a mixture of
tar collection system was about 90%. The molar composi- H2, CO, CO2, CH4, C2H4, C2H6 and C3H6, should be first
tion of product gas was measured in the same micro GC gradually more and then gradually less with the progress of
adopted for the batch gasification test, while the product reactions. The concentrations of CO and H2 in the outlet
gas volume was determined according to the rate of argon gas (Fig. 6a) thus both exhibited a peak and finally
stream used to carry fuel. approached values nearby zero. On account of the much
Extraction of the tars trapped in the condenser and slower char gasification in steam than fuel pyrolysis, we
water bubblers was conducted by following successively took the end of quick product gas release marked with
the steps of collecting the tarry water, washing the vessels the vertical broken line in Fig. 6 to represent the finish of
(using acetone), filtrating the collected water–acetone fuel pyrolysis. According to this figure, the pyrolysis ended
liquid, vacuum vaporization of water and acetone at about 43 s after dropping 1.0 g fuel into the reactor. This
(<333 K) and drying tars (the vaporization residue) in time is longer than our anticipated values necessary to fuel
warm airflow (<323 K). The resulting dry tars were then pyrolysis at the tested temperature of 1073 K. In a tubular
weighed to estimate the tar content in the product gas reactor and N2 atmosphere, Bingyan et al. [9] reported a
and analysed for its element composition (see Table 2). time of about 15 s for finishing the biomass pyrolysis at
The tar content, however, will not be reported in this article 1073 K. Here the longer time was considered to be due to
because of its little relevance to the article’s purpose. the too big reactor used. Nonetheless, our purpose was
other than measuring the time required for finishing fuel
3. Results and discussion pyrolysis. It was for determination of the conversions of
C and H into gas and the corresponding distribution of
3.1. Pyrolytic characteristics in steam-blown fluidized bed pyrolytic products, i.e. gas, char and tars, in steam-blown
fluidised bed. Thus, the reactor size would not much affect
As indicated in Section 2.1, gasification tests in the batch these intended data.
fluidized bed gasifier (Fig. 2) were conducted to demon- Fig. 6b shows the integrative C and H conversions (into
strate the pyrolytic characteristics of the fuel coffee gas) corresponding to the gas molar composition exempli-
grounds. Fig. 6 exemplifies the time series of molar gas fied in Fig. 6a. The integrative conversion for time t was
composition (Fig. 6a) measured at the gasifier exit and calculated with (taking C as the example)
250 T. Murakami et al. / Fuel 86 (2007) 244–255

ðMoles of released C until time tÞ the fuel. Table 2 verifies this by showing that the C/H ratio
; ð2Þ
ðMoles of C in the 1:0 g fuelÞ is 64.4/6.23 for tars but 54.9/6.12 for the tested coffee
grounds (no. 2 in Table 1).
where, the moles of released C as well as H until time t were
estimated from the time-series gas composition and dry gas 3.2. Theoretically possible gasification efficiency
flow rate determined according to the flow rate of N2 tra-
cer. Nonetheless, the H conversion defined via Eq. (2) re- The preceding batch test was for a dried fuel and also
fers to an explicit value because, on the one hand, part of did not consider the heat balance between the fuel gasifier
the fuel’s H would be converted into water in fuel pyrolysis and char combustor. Using the result of such a batch test
and, on the other hand, the product gas contains also the H (in Table 2) as the required known input, a process simula-
taken from steam via, for example, char steam gasification, tion with ASPEN according to the process and chemical
tar/hydrocarbon reforming and water gas shift. models outlined in Figs. 3 and 4 was conducted to deter-
After fuel feed, the integrative conversions of C and H mine the possibly available gasification efficiencies for the
increased with reaction time, in response to the gradually whole plant of Fig. 1.
more product gas released from the reactor. At the end Table 3 summarizes the simulated conditions and the
of fuel pyrolysis (broken-line annotated) they reached required known input. The fuel treated was the same coffee
about 63% and 89%, respectively (see Table 2). These cor- grounds no. 2 specified in Table 1 but with water contents
responded to a fuel mass conversion of about 75 wt.%. The varying in 3.0–45.0 wt.%. The wet-base fuel feed rate was
remaining 25 wt.% should be char and tars. In order to 250 kg/h. Steam at 773 K was supplied into the gasifier
determine their partition fractions we supposed that the according to a mass ratio of 1.0 kg/kg of steam to dry fuel.
tar yield is the same as that measured hereafter via pilot Water with the fuel was supposed to be completely vapor-
gasification test under the same 1073 K. There, it was ized inside the pyrolyzer (see chemical models in Fig. 4),
found that the evolved tars amounted to 2.9 wt.% of dry causing it to mix with the supplied steam and in turn to
fuel so that the char production became 22.1 wt.%. In this raise its temperature to the gasification temperature of
mass balance the produced H2O in fuel pyrolysis was sug- 1073 K. The employed heat carrier particles were silica
gested negligible and fuel ash was thought to be fully with sand of 190 lm in Sauter mean diameter. An air ratio of
char. The molar composition of product gas listed in Table 1.5 against the char amount from the gasifier was taken
2 refers to the measured values but excluding N2 tracer, to determine the airflow rate into the char combustor.
whereas the element composition of char was determined Table 3 mentions also several other characteristic tempera-
according to element mass balances. For this, the element tures, which were all from the process model outlined in
content of tars, shown as well in Table 2, was treated as Fig. 3. The data about fuel pyrolysis were based on Table
known data, which was obtained by measuring the tar sam- 2.
ple taken in the pilot gasification test at 1073 K shown Fig. 7 shows the simulation results acquired, with
lately in Section 3.3. Fig. 7a for the available cold gas efficiency (left Y, based
The similar batch gasification tests were conducted also on HHV) and C and H conversions (right Y) and Fig. 7b
at 973 and 1023 K (the other conditional parameters being for the necessarily required conditions that assure the effi-
the same as at 1073 K), revealing obvious decreases in the ciency and conversions. The conditions were presented
C and H conversions (into gas) and in the gas production with circulation rate of heat carrier particles (left Y) and
percentage with reducing the reaction temperature. The auxiliary fuel amount fed to the combustor (right Y). All
acquired data clarified also that for realizing C conversions the parameters were expressed as a function of fuel water
higher than 60% the temperature has to be over 1073 K. content. While the displayed circulation rate refers to a spe-
Besides, the H conversion was found to be always higher cific value against the treated dry fuel amount, the amount
than C conversion. While this reflects a general phenome- of auxiliary fuel (with the same water content as that fed to
non of biomass pyrolysis, in our case the presence of steam
may make the H conversion (against fuel H only) rather
Table 3
large because steam reforming of tars and hydrocarbons Conditions and known input for process simulation
would occur to convert some H of steam into product
Fuel: coffee grounds no. 2, 250 kg/h, with water content varied in
gas. The H conversion was lower than 100%. This is due 3.0–45.0 wt.%
to the fact that part of the fuel H had to become H2O dur- Gasifier: temperature = 1073 k; Steam/dry-fuel = 1.0 (mass ratio);
ing fuel pyrolysis. On the other hand, it shows also that the Steam temperature = 773 K
accompanying steam reforming reactions for the tested Combustor: temperature = 1173 k; Air ratio = 1.5;
pyrolysis in steam did not occur to an extensive degree to Air temperature = Heated via hot product gas
Heat carrier particles (sand): Sauter mean dp = 190 lm; Gs: by heat
make the H conversion over 100%. In the test we also balance
observed serious tar deposition onto pipes behind the gas- Exhaust temperatures: flue gas temperature = 423 K; Product gas
ifier, which made the pipes brown, even black. Because of temperature = 573 K
the higher conversion of H than of C into the product Pyrolysis characteristics (product allocation, gas/tar composition): in
gas, the tars must have a C-to-H ratio higher than that in Table 2
T. Murakami et al. / Fuel 86 (2007) 244–255 251

100 150 quently the C amount for gas production. The converted
a H also decreases because the decreased C amount for steam

Element conversion [%]


Cold gas efficiency [%]
90 gasification reduces the converted H from H2O (i.e. steam).
120 Corresponding to the simulation conditions that 99% tars
80
are reformed and the combusted char is almost free of H,
Cold gas efficiency
70 C conversion
the resulting H conversion is always higher than 100%.
90
H conversion When no heat loss is considered, the available cold gas effi-
60
ciency can be over 80% for fuels with water content lower
50 60 than 40 wt.%. In order to reach a cold gas efficiency of
85%, the fuel’s water content has to be lower than 10 wt.%.
50 20
Circulation rate / Dry fuel feed rate As the necessary conditions assuring the gasification effi-
b

Auxiliary fuel [wet-kg/h]


[-]

40
Auxiliary fuel ciencies mentioned above, Fig. 7b demonstrates that with
15 raising the fuel’s water content the entailed relative circula-
Dry fuel feed rate
Circulation rate

30 tion rate of heat carrier particles against the treated dry


10 fuel amount becomes higher (left Y). Meanwhile, until a
20
water content of 25 wt.% no auxiliary fuel is required (right
5 Y), whereas further increase of the fuel water content has
10
to lead to gradually higher consumption of auxiliary fuel.
0
0 10 20 30 40 50
0 These clarifications are obviously reasonable because they
Fuel's water content [wt%] responded just to the change of fuel enthalpy with fuel
water content.
Fig. 7. Simulation results of (a) gasification efficiency shown with cold gas
efficiency and C and H conversions and (b) the correspondingly required
Nonetheless, the C and H conversions show in Fig. 7a
particle circulation rate and auxiliary fuel feed to the combustor. are higher than the experimental data indicated in Table
2 (from batch gasification test). The steam gasification of
the gasifier) indicates the value without heat loss from the char without kinetic limit, complete tar reforming (99%
system. Hence, the unavoidable heat loss in practical plants reformed) and zero heat loss considered in the simulation
would make both the parameters surely higher, as will be should be the causes. Although these assumptions are too
clarified in Fig. 8. ideal, the simulation itself shows its significance by clarify-
As anticipated, the available cold gas efficiency and C ing theoretically the available highest gasification efficien-
and H conversions (Fig. 7a) all decrease with raising the cies for the examined fuels and technology.
water content in the fuel. The higher the fuel water content, On the other hand, based on Fig. 7 we can simply inves-
the more the heat required to vaporize the water and to tigate how heat loss affects the gasification efficiency and its
heat the resulting steam to the gasifier temperature (here required conditions. Fig. 8 presents the variations of the
1073 K). On the other hand, this part of heat cannot be available cold gas efficiency (left Y) and entailed auxiliary
completely recovered from the product gas because its out- fuel feed (right Y) with heat loss of up to 30% of the fuel
let temperature is 573 K. It was considered that until down- energy inputted into the gasifier. The considered case was
stream scrubber the product gas is better to be over such a for a water content of 10 wt.%. The revised cold gas effi-
temperature in order to avoid substantial deposition of tars ciency was computed from
on pipeline. Hence, the higher water content means the ðEfficiency wihout heat lossÞ=ð1 þ heat loss percentageÞ;
more C needed to be combusted to maintain the tempera-
ð3Þ
tures of the combustor and fuel gasifier, lowering conse-
while the modified necessary auxiliary fuel feed was deter-
mined as
100 25 8
Fuel's water content : 10 wt. %
> 0 if equivalent energy of gasified char
Cold gas efficiency [%]

Auxiliary fuel [wet-kg/h]

Cold gas efficiency >


>
90 Auxiliary fuel 20 >
< at zero heat loss > Actual heat loss
80 15 > ‘Actual heat loss’  ‘Equivalent energy of gasified char
>
>
>
:
70 10 at zero heat loss’; Otherwise:
ð4Þ
60 5
As expected, raising the heat loss surely decreases the cold
50 0
0 5 10 15 20 25 30 gas efficiency and increases the required auxiliary fuel feed.
Heat loss [%] If supposing that the heat loss varies generally from 5% to
Fig. 8. Theoretically possible cold gas efficiency with consideration of heat
10% of the treated fuel energy, we can see that the cold gas
loss of up to 30% of fuel energy fed into the gasifier under a fuel water efficiency without kinetic constrainments can be still 77%–
content of 10 wt.%. 81%. On this basis, we suggest that the water content of the
252 T. Murakami et al. / Fuel 86 (2007) 244–255

coffee grounds fuel for gasification is better to be about 1180


Gasifier Combustor
10 wt.% in order to assure the gasification efficiencies above 0.30m 0.30m
75%. Although no auxiliary fuel feed is needed in Fig. 7 for 1160
0.65m 1.40m

Temperature [K]
the fuel containing 10 wt.% water, the presence of heat loss 1.40m 2.50m
1140
makes it absolutely necessary when the loss is over 6%. The 4.80m
result judges further that the fuel’s water content should be 1120
around 10 wt.% in order to achieve energy conversion effi-
ciencies over 75%. This provides actually a design/opera- 1100
tion standard for the upstream fuel drier.
1080

3.3. Necessary reaction time 1060


0 10 20 30 40 50
How long should the reaction time be for fuel particles Time [min]
inside the gasifier in order to finish fuel pyrolysis and to Fig. 10. Typical time series of the temperatures inside the BFB gasifier
realize the cold gas efficiency desired, such as over 75%. (d,j, m) and riser combustor (h, s, n, e) of the employed pilot dual bed
This is actually another unavoidable problem that has to facility during a test.
be answered for developing the devised gasification plant.
For this answer a few tests with the coffee grounds no. 1
were conducted in the pilot dual fluidised bed gasification was for an explicit residence time of 160 s of fuel particles
facility sketched in Fig. 5 by varying the residence time inside the gasifier determined according to formulation (1).
of fuel particles inside the gasifier. According to formula- The correspondingly required airflow into the riser was
tion (1) the different residence times were realized by 70 nL/min. Fig. 10 shows the time series of a few typical
adjusting the gas velocity in the riser combustor to vary temperatures measured for this test. The displayed temper-
the particle circulation rate. From these gasification tests atures were for different bed heights in the BFB gasifier
we also gained the tar sample analysed for the tar element (solid marks) and riser combustor (open marks), and the
composition listed in Table 2. time zero refers to the onset of fuel feed.
Fig. 9 exemplifies the time series of molar composition As for fuel gasifier we can see from Fig. 10 that its free-
(including tracer gas) and HHV of the produced gas board (m) remained in a relatively stable temperature,
(Fig. 9a) and their corresponding transient cold gas effi- whereas its particle bed (d j) exhibited an evident local
ciency and C and H conversions (Fig. 9b). The plotted test temperature drop with fuel feed. This is indicative of the
occurrence of endothermic fuel pyrolysis/gasification reac-
tions inside the particle bed. Being subject to electric heat-
60 4500 ing, the dropped temperatures began to rebound to higher
Gas heating value [kcal/mN ]

Gas heating value


a values since 5 min and in turn reached relatively steady
Gas concentration [%]

50 H2 CH4
CO C2H4+C2H6 4200 values at about 30 min after the fuel feed. At last the upper
40 CO2 C3H6
3900
particle bed (j) and freeboard (m) stagnated at nearly the
30 same temperature, which was the situation expected in set-
3600 ting. The bed bottom (d, 0.3 m above the distributor) had
20
slightly lower temperatures, say, for about 15 K, which
10 3300 shown just an effect of the inlet steam. Inside the riser the
temperatures experienced a local drop as well with fuel feed
3

0 3000
by responding to the lowered particle temperatures inside
100 100
b
the gasifier. Nonetheless, the temperature drop itself was
Element conversion [%]
Cold gas efficiency [%]

90 90 smaller than in the gasifier, while its rebounding was much


80 80
quicker. This allowed the riser’s temperatures at different
elevations to reach their respective steady-states in about
70 70 15 min. Meanwhile, the riser bottom (s) possessed the low-
60 60 est temperature, showing just the process feature that the
Cold gas efficiency particles from the gasifier have the lowest temperature so
50 C conversion 50
H conversion that they need to be reheated in traveling through the riser
40 40 combustor. In regard to the exemplified cases the most
0 10 20 30 40 50
Time [min] intensive char combustion occurred possibly in the middle
section of the rise, causing the highest temperature inside
Fig. 9. Time series of molar composition of rude product gas (with tracer) the riser to emerge at the height of about 2.5 m (n).
and the corresponding transient HHV, cold gas efficiency and C and H
conversions measured in the dual bed pilot gasification facility under an
The gas compositions in Fig. 9 show evidently that the
explicit fuel particle residence time of 160 s inside the gasifier (the other operation trended to become quasi-steady since 10 min
conditional parameters being detailed in Section 2.3). after fuel feed (based on the displayed conversions and
T. Murakami et al. / Fuel 86 (2007) 244–255 253

HHVs). This, compared to the temperature profiles in 100 100


Fig. 10, reveals that the reaction temperature variation in

Element conversion [%]


Cold gas efficiency [%]
90 90
about 10 K after 10 min of fuel feed did not greatly affect
the gas generation characteristics. Under the tested condi- 80
80
tions (see the Section 2.3) CO took the highest concentra- 70
tion of about 28 vol.% in the product gas (for the tested 70
case the tracer concentration being 24 vol.%). Following 60
Cold gas efficiency
this were H2, CH4, CO2, C2 and C3 hydrocarbons in succes- 50 C conversion 60
sion (left Y, Fig. 9a). The molar ratio of H2 to CO was H conversion
about 0.6, with the absolute H2 concentration being about 40 50
0 300 600 900 1200
15 vol.%. Corresponding to the mentioned gas composition Fuel residence time in gasifier [sec]
the gas’ HHV (right Y, Fig. 9a) was about 3900 kcal/nm3.
Concerning the product gases from biomass gasification Fig. 11. Steady-state cold gas efficiency and its corresponding C and H
conversions under different explicit residence times of fuel particles inside
this HHV represents a high caloric value, verifying the dis-
the gasifier of the pilot dual bed facility.
tinctive merit of the dual fluidised gasification technology
mentioned in Section 1. The concentration diagrams in
Fig. 9a clarify also that the product gas contained hydro- whose test conditions are similar to those employed here.
carbons (including CH4) more than 20 vol.%. In the view That is, except for the similar reaction temperature of
of H2 production it is important to convert these gas 1073 K, the fuel feed rate was also about 3.5 kg/h and
species into CO and H2 via reforming. steam-to-fuel mass ratio was around 1.0.
Fig. 9b demonstrates that the achieved cold gas efficiency The figure shows that with increasing the fuel particle
was 69% or so, corresponding to C and H conversions of residence time the conversions of C and H and their corre-
about 65% and 88%, respectively. These conversions are sponding cold gas efficiency increased gradually. However,
somehow very close to the values shown in Table 2 for fuel when prolonging the time from 160 to 1200 s the increases
pyrolysis obtained in the batch gasification tests under a in conversion and efficiency were confined to a few percents
similar temperature. It indicates essentially that the reac- (<5.0%). Corresponding to this, the increase in the resi-
tions occurring inside the gasifier of the pilot gasification dence time of fuel particles from 2.3 s to 160 s induced dis-
facility (Fig. 5) were principally fuel pyrolysis, just aligning tinctive increases in both efficiency and conversion. While
with our general anticipation about biomass gasification. the H conversion (n m) elevated to 88% from 62%, the
That is, its gas production should mainly rely on fuel pyro- cold gas efficiency (s d) and C conversion (h j) exhibited
lysis rather than on char gasification. Nonetheless, the increases of about 13%. Consequently, for the devised
reported test suffered probably a fuel loss of about 5% of gasification plant (Fig. 1) the required explicit fuel particle
the original feed because the feed of fuel particles into the residence time inside its gasifier can be lower than 160 s.
gasifier’s freeboard would lead to an easy elutriation of fine Rather longer residence time is beneficial to fuel conver-
particles. In a strict sense we thus believe that the C and H sion, but the actually available benefit is slight. On the
conversions and cold gas efficiency attainable in the pilot other hand, the gasification inside the riser cannot provide
gasification test are slightly higher than those from fuel the fuel sufficiently long time to finish even pyrolysis
pyrolysis shown in Table 2. This tokens the occurrence of because inside the riser the available average residence time
steam reforming of C inside the gasifier. The longer reaction for the gas and particles is below 10 s [11]. Examining
time in this case (160 s) than in Fig. 6 (concerned time: 45 s) Fig. 11 we can see actually that the C and H conversions
should be the cause. indicated with the keys s, h and n are truly much lower
A test similar to the illustration of Figs. 9 and 10 was than the corresponding values mentioned in Table 2 for
conducted also at an explicit residence time of 1200 s of fuel pyrolysis. This, in essence, verifies that in this case
fuel particles inside the gasifier. Meanwhile, another recent the fuel pyrolysis was not yet completed. This is also why
publication of ours [10] had reported a set of data obtained the design of the gasification plant conceptualized in
by gasifying the same fuel in the riser of the same facility. Fig. 1 deploys its fuel gasification into a BFB and its char
In this case, the riser becomes the gasifier so that the fuel combustion into a riser coupled to the BFB. Regarding this
particle residence time is basically equal to the gas resi- technical choice we had presented another publication [10]
dence time because particles move with the gas. At the to compare the available efficiencies and conversions when
reported superficial gas velocity of 2.82 m/s (at bed temper- gasifying the same fuel inside the riser and BFB of a dual
ature) the fuel particle residence time becomes 2.3 s or so fluidized bed system, respectively.
(riser height being 6.4 m). By summarizing all of these men- Nonetheless, the element (C and H) conversions and
tioned data we have then Fig. 11 where the steady-state cold gas efficiency shown here are much lower than those
cold gas efficiency and its corresponding C and H conver- indicated in Figs. 7a and 8 (left Y). For example, at the fuel
sions are displayed as functions of the explicit fuel particle water content of 10 wt.% the available C and H conver-
residence time. In the plot the data indicated by the keys s, sions and cold gas efficiency can be, respectively, 75%,
h and n are from the gasification test inside the riser [10], 140% and 85% in Fig. 7a but only 68%, 92% and 72% in
254 T. Murakami et al. / Fuel 86 (2007) 244–255

Fig. 11 at the residence time of 1200 s. The production of with the product gas from the BFB gasifier and that
tars and a low degree of C steam gasification involved in with the flue gas from the char combustor are effi-
the gasification test should be the reasons for the men- ciently recovered to produce steam reagent and to
tioned lower experimental fuel conversion and cold gas effi- preheat combustion air, and (b) the tars generated
ciency on comparison of their theoretical values predicted in fuel pyrolysis are mostly reformed. The simulation
in the process simulation. As early mentioned in Section clarified also that with increasing the fuel’s water con-
3.2, the preconditions for the simulation results shown in tent the available cold gas efficiency and C and H
Figs. 7 and 8 are that 99% of the tars generated in fuel conversions decrease, whereas the required specific
pyrolysis is reformed and all unburned char (i.e. C) is gas- circulation rate of heat carrier particles with respect
ified through reaction with steam. The fact in the experi- to the treated dry fuel amount increases. When con-
ment was that up to 3.0 wt.% of the fuel mass was sidering the practical applications with more than
present as tars (at 1073 K), and the slow char gasification 3.0% of fuel’s enthalpy as heat loss, we suggested that
reaction might be unable to afford to a desired amount the fuel for treatment should contain water not much
of char to be gasified. Hence, the fuel conversion from over 10 wt.% in order to maintain the available cold
experiment must be lower than the predicted value. Com- gas efficiency higher than 75%.
paring Figs. 11 and 7a clarifies that the C conversion (3) Gasifying dried coffee grounds in a pilot dual fluid-
shared differences only of a few percents (at a water content ized bed gasification facility simulating the devised
of 10 wt.%), but the differences in H conversion are over gasification plant clarified that the explicit residence
40%. This is indicative of the low degree of tar reforming time of fuel particles inside the BFB gasifier (using
and char steam gasification reactions occurring in the steam as the gasification reagent) calculated accord-
experiments. On the other hand, we should also note that ing to a plug granular flow assumption could be
in the presented tests a few percents of supplied fuel (up shorter than 160 s. This time can guarantee the finish
to 5.0%) were entrained with the gas flow from the gasifier of fuel pyrolysis so that further prolonging the fuel’s
because the fuel was fed into the gasifier’s freeboard. This residence time did not much improve the realized fuel
further lowered the experimental fuel conversion into gas. conversion and cold gas efficiency. The result indi-
cates essentially that as for biomass gasification the
4. Conclusions generated product gas comes basically from fuel
pyrolysis. Through the tests we demonstrated also
With both experimental tests and process simulation, that the PTR reactor of the experimental facility
the present study clarified the following engineering funda- was unable to provide the fuel sufficiently long reac-
mentals relative to chemical reactions and heat/mass bal- tion time to complete even fuel pyrolysis. This dem-
ances that are necessary to the development of a newly onstration justifies the superior technical choice of
devised dual bed gasification plant for biomass which the devised dual bed gasification plant, which should
adopts a bubbling fluidized bed (BFB) as its gasifier and arrange its BFB as the fuel gasifier and its PTR as the
a pneumatic transported riser (PTR) coupled to the BFB char combustor.
as its char combustor.
Acknowledgement
(1) The gasification of, for example, 1.0 g biomass fuel in
a steam-blown fluidized bed in batch format The work was conducted during a research program fi-
appeared to be effective to measure the fuel pyrolysis nanced by The New Energy and Industrial Technology
characteristics in BFB reactors and steam atmo- Development Organization (NEDO), Japan, on developing
sphere. As for our tested biomass fuel, a kind of dried an advanced upgrading and pyrolytic gasification system
coffee grounds (water content being 10 wt.%), it was for biomass with high water content. The authors are also
demonstrated that the fuel pyrolysis under a reaction grateful to Mr. Minoru Asai and Mr. Shigeru Kitano of the
temperature of 1073 K and a steam atmosphere was same company for their helps in experiment.
able to convert 63% of fuel C into product gas. The
corresponding explicit H conversion with respect to
fuel H reached about 90%. Both of these led about References
75% of the fuel mass to be converted into product
[1] The Ministry of Agriculture, Forestry, Fisheries of Japan, Overall
gas. strategy of Japan on biomass, 2001 [in Japanese].
(2) With the fuel pyrolysis characteristics clarified above [2] Mito Y, Komatsu N, Hasegawa I, Mae K. Slurry dewatering process
as the necessarily required known input, a process for biomass. In: Proc int conf on coal sci technol (ICCS&T), IEA-
simulation using the software package ASPEN dem- Clean Coal Center, 2005, Paper 2E01.
[3] Pfeifer C, Rauch R, Hofbauer H. In-bed catalytic tar reduction in a
onstrated that the devised dual fluidised bed gasifica-
dual fluidized bed biomass steam gasifier. Ing Eng Chem Res
tion plant can sustain its independent heat and mass 2004;43:1634–40.
loops with a cold gas efficiency over 75% at a reaction [4] Paisley MA, Farries MC, Black JW, Irving JM, Overend RP.
temperature of 1073 K, provided (a) the heat carried Preliminary operating results from the Battelle/FERCO gasification
T. Murakami et al. / Fuel 86 (2007) 244–255 255

demonstration plant in Burlington, Vermont, USA. In: First word [8] Xu G, Murakami T, Suda T. An apparatus for measuring parti-
congress and exhibition on biomass for energy and industry, Sevilla, cle flow rate, Japanese patent 2005; Appl. no. 2005-295036 [in
June, 2000. Japanese].
[5] Xu G, Murakami T, Suda T, Kusama S, Fujimori T. Distinctive of [9] Bingyan X, Chuangzhi W, Zhengfen L, Xiguang Z. Kinetic study on
CaO additive on atmospheric gasification of biomass at different biomass gasification. Solar Energy 1992;49:199–204.
temperatures. Ind Eng Chem Res 2005;44:5864–8. [10] Xu G, Murakami T, Suda T, Matsuzawa Y, Tani H. The superior
[6] Murakami T. Biomass gasification characteristics in a pyrolytic technical choice for dual fluidised bed gasification of biomass. Ind
fluidised bed gasifier, Presented in tutorial meeting of gasification Eng Chem Res 2006;45:2281–6.
section of the Japan institute of energy, January 2006, Tokyo. [11] Petersen I. Sewage sludge gasification in the circulating fluidized bed –
[7] Xu G, Murakami T, Suda T, Matsuzawa Y. Reactor siphon and its Experiments and modeling. Doctoral thesis; Hamburg: Technical
control of particle flow rate when integrated into a circulating University Hamburg-Harburg (TUHH); 2004.
fluidized bed. Ind Eng Chem Res 2005;44:9347–54.

Vous aimerez peut-être aussi