Está en la página 1de 10

Energy 50 (2013) 333e342

Contents lists available at SciVerse ScienceDirect

Energy
journal homepage: www.elsevier.com/locate/energy

Characterisation of tars from biomass gasication: Effect of the


operating conditions
J.J. Hernndez*, R. Ballesteros, G. Aranda
Escuela Tcnica Superior de Ingenieros Industriales, Universidad de Castilla-La Mancha, Avda. Camilo Jos Cela s/n (edif. Politcnica), 13071 Ciudad Real, Spain

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 20 July 2012
Received in revised form
5 December 2012
Accepted 7 December 2012
Available online 5 January 2013

Tars formation remains the main technical hurdle in the development of biomass gasication at
commercial scale, due to its associated operating problems (condensation, catalyst deactivation,
polymerisation). This work aims to study the effect of the operating conditions (biomass/air ratio,
temperature and gasifying agent) on the production and the technical and environmental hazard
(tar dew point and carcinogenic potential) of tars produced in a small-scale drop-tube gasier,
dealcoholised marc of grape being used as biomass fuel. An analytical HPLC method has been used for the
detection of polycyclic aromatic hydrocarbons (PAHs), benzene, toluene, ethyl-benzene and xylenes
(BTEX), phenol, and pyridine. Results show that an increase in the relative biomass/air ratio, a decrease in
temperature, and higher steam content lead to a higher tar production. BTEX have been found as the
main constituents of tars (60e70% wt.), whereas among PAHs, the most signicant are the lighter ones
(naphthalene, acenaphtylene, acenaphtene). Phenol production is favoured at lower temperatures or/and
higher steam content (around 50% wt. of the tar mixture at 750  C and steam gasication). A progressive
aromatisation of tars has been observed when increasing the temperature, the effect of steam addition
on tar composition being not signicant at temperatures higher than w1000  C.
2012 Elsevier Ltd. All rights reserved.

Keywords:
Biomass gasication
Tar characterisation
Temperature
Relative fuel/air ratio
Steam content

1. Introduction
Gasication is a complex thermochemical process in which
a carbonaceous solid fuel (coal, biomass, wastes) is transformed at
high temperatures (700e1500  C) and in the presence of a gasifying
agent (air or oxygen under sub-stoichiometric conditions, steam,
carbon dioxide, hydrogen, and mixtures of them) into a gas with
a useful heating value, called producer gas or synthesis gas. The key
advantage of gasication is the possibility of converting a solid fuel
into a gas (easier to clean, transport and burn efciently) which
keeps 70e80% of the chemical energy of the original fuel [1].
Moreover, gas from gasication can be used in a wide range of
applications: production of heat and power, and as feedstock for
the synthesis of fuels and chemicals (the latter in the cases of
oxygen and/or steam gasication) [2e7]. In the case of biomass
gasication, the advantages associated with the use of biomass (an
abundant, widespread renewable energy source) are added into
those of the gasication technology. However, during the gasication process, part of the fuel (around 20%, although the amount
depends on the operating conditions and the type of gasier) is not
* Corresponding author. Tel.: 34 926 295 300x3880.
E-mail address: JuanJose.Hernandez@uclm.es (J.J. Hernndez).
0360-5442/$ e see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.energy.2012.12.005

transformed into producer gas, but remains as carbonaceous solid


residue (char and soot particles), and as a condensable product
(tars).
Tars are the most troublesome pollutants of producer gas, being
the main technical hurdle for the commercial implementation of
biomass gasication [1,8,9]. Tars consist of a complex mixture of
organic compounds (including aromatic and heteroaromatic
species as well as polycyclic aromatic compounds, PAHs) with high
boiling points, which is produced during the devolatilisation stage
[9,10,11]. The main technical problems caused by biomass gasication tars include tar condensation (which causes plugging and
fouling problems, thus giving place to unwanted plant shutdowns),
tar polymerisation at high temperatures (which produces polycyclic compounds and even soot in extreme cases), the need for
managing hazardous residual efuents derived from wet cleaning
systems, and catalyst deactivation due to tar deposition [8,12].
The gravimetric amount of tars produced is claimed not to be
enough to assess the performance of the gasication process.
Rather, the criterion of the quality of the product gas and the
efciency of the gas cleaning system that is increasingly receiving
attention is the tar dew point, dened as the temperature at which
condensation is produced when cooling the gas. Tar dew point
mainly depends on tar composition, in particular on the amount of

334

J.J. Hernndez et al. / Energy 50 (2013) 333e342

Table 1
Thermochemical characterisation of dealcoholised marc of grape.
Fuel

Dealcoholised marc of grape

Ultimate analysis (% wt., dry basis)


C (%)
H (%)
N (%)
S (%)
O (%)
Proximate analysis (% wt., dry basis)
Volatile matter
Fixed carbon
Ash
LHV (MJ/kg, dry basis)
Stoichiometric fuel/air ratio,
Fstoic (dry, ash-free basis)

52.06
5.75
2.05
0.14
32.06
64.02
28.04
7.94
20.41
0.14

Tar properties have been evaluated through several parameters:


individual production of the tar components, TPi (expressed in
mg/Nm3 of product gas), total tar production, TP (in mg/Nm3), tar
dew point (in  C) and tar carcinogenic potential, KE (in mg/Nm3).
These parameters have been dened in Section 2.3 and, as
mentioned before, they affect the selection of the gas cleaning
system, the nal use of the gasication gas and the management of
the efuents from tar removal techniques. Therefore, one of the
most interesting features of this work is the extensive and
combined analysis of the tar properties (including a detailed
speciation) and the gasication performance when changing the
operating conditions of a drop-tube gasier.
2. Experimental

heaviest compounds [13,14], while tar composition and amount


depend on several factors, namely the nature of the initial feedstock
(coal, biomass) [9,10,15], fuel properties (in particular, moisture
content [11,16e19]), the operating conditions (temperature, pressure, residence time) [8,9,12,14,20e25], the type of gasier [17], the
gasifying agent used [12,15,22,26,27], the use of catalysts, and the
method of tar sampling and analysis [8,9,10,20].
In this work, the effect of the main operating conditions of the
gasication process (relative fuel/air ratio, operating temperature
and composition of air and steam in the gasifying agent) on the
production and properties of the resulting tars has been analysed.
Air and steam are inexpensive, extensively used gasifying agents in
small- and medium-scale gasiers for energy production. With this
purpose, an analytical method based in liquid chromatography
(HPLC) for the detection of the main species present in gasication
tars has been used. HPLC is acknowledged as a feasible technique
for the identication of both polar and non-polar compounds obtained from biomass pyrolysis/gasication [28e30]. However, the
use of HPLC for the detection of tars produced in biomass gasication is less extended, since the most important off-line methods
for tar sampling (the guideline method, and the solid phase
adsorption method, SPA) use gas chromatography coupled to mass
spectrometry (GCeMS) [14,31e36]. The HPLC method developed is
able to detect 16 PAHs compounds, BTEX, phenol, and pyridine,
these compounds being considered as the most abundant at hightemperature gasication [35]. Moreover, the PAHs detected (listed
in Table 4) have also been speciated for being the group pointed by
the US Environmental Protection Agency (EPA) as priority pollutants due to their higher carcinogenic potential. Phenol and pyridine
can be considered as representative of oxygen-containing and
nitrogen-containing heteroatomic compounds, respectively.

2.1. Biomass feedstock and experimental schedule


The biomass fuel used in this work has been dealcoholised marc
of grape, an agro-industrial residue obtained in the winery and
distillery industries from the water washing of marc of grape,
material consisting of skins, stalks and seeds of grape. This feedstock has been selected because of its high energy potential in the
central and southern regions of Spain. Table 1 contains the thermochemical characterisation of dealcoholised marc of grape.
2.2. Gasication facility and test procedure
All the tests performed have been carried out in a small-scale
drop-tube gasication pilot plant located at the laboratories of
the Fuels and Engines Group of the University of Castilla-La Mancha
(GCM-UCLM), described elsewhere [37e39] and shown in Fig. 1.
The reactor consists of an alumina tube (60 mm internal diameter,
1.2 m long). Given the small dimensions of the gasier, it is
necessary to externally heat it in order to compensate for heat
losses. Therefore, the tube is surrounded by an electrical furnace
with three independent temperature zones, which allows the
separate analysis of the operating conditions (unlike usual
commercial auto-thermal facilities, where fuel/air ratio and
temperature are variables closely related). A steam production
system, composed of a water pump provided with a frequency
controller and a 3 kW steam generator allows one to select and
keep a constant steam ow up to 6.7 L/h. In addition, a cold trap
(consisting of a sealed stainless steel container surrounded by
a water jacket at room temperature) installed at the gas outlet
enables to collect the condensates (water, tars and entrained
particles) contained in the product gas. Lastly, a gas metre makes

Table 2
Experimental tests carried out for tar sampling and analysis.
Fuel

T ( C)

_ f (kg/h)
m

_f
m

Dealcoholised
marc of grape

1200

0.61
1.33
2.04
1.4
1.11
1.11
1.105
0.99
0.90
0.925

0.51
1.14
1.74
1.19
0.94
0.94
0.94
0.84
0.76
0.79

750
750
1050
1200
750
1050
1200

_ f : fuel mass ow rate in kg/h.


m
_ f daf : dry, ash-free fuel mass ow rate in kg/h.
m
_ s : steam mass ow rate in kg/h.
m
_ a : air mass ow rate in kg/h.
m
YH2 O : steam content in the gasifying agent in % wt.
Frg : relative fuel/air ratio (with respect to the stoichiometric one).

daf

(kg/h)

_ a (kg/h)
m

_ s (kg/h)
m

2.4

2.4
1.08

0
0.9

1.6

YH2 O (% wt.)
0

0
45.4

100

Frg
1.53
3.38
5.17
3.54
6.23
6.24
6.22
e
e
e

J.J. Hernndez et al. / Energy 50 (2013) 333e342

335

Table 3
List of detected compounds in the HPLC analytical methods developed.
Group

Compound name

Abbreviation

Standard
concentration
(mg/ml)

TEF [37]

PAHs

Acenaphtene
Acenaphtylene
Anthracene
Benz(a)anthracene
Benz(b)uoranthene
Benz(k)uoranthene
Benz(g,h,i)perylene
Benz(a)pyrene
Chrysene
Dibenz(a,h)anthracene
Fluoranthene
Fluorene
Inden (1,2,3-cd)pyrene
Naphtalenea
Phenanthrene
Pyrene
Benzene
Toluene
Ethylbenzeneb
Xylene (o-, m-, p-)
Phenol
Pyridine

Ac
Acy
Anth
BaA
BbF
BkF
BghiPe
BaPy
Chy
DahA
FluA
Fluo
InPy
Naph
Phe
Pyr
Ben
Tol
Et-Ben
Xyl
Phen
Pyrd

1000
1000
100
100
200
100
200
100
100
200
200
200
100
1000
100
100
2000
2000
2000
2000 (each one)

0.001
0.001
0.001
0.1
0.1
0.1
0.01
1
0.01
1
0.001
0.001
0.1
0.001
0.01
0.001
e
e
e
e
e
e

BTEX

Heteroatomic compounds
a
b

1 mL, pure

Coeluted with ethylbenzene.


Coeluted with naphthalene.

possible to obtain the ow rate of the non-condensable fraction of


the product gas ow. The dry-basis gas composition is determined
on-line every 2 min by a gas micro-chromatograph VARIAN CP4900 PRO located at the end of a sampling line.
The experimental procedure in the gasication tests was as
following: the biomass, previously ground and homogenized below
0.5 mm, was weighed and loaded into the feeding hopper. After
reaching the set temperature point in the electrical furnace, and
having adjusted the air and steam ows according to the experimental schedule shown in Table 2, the screw feeder was turned on.
In all tests the steam inlet temperature was set at 400  C, and the
maximum steam ow rate has been 1.6 kg/h. During the test, both

the gas composition and the gas outlet temperature were periodically registered. Once the operation was stabilised, several values
of the gas metre were registered in order to calculate the average
producer gas ow rate. After 20e30 min, when the gas composition
detected by the micro-GC (in dry basis) was steady, the test was
nished and the duration of the test was registered. All ows
(biomass, air and steam) were stopped. Both the ash-char hopper
and the cold trap were taken out, their contents being weighed (in
order to perform mass balances) and sampled (the method for
sampling and analysis of tars is detailed in Section 2.3).
For the overall study of the gasication process, it is necessary to
carry out an analysis of both the process performance and the tar

Table 4
Effect of the operating temperature and the gasifying agent on the speciation of tars.
Group

Compound

TPi (mg/Nm3)
Air

PAHs

BTEX

Heteroatomic
compounds
N.D.: not detected.

Naph Et-Ben
Acy
Ac
Fluo
Phe
Anth
FluA
Pyr
BaA
Chy
BeBF
BeKF
BaPy
DahA
BghiPe
InPy
Ben
Tol
Xyl (o- m p-)
Phen
Pyrd

54.6 Air/45.4 steam (% wt.)

Steam

T 750  C

T 1200  C

T 750  C

T 1050  C

T 1200  C

T 750  C

T 1050  C

T 1200  C

297.2  0.6
248.6  6.8
96.0  1.2
88.2  0.2
77.4  0.3
25.5  0.6
75.1  31.6
43.9  0.4
14.8  1.2
10.8  0.3
11.75  1.34
4.37  1.25
0.81  0.48
7.08  0.73
5.94  0.3
4.7  0.7
5.9  2.7
4.7  2.7
2598.6  0.6
1584.74  0.6
N.D.

211.3  0.6
220.2  0.2
42.2  1.2
15.7  0.3
44.9  0.1
11.0  0.8
39.3  0.8
30.0  0.3
6.4  1.0
2.7  1.4
0.61  0.76
1.66  0.17
0.68  0.53
2.71  0.40
3.09  0.9
1.6  1.9
3.1  2.8
1.6  0.5
1847.6  0.6
162.4  1.1
N.D.

302.2  1.5
249.6  1.0
381.1  1.7
67.9  1.2
47.3  2.4
16.2  1.9
61.2  12.9
60.0  2.2
15.7  84.9
6.2  43.0
19.9  21.8
4.8  36.8
19.9  61.6
4.8  76.3
2.1  48.7
6.1  73.5
3739.5  6.6
1377.0  4.8
2642.8  1.5
5883.5  0.6
N.D.

1039.2  1.1
684.3  5.8
201.9  7.5
185.2  0.9
229.9  0.1
62.7  0.5
152.8  0.2
94.9  2.6
33.9  21.2
19.6  2.5
19.0  9.40
8.3  19.7
19.0  19.0
8.3  20.8
2.2  14.8
17.9  2.5
1633.4  14.0
297.7  22.7
9087.6  1.4
1574.3  0.5
N.D.

763.8  4.8
531.0  8.3
159.2  20.7
57.3  72.7
119.6  3.2
33.3  5.1
77.4  5.2
52.0  8.3
11.8  8.9
9.7  9.6
11.5  21.7
8.1  36.0
11.5  27.0
8.1  48.5
1.4  85.3
24.4  86.2
1000.8  20.5
738.2  60.1
6679.3  4.8
962.3  0.3
N.D.

7418.2  6.0
10782.5  4.1
5525.5  9.6
593.1  29.9
2078.0  0.5
515.0  0.5
1524.9  1.7
1125.5  2.1
343.4  33.0
230.5  35.5
300.4  9.6
155.3  42.8
300.4  71.2
155.3  5.0
21.2  15.3
578.5  5.6
566.3  21.9
331.3  22.1
64872.2  6.0
189914.9  0.6
N.D.

4374.3  0.9
2209.1  0.3
701.3  3.9
920.4  0.5
792.4  0.7
234.5  0.7
501.8  0.2
305.3  1.1
139.1  1.8
67.6  2.9
70.9  6.6
39.5  13.3
70.9  10.5
39.5  28.9
6.9  33.5
69.2  15.7
52.5  3.4
35.4  6.3
38253.7  0.9
7343.7  0.5
N.D.

2784.9  1.6
1875.3  1.5
1773.2  9.0
238.8  0.3
525.7  0.5
137.3  0.5
467.1  0.6
253.8  0.9
81.5  18.8
52.8  4.5
52.2  1.0
29.8  8.2
52.2  13.6
29.8  35.7
3.2  11.8
160.4  2.2
86.6  4.2
52.1  10.7
24354.2  1.6
3239.7  0.3
N.D.

336

J.J. Hernndez et al. / Energy 50 (2013) 333e342

Fig. 1. Biomass entrained-ow gasication pilot plant.

production. The former issue is determined through several gasication parameters presented in previous works [38]: producer gas
net heating value (LHVpg, in MJ/kg), gas yield (GY, in Nm3/kgdaf,
dened as the dry, N2-including gas ow rate produced for every
kilogramme of dry, ash-free fuel introduced), and cold gas efciency (CGE, in %, dened as the chemical energy content of the
product gas compared to that of the original solid fuel).
Table 2 sums up the gasication tests performed in this work,
_ f daf is the dry, ash_ f is the fuel mass ow rate in kg/h; m
where m
_ a represent the steam
_ s and m
free fuel mass ow rate in kg/h; m
and air mass ow rates in kg/h, respectively; YH2 O is the steam
content in the gasifying agent in %wt., and Frg is the relative fuel/air
ratio (dened with respect to the stoichiometric one). Although
some operating conditions (in particular, high steam content and
low temperature) lead to an excessive tar production and thus are
not realistic conditions of commercial gasication plants, it is worth
remarking that the objective of this work is not tar removal, but an
extensive comparison of tar properties (composition, technical and
environmental hazard) when modifying the gasier operating
conditions, for which extreme values have also been evaluated. In
addition, the low residence times of reactants within this smallscale reactor leads to lower fuel conversion and higher tar
production than those typical of commercial drop-tube gasiers.
2.3. Tar sampling and analysis
The procedure used to sample and analyse the tars produced in
the gasication process is as follows: the content of the cold trap is
weighed and extracted after each gasication test. This fraction
contains water, some tars and particles dissolved. Next, the trap is
rinsed three times with 500 mL of 2-propanol (solvent recommended by the guideline method) [35]. This second fraction
contains non-polar tars and some remaining particles dissolved in
2-propanol. After both fractions are separately ltered and their
volume measured, they are homogenised and stored in dark bottles

at 4  C. The lters containing the solid particles are Soxhlet


extracted with 250 mL of 2-propanol for 6e8 h, until the solvent
drops are clear. The volume of the liquid fraction obtained after
the Soxhlet extraction is measured, and added into the rest of
sample, whereas the particles extracted are oven-dried at 105  C
for 3e4 h, and then weighed. The volume and temperature of the
total tar sample are measured after being homogenised. 100 mL of
the total tar sample are introduced in an atmospheric distillation
facility in order to determine the gravimetric tar production.
Simultaneously, 1 mL of the tar sample is microltered before
being introduced into vials and HPLC analysed. Analysis vials
made of dark glass are kept in the freezer in order to prevent tar
polymerisation until they are completely HPLC analysed. An HPLC
Shimadzu system with ultravioletevisible detector was used for
the determination of the tar composition. Two different analytical
methods (each one with a different chromatographic column) were
developed. The rst one was used for the analysis of PAHs and BTEX,
whereas the second method detects phenol and pyridine. After the
analytical method for PAH and BTEX detection is started up and all
tar samples obtained from gasication tests are properly sampled
and HPLC analysed, the HPLC column is replaced and the analytical
method for phenol and pyridine detection is developed and calibrated. After the method is ready to use, all the analysis samples
(which, in the meanwhile, had been stored in dark vials at very low
temperature) are injected again according to the analytical conditions of each column. PAHs and BTEX analytical method uses
a SUPELCO Analytical 5 mm 250  4.6 mm column, a mobile phase of
water/acetonitrile at a ow rate of 1.5 mL/min, the analysis
temperature and the analysis time being 25  C and 50 min respectively. On the other hand, phenol and pyridine analysis is carried out
with a Supelcosil LC8 3 mm 150  4.6 mm, at 30  C and with a mobile
phase of water/methanol/acetic acid (1 mL/min), the analysis time
being 25 min. The detector wavelength has been 254 nm in both
methods. As recommended in [35], the elapsed time between
sampling and HPLC analysis did not exceed a month in any case.

J.J. Hernndez et al. / Energy 50 (2013) 333e342

337

Table 3 contains the compounds detected, as well as the name


abbreviations used in the gures, the concentration of every individual compound in the standard mixtures, and the values of their
toxicity factor (TEF). The standards mixtures injected for calibration
purposes were the following:
- A PAHs standards mixture (SUPELCO EPA 610 Polynuclear
Aromatic Hydrocarbons Mix 100e2000 mg/mL methanol:
chloroform 1:1, 1 mL).
- A certied BTEX standard mixture (HC BTEX Mix 2000 mg/mL in
methanol, 1 mL). The mixture contained benzene, toluene,
ethyl-benzene and o-, m-, and p-xylene.
- Individual standards of benzene, toluene, and o-, m-, and
p-xylene (5000 mg each one) in order to identify the order or
the peaks detected in the BTEX standard mixture.
- A certied phenols standard (SUPELCO EPA Phenols Mixture
500e2500 mg/L in methanol, 1 mL).
- An individual standard of phenol (1000 mg).
- A standard of pure pyridine (1 mL).
Calibration of the HPLC system was carried out using the standards at different concentrations (350, 300, 200, and 50 mL of
standard mixture diluted in acetonitrile or methanol). Each calibration sample was injected three times in order to check the
analysis repeatability.
The correlation factor obtained after the calibration procedure
was R2 0.999 for all the PAH compounds, except for anthracene
and chrysene, with lower but admissible correlation factors
(R2 0.995 and R2 0.989, respectively). In the case of BTEX
calibration, correlation factors ranged from R2 0.966 (p-xylene) to
R2 0.998 (toluene). Concentrations lower than 0.3 mg/mL were not
considered by the software (lower detection limit of the HPLC
system). It is worth mentioning that both ethyl-benzene and
naphthalene co-elute, thus, the detection peak corresponds to the
sum of both compounds.
As for the analysis of the tar obtained from gasication tests,
each tar sample was analysed three times in order to check the
repeatability of the test. Calibration was controlled before each
analysis introducing a standard sample. In case any compound was
wrongly detected, the calibration process was repeated.
Tar production was evaluated through several parameters:
- Individual production of tar compounds detected by liquid
chromatography, TPi (expressed in mg/Nm3 product gas).

Fig. 2. Effect of the relative fuel/air ratio (Frg) on the total tar production.

Fig. 3. Effect of the relative fuel/air ratio (Frg) on the composition of the tars produced:
(a) PAHs, and (b) BTEX and phenol.

TPi

iHPLC $Vtotal $103


V_ pg $ttest

(1)

being [i]HPLC the concentration of the individual compounds from


HPLC, in mg/mL, obtained from the chromatographic analyses; Vtotal
the measured volume of total tar sample in mL; V_ pg the product gas

Fig. 4. Effect of the relative fuel/air ratio (Frg) on the tar dew point and the carcinogenic potential of the tars produced.

338

J.J. Hernndez et al. / Energy 50 (2013) 333e342

ow rate in Nm3/h, measured from the gas metre; and ttest the
measured duration of the gasication test in h.
- Total tar production, TP (in mg/Nm3, calculated as the sum of
the individual productions).

TP

TPi

(2)

- Tar dew point (in  C), calculated from the complete model
developed and validated by the ECN [13].
- Tar carcinogenic potential, KE (in mg/Nm3), calculated from the
tar production of the individual PAH compounds and their
respective factors of toxicity, TEF, as proposed by Zorn et al. [40]
(see Table 3):

KE

TPi $TEFi

(3)

being in this case the subscript i each one of the PAH compounds
detected.
3. Results and discussion
3.1. Effect of the relative fuel/air ratio
The relative fuel/air ratio, Frg (or inversely, the equivalent ratio
ER), which is dened as the fuel/air mass ratio respect to the stoichiometric one, is considered one of the main operating parameters
of the air- and air-steam gasication processes [18,39,41e44], since
it inuences fuel conversion, the heating value of the product gas
and tar production.
For this section, the experimental gasication tests were carried
out using air as gasifying agent. The operating temperature
(measured as the external reactor temperature) was kept constant
at 1200  C, typical value of entrained-ow gasiers. The relative
fuel/air ratio Frg was modied by adjusting the fuel ow rate in
order to keep as constant as possible the space residence time
(which has a signicant effect on the fuel conversion). Frg values
tested range from Frg 1.5 (value close to a combustion process) to
Frg 5.2 (high value near pyrolysis conditions), including an
intermediate value typical of gasication processes (Frg 3.4). The
results obtained in tar characterisation are displayed in Figs. 2e4.
In Fig. 2 it can be observed that tar production increases nonlinearly with the relative fuel/air ratio, rising from 1 g/Nm3 gas
(at Frg 1.5) to 15.3 g/Nm3 (Frg 5.2). Similar values were obtained

by Narvez et al. [41]. The non-linear increase occurs for all the
families considered in the tar: PAHs, BTEX, and phenol (it is worth
remarking that pyridine was not detected in any test). Moreover, it
can be checked that when increasing the relative fuel/air ratio from
1.5 to 3.4, the contribution of BTEX in the total tar amount suffers an
increase, whereas both PAHs and phenol slightly decrease.
However, beyond Frg 4, an increase in the relative fuel/air ratio
does not inuence tar distribution. Moreover, it is worth remarking
that the BTEX group is the most abundant in the tar mixture,
accounting for 60e70% of the total tar production, whereas phenol
contributes in all cases with less than 10% of the total amount. PAHs
keeps around 20e30% in the range of Frg tested.
Fig. 3 details the effect of the relative fuel/air ratio on the tar
composition (PAHs, BTEX, and phenol). Although it had been
already observed in Fig. 2 that BTEX is the most signicant group in
tars, among them xylenes are by far the most abundant
compounds, reaching values around 10000 mg/Nm3, whereas
benzene and phenol production are of the same magnitude. The
most abundant PAHs produced are by far lighter ones such as
naphthalene and acenapthylene, whose production is similar to
benzene and phenol, around 1000 mg/Nm3 at Frg 5.2. Similar
naphthalene production was obtained by Rapagn et al. [30],
although in their work, xylenes were not analysed by HPLC and
toluene was the most abundant tar compound. PAHs higher than
chrysene have a very low concentration in the mixture. Lastly, in
Fig. 3 it can be appreciated that there is a non-linear increase of all
the tar components with Frg. These results are consistent with the
literature, since a higher availability of oxygen in the process
(higher when reducing the Frg values) was found to favour the
destruction via oxidation of the volatile matter released in the
pyrolysis stage [9,12,14,26].
Fig. 4 details the effect of the relative fuel/air ratio on the tar dew
point and the carcinogenic potential. Both parameters are closely
related to the tar composition, particularly to the production of
polycyclic aromatic hydrocarbons. It can be observed that the
increase in Frg leads to a slight rise in the dew point from 174  C to
210  C, and to an exponential boost in the carcinogenic potential for
Frg values higher than 3.4, the latter being caused by the non-linear
increase of heavier PAHs (Fig. 3).
However, in order to determine the optimal operating conditions, it is necessary to take also into account the overall effect of the
relative fuel/air ratio on the gasication process performance. With
that purpose, Fig. 5 plots the effect of Frg on the heating value of the
producer gas (LHVpg), the gas yield (GY) and the cold gas efciency
(CGE). When increasing Frg, the relative amount of air available in
the process decreases, and thus char gasication reactions occur at

Fig. 5. Effect of the relative fuel/air ratio (Frg) on the gasication performance: product gas heating value (LHVpg), gas yield (GY), and cold gas efciency (CGE).

J.J. Hernndez et al. / Energy 50 (2013) 333e342

339

a higher extent compared to oxidation reactions. Consequently, the


heating value of the product gas increases. Nevertheless, as Frg rises,
the process is increasingly approaching a pyrolysis; a higher amount
of fuel remains unreacted, thus decreasing the gas yield (GY). The
opposite trends of gas heating value and fuel conversion are
responsible of the existence of a maximum in the cold gas efciency
curve at intermediate Frg values.
As a conclusion, the operation at intermediate Frg values
(Frg w 3e4) leads to a trade-off between gasication performance
(product gas heating value, gas yield) and tar properties (relatively
low tar production and carcinogenic potential values).
3.2. Effect of the operating temperature and steam content in the
gasifying agent
As mentioned in Section 1, temperature is one of the main
operating parameters affecting tar production and composition.
Actually, temperature inuences all the stages taking place during
the thermochemical process of solidegas conversion: fuel devolatilisation, and char gasication reactions. For this section, the effect
of temperature (measured as the external temperature of the
reactor) has been evaluated using three different gasifying agents:
air (YH2 O 0% wt.), steam (YH2 O 100% wt.), and a mixture of air
and steam (YH2 O 45.4% wt.), as shown in Table 2. This latter
composition was selected within the range of optimal conditions
determined in a previous study [45]. In all tests, the fuel/gasifying
agent (air steam) mass ratio was kept as constant as possible,
around 0.5, in order to study the role of air and steam while keeping
constant the space residence time.
Fig. 6 plots the effect of temperature on the total tar production
TP for the three gasifying agents tested. It can be checked that the in
the case of using pure air or pure steam, the amount of tars
generated in the process is reduced as the temperature increases,
which indicates a higher extent in the thermal cracking and steam
reforming reactions. Moreover, the decrease in tar production in
the case of steam gasication seems to follow a non-linear trend,
since the reduction of tar production is much more notable when
increasing the temperature from 750  C to 1050  C than an increase
in temperature from 1050  C to 1200  C. However, in the case of
gasication with a mixture of air and steam (YH2 O 45.4% wt.),
a peculiar trend has been obtained: the reduction of tar production
starts to be signicant only at temperatures higher than 1050  C,
being the amount of tars produced similar at 750  C and 1050  C.
The reasons behind this behaviour must be further studied, but
they might be related to the combined effect of air and steam [45].
On the other hand, it is worth observing that there is a nonlinear increase of tar production when adding steam into the
gasifying agent. This behaviour might be related to the roles of air
and steam when introduced together into the process. The
combustion reactions associated to air prevent the temperature
within the reactor to drop excessively, as well as to contribute to the
production of CO and CO2 (via combustion, partial combustion
and Boudouard reactions), so that the net effect of air is the
improvement of devolatilisation, tar cracking and gasication
endothermic reactions. On the other hand, steam mainly promotes
the steam reforming of char and tars and the water-gas shift
reaction (thus enhancing H2 production), as pointed by several
authors [15e17,45e50]. When increasing the steam content (thus
decreasing the air proportion), several factors (namely, the reduction of the extent of the faster combustion reactions associated to
air, the strongly endothermic nature of steam reforming reactions,
and the presence of excess steam that does not take part in the
reactions) lead to a drop in the temperature within the gasier, as
observed by other authors [47,48,50e52], and therefore, tar
production increases.

Fig. 6. Comparison of the effect of the operating temperature and the gasifying agent
on the production of tars.

In Fig. 6 it can also be appreciated that BTEX constitute the most


abundant fraction in the tar mixture; nevertheless, as the steam
content and/or the operating temperature decrease, the proportion
of phenol in the tar samples is increasingly favoured (as shown in
Fig. 7). Thus, in the most extreme case (steam gasication at
750  C), phenol is observed to account for almost half of the tars
produced. In addition, the total tar production is clearly boosted

340

J.J. Hernndez et al. / Energy 50 (2013) 333e342

Fig. 7. Effect of the operating temperature on the evolution of tar composition.

when increasing the steam content of the gasifying agent. Therefore, both the temperature and the gasiyng agent inuence not
only the amount of tars, but also their composition. These results
are consistent with those found in the literature. As stated by other
authors, lighter hydrocarbons and oxygenated compounds are
transformed at higher temperatures into heavier, more stable PAHs,
which are precursors of soot [8,9,12,14,20e25]. Therefore, the
severity of the gasication process must be selected as a trade-off
between the extent of the cracking reactions and the composition
of the remaining tars (more difcult to destroy via catalytic techniques and precursors of soot) [9,14,20].
Table 4 provides detailed information on the effect of temperature on tar composition for the three gasifying agents tested (the
standard deviation of each component concentration is also shown
in the table). It can be seen that, in the case of air gasication, the
most abundant compounds at 750  C are benzene, xylenes and
phenol, whereas at 1200  C the most representative compounds of
the tar mixtures are xylenes. PAHs are the least abundant
compounds regardless of the operating temperature, keeping in
all cases below 300 mg/Nm3. Moreover, the production of light
PAHs (naphthalene, acenaphtylene) is signicantly higher than the
concentration of the rest of polyaromatic compounds detected. As
found in Fig. 6, it can be conrmed the reduction of the production of
all the groups present in the tar mixture when temperature
increases.
As for the case of a mixture of air and steam as gasifying
atmosphere (central columns of Table 4), it is worth remarking the
existence of non-linear trends: although compounds such as
benzene, phenol and acenaphtene decrease at higher temperatures,
most of the species detected present a maximum (e.g. xylenes,
naphthalene, acenaphtylene, uoranthene, phenanthrene, anthracene) or a minimum (toluene) at intermediate temperatures. As

previously observed in the air gasication process, not only is tar


production higher, but also tar composition seems to be affected by
increasing temperature: whereas at 750  C the most important
compounds are phenol and benzene, at 1200  C xylenes are by far
the most abundant species. Moreover, at high temperatures,
naphthalene concentration becomes of the same magnitude as
benzene, toluene and phenol. Lastly, it is worth noticing that, unlike
the results obtained in air- and steam gasication separately, higher
temperatures when gasifying with an air-steam mixture promote
the production of heavier PAHs such as dibenz(a,h)anthracene,
benz(g,h,i)perylene and inden(1,2,3-cd)pyrene (which might be
responsible for the increase in the carcinogenic potential of tars
produced from air-steam gasication at higher temperatures, as
observed in Fig. 8).
Lastly, in the speciation results obtained in the case of steam
gasication and shown in Table 4, it is remarkable the extremely
high production of phenol at low temperatures, which exceeds
180 g/Nm3, and which is by far the most abundant species of the tar
sample obtained at 750  C. Phenol is followed from benzene and
xylenes, being PAHs the least abundant compounds present in tars.
However, an increase in temperature leads to a dramatic decrease of
all the species detected in the tars. Moreover, in some cases (phenol,
benzene, toluene, and most of the PAHs) the decrease clearly follows
a non-linear trend. At 1200  C, the most abundant compounds are
xylenes. Production of PAHs is signicantly less important, being
lighter species such as naphthalene, acenaphtylene or acenaphtene
the most abundant among the polyaromatic compounds. Therefore,
it is evident that temperature affects both tar production and
composition. These results are consistent with those found in
literature: as already discussed in Section 1, higher temperatures
causes an increase in the extent of the reactions of thermal cracking,
thus destroying primary tars. However, the remaining tars are

Fig. 8. Effect of the operating temperature and the gasifying agent on the tar dew point and the carcinogenic potential of tars.

J.J. Hernndez et al. / Energy 50 (2013) 333e342

341

Fig. 9. Effect of the operating temperature on the gasication performance: product gas lower heating value (LHVpg) and cold gas efciency (CGE).

increasingly transformed into more refractory polyaromatic


aromatic compounds (the so-called tar maturation process) reported by Milne et al. [9]. Moreover, soot formation (secondary char)
via gas-phase nucleation mechanisms is favoured at high temperatures [23].
Fig. 7 plots the evolution of tar composition in function of the
operating temperature and the gasifying agent used. The results
observed in previous gures can be conrmed, namely, the
progressive aromatisation of tars when increasing the temperature
regardless of the gasifying agent used (an increase of the proportion of PAHs and BTEX, as well as a decrease in the content of
phenol). Nevertheless, it is interesting to observe both a higher
sensitivity of tar composition with temperature as the steam
content increases, and the effect of steam to be signicant only at
temperatures lower than 1000  C, the latter being maybe due to the
higher rate of steam reforming reactions at high temperatures
(comparable to that of combustion reactions).
Lastly, Fig. 8 plots the effect of temperature on the level of
technical and environmental hazard of tars produced, quantied by
the tar dew point and the carcinogenic potential, respectively. It can
be seen that in the cases of air- and steam gasication, higher
temperatures cause a slight decrease in the tar dew point, being
more notable in the case of steam gasication when increasing the
temperature from 750  C to 1050  C. On the contrary, combined
air-steam gasication leads to inverse trends, increasing the tar
dew point with the operating temperature. This behaviour, as
discussed previously in Table 4, might be due to the increase in
heavier PAHs, which exceeds the reduction in the overall tar
production. The same reason explains the increase in the carcinogenic potential KE, which again shows opposite trends in the case
of air-steam gasication. It is worth remarking that the toxicity of
tars produced from steam gasication (KE w 200e800 mg/Nm3 in
the range of temperatures tested) is dramatically higher than in
case of using air in the gasication process (keeping in the latter
case always below 40 mg/Nm3), although the reduction of KE with
temperature is more notable when using steam. This behaviour
might point out again at the roles of molecular oxygen O2 as
oxidiser of volatile matter released in the pyrolysis stage, and of
temperature as promoter of thermal cracking reactions. Therefore,
in autothermal gasiers the introduction of oxygen promotes the
tar destruction via two simultaneous mechanisms: the increase of
the extent of the oxidation reactions of volatile matter and the
associated increase of temperature.
The previously discussed results on tar production must be
compared with the effect of temperature on the gasication
performance, displayed in Fig. 9. It can be rstly observed that the

addition of steam leads to an increase of the product gas quality


owing to a less dilution in N2 from air, as well as to the modication
of the extent of the gasication reactions (steam reforming and
WGS reactions are enhanced, thus favouring the production of CO
and H2, and therefore, increasing gas heating value). Higher
temperatures lead to an overall improvement of the gasication
process regardless of the gasifying agent used. However, there is
a range of temperatures (T < 900  C), in which the cold gas efciency of air-steam gasication exceeds the values obtained in pure
steam gasication. On the contrary, at temperatures above 900  C,
the process efciency increases with the steam content of the
gasifying agent.
4. Conclusions
In this work, two HPLC analytical methods have been developed
for the detection of the most important compounds present in
biomass gasication tars: PAHs, BTEX, phenol, and pyridine. This
technique has been used for the determination of the tar composition in an experimental study of the effect of the main operating
conditions of a biomass entrained-ow gasication process. The
results obtained show that:
- An increase of the relative fuel/air ratio Frg leads to a non-linear
rise of the tar production due to the lower amount of available
oxygen to oxidise the volatile matter released from the fuel.
Moreover, a progressive increase in the BTEX fraction of tars
has been found when increasing Frg from 1.5 to 3.4. However,
taking into account the overall effect on the gasication
process, it is recommended to operate at intermediate Frg
values as a trade-off between gas quality, process efciency,
fuel conversion, and tar properties (production, and carcinogenic potential).
- Higher operating temperatures reduce the total amount of tars
produced; nevertheless, it takes place a progressive aromatisation of the remaining tars, as observed by the reduction in
the phenol proportion, and the increase in the PAH and BTEX
fractions. Both tar carcinogenic potential and dew point are
reduced (the latter in a lesser extent) when increasing
temperature for air- and steam gasication.
- The combination of high steam contents and low temperatures
favours the production of phenol in tars. Moreover, the use of
steam increases the sensitivity of tar composition to changes in
the operating temperature.
- The introduction of steam in the gasifying agent at a constant
temperature leads to a non-linear increase of tar. However, it

342

J.J. Hernndez et al. / Energy 50 (2013) 333e342

does not signicantly affect tar composition at high temperatures (higher than 1000  C). Therefore, as a trade-off between
process performance and tar production, it is recommended to
operate with air-steam mixtures with YH2 O  50% wt.
Acknowledgements
The Ministry of Education and Science of the Government of
Castilla-La Mancha is gratefully acknowledged for their nancial
support through the GENERBIO Research Project (Reference POII100128-01789). We also are grateful to ALVINESA for the biomass
supply. G. Aranda is indebted to the Spanish Ministry of Education,
Culture and Sport for a FPU Scholarship (ref. AP2007-02747).
References
[1] Higman C, van der Burgt M. Gasication. 2nd ed. GPP-Elsevier; 2008.
[2] Brown RC. Biomass reneries based on hybrid thermochemical-biological
processing e an overview. In: Kamm B, Gruber P, Kamm M, editors.
Bioreneries e industrial processes and products. Status quo and future
directions, vol. 1. Wiley-VCH Verlag GmbH & Co.KGaA; 2006 [chapter 11].
[3] Milne T, Elam C, Evans R. Hydrogen from biomass. State of the art and research
challenges. International Energy Agency (IEA/H2/TR-02/001), http://www.
ieahia.org/pdfs/hydrogen_biomass.pdf; 2002 [last accessed in October 2012].
[4] Sims R, Taylor M, Saddler J, Mabee W. From 1st- to 2nd-generation biofuel
technologies. An overview of current industry and RD&D activities. International Energy Agency [last accessed in October 2012], http://www.iea.org/
publications/freepublications/publication/2nd_Biofuel_Gen.pdf; 2008.
[5] Ma Z, Zhang Y, Zhang Q, Qu Y, Zhou J, Qin H. Design and experimental
investigation of a 190 kWe biomass xed bed gasication and polygeneration
pilot plant using a double air stage downdraft approach. Energy 2012;46:
140e7.
[6] Nzihou A, Flamant G, Stanmore B. Synthetic fuels from biomass using
concentrated solar energy e a review. Energy 2012;42:121e31.
[7] Karamarkovic R, Karamarkovic V. Energy and exergy analysis of biomass
gasication at different temperatures. Energy 2010;35:537e49.
[8] Gil J. El problema de los alquitranes en la gasicacin de biomasa. Infopower:
Actualidad y tecnologa de produccin y uso eciente de la energa 2005;79:
88e94 [in Spanish].
[9] Milne TA, Evans RJ, Abatzoglou N. Biomass gasier tars: their nature,
formation, and conversion. National Renewable Energy Laboratory; 1998.
NREL/TP-570e25357.
[10] Li C, Suzuki K. Tar property, analysis, reforming mechanism and model for
biomass gasication e an overview. Renew Sust Energ Rev 2009;13:594e604.
[11] Reed TB, Das A. Handbook of biomass downdraft gasier engine systems.
Solar Energy Research Institute, US Department of Energy. SERI/SP-271e3022,
http://taylor.ifas.u.edu/documents/Handbook_of_Biomass_Downdraft_
Gasier_Engine_Systems.pdf; 1988 [last accessed in October 2012].
[12] Devi L, Ptasinski KJ, Janssen FJJG. A review of the primary measures for tar
elimination in biomass gasication processes. Biomass Bioenerg 2003;24:
125e40.
[13] Thersites. The ECN tar dew point site. http://www.thersites.nl [last accessed in
October 2012].
[14] van Paasen SVB, Kiel JHA. Tar formation in a uidised-bed gasier. ECN-Ce04013, ftp://nrg-nl.com/pub/www/library/report/2004/c04013.pdf; 2004 [last
accessed in October 2012].
[15] Vreugdenhil BJ, Zwart RWR. Tar formation in pyrolysis and gasication. ECNEe08-087, http://www.ecn.nl/docs/library/report/2008/e08087.pdf; 2009
[last accessed in October 2012].
[16] Franco C, Pinto F, Gulyurtlu I, Cabrita I. The study of reactions inuencing the
biomass steam gasication process. Fuel 2003;82:835e42.
[17] McKendry P. Energy production from biomass (part 3): gasication technologies. Bioresour Technol 2002;83:55e63.
[18] Mevissen N, Schulzke T, Unger CA, Mac an Bhaird S. Thermodynamics of autothermal wood gasication. Environ Prog Sustain Energ 2009;28(3):347e54.
[19] van Loo S, Koppejan J, editors. The handbook of biomass combustion and
co-ring. Earthscan; 2008.
[20] Di Blasi C. Combustion and gasication rates of lignocellulosic chars. Prog
Energ Combust 2009;35:121e40.
[21] Adegoroye A, Paterson N, Li X, Morgan T, Herod AA, Dugwell DR, et al. The
characterisation of tars produced during the gasication of sewage sludge in
a spouted bed reactor. Fuel 2004;83:1949e60.
[22] Han J, Kim H. The reduction and control technology of tar during biomass
gasication/pyrolysis: an overview. Renew Sust Energ Rev 2008;12:397e416.
[23] Neves D, Thunman H, Matos A, Tarelho L, Gmez-Barea A. Characterization
and prediction of biomass pyrolysis products. Prog Energ Combust 2011;37:
611e30.
[24] Ponzio A, Kalisz S, Blasiak W. Effect of operating conditions on tar and gas
composition in high temperature air/steam gasication (HTAG) of plastic
containing waste. Fuel Process Technol 2006;87(3):223e33.

[25] Zwart R, van der Heijden S, Emmen R, Bentzen J, Ahrenfeldt J, Stoholm P, et al.
Tar removal from low-temperature gasiers. ECN-Ee10-008, http://www.ecn.
nl/docs/library/report/2010/e10008.pdf; 2010 [last accessed in October 2012].
[26] Kinoshita CM, Wang Y, Zhou J. Tar formation under different biomass gasication conditions. J Anal Appl Pyrol 1994;29:169e81.
[27] Gil J, Corella J, Aznar M, Caballero M. Biomass gasication in atmospheric and
bubbling uidized bed: effect of the type of gasifying agent on the product
distribution. Biomass Bioenerg 1999;17:389e403.
[28] Bahng M-K, Mukarakarate C, Robichaud DJ, Nimlos MR. Current technologies
for analysis of biomass thermochemical processing: a review. Anal Chim Acta
2009;651:117e38.
[29] Dez MA, lvarez R, Gayo F, Barriocanal C, Moinelo SR. Study of the
composition of tars produced from blends of coal polyethylene wastes
using high-performance liquid chromatography. J Chromatogr A 2002;945:
161e72.
[30] Rapagn S, Gallucci K, Di Marcello M, Matt M, Nacken M, Heidenreich S, et al.
Gas cleaning, gas conditioning and tar abatement by means of a catalytic
lter candle in a biomass uidized-bed gasier. Bioresour Technol 2010;101:
7123e30.
[31] Grootjes AJ. Tar measurement by the solid phase adsorption (SPA) method.
ECN-Le11-064,
http://www.ecn.nl/docs/library/report/2011/l11064.pdf;
2011 [last accessed in October 2012].
[32] Brage C, Yu Q, Chen G, Sjstrm K. Use of amino phase adsorbent for biomass
tar sampling and separation. Fuel 1997;76(2):137e42.
[33] Dufour A, Girods P, Masson E, Normand S, Rogaume Y, Zoulalian A. Comparison of two methods of measuring wood pyrolysis tar. J Chromatogr A 2007;
1164:240e7.
[34] BTG Biomass Technology Group. Tar and tar measurement http://www.
btgworld.com/en/rtd/analysis/tar-and-tar-measurement [last accessed in
October 2012].
[35] CEN/TS 15439:2006. Biomass gasication e tar and particles in product gases
e sampling and analysis; 2007.
[36] Masson E, Ravel S, Thiery S, Dufour A. Tar analysis by solid phase adsorption
(SPA) associated with thermal desorption (TD) and gas chromatography (GC)
analysis, international Workshop Measurement, analysis and monitoring of
condensable gas components (especially tar) in product gases from biomass
gasication and pyrolysis at 19thEU Biomass Conference and Exhibition, ICC
Berlin (2011) http://www.evur.tu-berlin.de/leadmin/fg45/Workshops/Tar_
Workshop/05_Masson_crittbois_France_Workshop_Berlin_2011_06_08.pdf
[last accessed in October 2012].
[37] Hernndez J, Aranda-Almansa G, Bula A. Gasication of biomass wastes in an
entrained-ow gasier: effect of the particle size and the residence time. Fuel
Process Technol 2010;91:681e92.
[38] Hernndez J, Aranda-Almansa G, Serrano C. Co-gasication of biomass wastes
and coal-coke blends in an entrained-ow gasier: an experimental study.
Energ Fuel 2010;24:2479e88.
[39] Lapuerta M, Hernndez JJ, Pazo A, Lpez J. Gasication and co-gasication of
biomass wastes: effect of the biomass origin and the gasier operating
conditions. Fuel Process Technol 2008;89(9):828e37.
[40] Zorn C, Khler M, Weis N, Scharenberg W. Proposal for assessment of indoor
air polycyclic aromatic hydrocarbon (PAH), proceedings: indoor air; 2005. p.
2535e40.
[41] Narvez, Oro A, Aznar MP, Corella J. Biomass gasication with air in an
atmospheric bubbling uidized bed. Effect of six operational variables on the
quality of the produced raw gas. Ind Eng Chem Res 1996;35:2110e20.
[42] Lv P, Yuan Z, Ma L, Wu C, Chen Y, Zhu J. Hydrogen-rich gas production from
biomass air and oxygen/steam gasication in a downdraft gasier. Renew
Energ 2007;32:2173e85.
[43] Zhao Y, Sun S, Tian H, Qian J, Su F, Ling F. Characteristics of rice husk
gasication in an entrained ow reactor. Bioresour Technol 2009;100:
6040e4.
[44] Roll H, Hedden K. Entrained ow gasication of coarsely ground Chinese reed.
Chem Eng Process 1994;33:353e61.
[45] Hernndez JJ, Aranda G, Mendoza JM. Effect of steam content in the
air-steam ow on biomass entrained ow gasication. Fuel Proces Technol
2012;99:43e55.
[46] Prins M, Ptasinski K, Janssen F. Thermodynamics of gas-char reactions: rst
and second law analysis. Chem Eng Sci 2003;58:1003e11.
[47] Campoy M, Gmez-Barea A, Vidal F, Ollero P. Air-steam gasication of
biomass in a uidised bed: process optimisation by enriched air. Fuel Process
Technol 2009;90:677e85.
[48] Lucas C, Szewczyk D, Blasiak W, Mochida S. High-temperature air and steam
gasication of densied biofuels. Biomass Bioenerg 2004;27:563e75.
[49] Yan F, Luo S, Hu Z, Xiao B, Cheng G. Hydrogen-rich gas production by steam
gasication of char from biomass fast pyrolysis in a xed-bed reactor: Inuence of temperature and steam on hydrogen yield and syngas composition.
Bioresour Technol 2010;101:5633e7.
[50] Chen W, Annamalai K, Ansley RJ, Mirik M. Updraft xed bed gasication of
mesquite and juniper wood samples. Energy 2012;41:454e61.
[51] Lv PM, Xiong ZH, Chang J, Wu CZ, Chen Y, Zhu JX. An experimental study on
biomass air-steam gasication in a uidized bed. Bioresour Technol 2004;95:
95e101.
[52] Gordillo G, Annamalai K, Carlin N. Adiabatic xed-bed gasication of coal,
dairy biomass, and feedlot biomass using an air-steam mixture as an oxidizing
agent. Renew Energy 2009;34:2789e97.

También podría gustarte