You are on page 1of 8

Fuel 92 (2012) 346–353

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Gasification reactivity of char from dried sewage sludge in a fluidized bed


Susanna Nilsson ⇑, Alberto Gómez-Barea, Diego Fuentes Cano
Chemical and Environmental Engineering Department, Escuela Superior de Ingenieros (University of Seville), Camino de los Descubrimientos s/n, 41092 Seville, Spain

a r t i c l e i n f o a b s t r a c t

Article history: The gasification reactivity of char from dried sewage sludge (DSS) applicable to fluidized bed gasification
Received 19 May 2011 (FBG) was determined. The char was generated by devolatilizing the DSS with nitrogen at the selected
Received in revised form 17 July 2011 bed temperature and was subsequently gasified by switching the fluidization agent to mixtures of CO2
Accepted 18 July 2011
and N2 (CO2 reactivity tests) and steam and N2 (H2O reactivity tests).. The tests were conducted in the
Available online 31 July 2011
temperature range of 800–900 °C at atmospheric pressure, using partial pressure of the main reactant
in the mixture (CO2 or H2O) in the range of 0.10–0.30 bar. Expressions for the intrinsic reactivity (free
Keywords:
of diffusion effects) as a function of temperature, partial pressure of gas reactant (CO2 or H2O) and degree
Gasification
Kinetics
of conversion were obtained for each reaction. For the whole range of conversion it was found that the
Sewage sludge char reactivity in an H2O–N2 mixture was roughly three times higher than that in a mixture with the cor-
Char responding partial pressure of CO2. The reactivity was only influenced by particle size greater than
Modeling 1.2 mm in the tests with steam at 900 °C. It was demonstrated that the method of char preparation
greatly influences the reactivity, highlighting the importance of generating the char in conditions similar
to that in FBG.
Ó 2011 Elsevier Ltd. All rights reserved.

1. Introduction have a great influence on the reactivity of char. For fuels with high
volatile content, such as biomasses and wastes, the structure of
The kinetics of char gasification is a key aspect when designing char generated is strongly affected by devolatilization conditions.
a fluidized bed gasifier (FBG) because the reactions of char with Also, deactivation of chars exposed to high temperatures has been
CO2 and H2O are slow compared to devolatilization and gas phase observed [21]. The variation of the reactivity during conversion can
reactions. In most cases, conversion of char is not completed in be taken into account by using structural profiles determined
FBG, thus reducing the process efficiency. experimentally. Different structural profiles from literature applied
The reactivity of char depends on the type of fuel and on how to various types of char have recently been reviewed in [1]. The
the char is prepared. Extrapolation of char reactivity data from most important parameters affecting the properties of the char
one fuel to another is, therefore, questionable [1]. A great deal of generated after devolatilization are the temperature and particle
research exists detailing measurements of the char gasification heating rate [22]. The composition of the fluidizing gas used during
reactivity of a variety of fuels, including coal [2–9] and biomass devolatilization is assumed to only slightly affect the char gener-
[10–13]. Comparatively less work exists on the gasification of char ated because the high flow of volatiles released from the solid par-
from contaminated biomass, residues and wastes [14–19]. Little ticle makes penetration of the fluidizing gas difficult [1].
research has dealt with the gasification of char from dried sewage In commercial FBG units, mm-sized particles are used, the tem-
sludge (DSS) [14,18]. It was found that char from DSS had higher perature is in the range of 750–900 °C and the heating rate at
reactivity with CO2 than chars from coal and car tires [14]. The rea- which the fuel is devolatilized is in the order of 100–1000 °C/s.
son may be the effect of catalytic compounds [2,11,12,20], which Then, to obtain reactivity useful for application in FBG units, the
are found in large proportions in DSS char. However, it is difficult char should be generated at a high temperature and heating rate.
to conclude general trends based on existing char gasification or In laboratory studies these conditions are often not fulfilled. More-
combustion tests from DSS found in the literature because the over, there are other processes that make it difficult to prepare the
chemical and physical characteristics of DSS may vary significantly char in the laboratory to be useful for FBG. A significant reduction
depending on its origin. in char reactivity after cooling the char has been reported [2,6].
It is known that the method of char generation and preparation However, in most kinetics studies the char generated in the labora-
affects the reactivity [1]. Pore distribution and internal surface tory is cooled down to room temperature before conducting the
char gasification tests [5,7–9,13–15,18].
In the present work, the reactivity (or conversion rate), r is de-
⇑ Corresponding author. Tel.: +34 954 482163; fax: +34 954 461775.
fined as:
E-mail address: sln@esi.us.es (S. Nilsson).

0016-2361/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.fuel.2011.07.031
S. Nilsson et al. / Fuel 92 (2012) 346–353 347

Nomenclature

a, b, c fitting parameters in Eq. (7), – rx reactivity at reference conversion, s1


De effective diffusivity, m2/s rp particle radius, m
dp particle diameter, m T temperature, K
Ea activation energy, J/mol t time, s
F(x) function that expresses the variation of reactivity with umf minimum fluidization velocity, m/s
char conversion, – x char conversion, –
k0 preexponential factor, barn s1
mC mass of carbon in char at any time, g Greek letters
mC0 initial mass of carbon in char produced after devolatil- d variable for integration, –
ization, g g global effectiveness factor, –
n reaction order, – ge external effectiveness factor, –
pCO2 partial pressure of CO2 in the feed gas during CO2 gasi- gi internal effectiveness factor, –
fication experiments, bar s characteristic time for intraparticle mass transfer, s
pH2 O partial pressure of H2O in the feed gas during H2O gas-
ification experiments, bar Abbreviations
R ideal gas constant, 8.314 J/(mol K) DSS dried sewage sludge
r reactivity, s1 FB fluidized bed
r30 reactivity at x = 0.30, s1 FBG fluidized bed gasification or fluidized bed gasifier

1 dmC dx tivity of char from DSS for both gasification with CO2 and H2O (in
r¼ ¼ ð1Þ
mC0 dt dt mixtures where only a single reactant, CO2 or H2O, is present) based
on the form given in Eqs. (2) and (3), assessing the effects of differ-
mC0 and mC being the mass of carbon in the char at initial time and ent factors and assumptions on the expressions obtained.
at conversion x, respectively. To obtain the reactivity of a char pre-
pared from a given fuel in a mixture of CO2–N2 (reactivity with CO2)
or in a mixture of H2O–N2 (reactivity with H2O), at temperature T, 2. Experimental
the following expression can be used to fit the measurements [1]:
2.1. Apparatus and materials
r ¼ r x ðT; pCO2 =H2 O Þ FðxÞ ð2Þ
Experiments have been carried out in a laboratory fluidized bed
where rx is the reactivity at reference state of conversion x. rx ac-
(FB) reactor. The experimental setup is represented in Fig. 1. The
counts for the dependence of reactivity on temperature and partial
reactor is made of stainless steel. It has a preheating section, an
pressure of gas reactant (pCO2 or pH2 O ), whereas F(x) is a function tak-
FB section with a 51 mm internal diameter and a freeboard section
ing into account the variation of reactivity with conversion [1]. pCO2
with an 81 mm internal diameter. The reactor is surrounded by a
stands for the partial pressure of CO2 in the CO2–N2 mixtures used
10 kW electrical oven and is equipped with four thermocouples
during CO2 reactivity tests and pH2 O is the partial pressure of H2O in
and two controllers, allowing the control of temperature in the
the H2O–N2 mixtures used during H2O reactivity tests. The approx-
bottom bed and freeboard.
imation assumed in Eq. (2) by which the reactivity can be written as
the product of two functions rx (T, pCO2 =H2 O ) and F(x) has been shown
DSS
to be valid within a specific range of operating conditions (espe-
cially within a specific temperature interval) [1,23].
The gasification reactions are governed by surface processes so
Langmuir–Hinshelwood kinetics [14,24] has shown to represent
the char reactivity [25], which accounts for the observed inhibitory
effects of CO and H2 [23,25]. However, at moderate pressure and Electrical
Gas oven
when the partial pressure of CO and H2 are not very high, the reac- Cleaning
tivity at a given conversion, rx, can be simplified by nth-order kinet-
ics. The validity of this simplification can be verified
T
experimentally by confirming that the order of reaction neither
Gas Analyzer Freeboard
varies with temperature nor with degree of conversion. Assuming P
this simplification to be valid, rx can be represented using the fol-
lowing expression: T
FB

Ea n P
r x ¼ k0 exp pCO2 =H2 O ð3Þ H2O
RT

Eqs. (2) and (3) allow for a simplified and practical expression to fit
Mass
the measurements. However, the validity of the assumptions made
Flow
to derive Eqs. (2) and (3) should be verified. Moreover, the reactivity
Controllers
(kinetic parameters and structural profile) should be obtained with-
out diffusion effects, i.e. intrinsic reactivity. In this way, the reactiv- N2 CO2
ity could be used as part of a char particle model to predict the
conversion in practical equipment (with a variety of char sizes
and operating conditions). In this work, we derive the intrinsic reac- Fig. 1. Experimental setup.
348 S. Nilsson et al. / Fuel 92 (2012) 346–353

Two types of mixtures were used for the char reactivity tests: Table 2
CO2–N2 (CO2 char reactivity tests) and H2O–N2 (H2O char reactivity Content of main metal elements in DSS.

tests). Steam was generated by instantaneously vaporizing a fixed Element mg/g of DSS
flow of water. The steam generated was mixed with the N2 and the Fe 38.91
mixture was fed to the reactor. The flows of N2 and CO2 were ad- Na 38.32
justed by means of mass-flow controllers, whereas the flow of Al 31.51
water was adjusted by a peristaltic pump, which was calibrated Mg 8.92
K 7.81
before each test. The composition of the exit gas was measured Ca 4.79
by a Siemens analyzer using a non-dispersed infrared method for Ni 1.71
CO, CO2 and CH4 and thermal conductivity and paramagnetic Zn 1.40
methods for H2 and O2, respectively.
The proximate and elemental analyses of DSS and the composi-
tion of DSS char are shown in Table 1. The content of the main me- Table 3
tal constituents of DSS is given in Table 2 and the particle size Particle size distribution of DSS.
distribution of DSS is found in Table 3. It can be seen in Table 1 that
Sieve size, mm Mass fraction, %
DSS char contains mainly ash and carbon. The DSS particles were
>5 0.75
employed in the experiments as received, i.e. no size reduction
4–5 0.28
was applied. The bed material employed was bauxite with a size 2.8–4 54.86
in the range of 250–500 lm and density of 3300 kg/m3. The mini- 2–2.8 39.44
mum fluidization velocity measured for the bauxite was 0.20 m/s 1.4–2 3.70
and the terminal velocity (calculated for the mean particle size, 1–1.4 0.71
0.5–1 0.21
i.e. 375 lm) was 4.3 m/s. The mass ratio of DSS to bauxite in the
<0.5 0.05
bed was close to 1/100 in all the tests, so the fluid-dynamics of
the bed is assumed to be governed by the bauxite.

mixing in both the gas feed and exit lines. In the first blank test,
2.2. Experimental procedure a certain flow of CO2 was injected into the fluidization gas (pure
N2) in a port situated in the gas feed line. In the second blank test,
First, the reactor was heated by setting the test temperature in the injection of CO2 was made in a port situated in the upper part
the oven. During the heating period a continuous flow of air was of the FB. The effect of mixing was well characterized by a first or-
fed. Once the desired temperature was reached, the fluidizing gas der model, with time constants equal to 1.9 s and 8.1 s for the feed
was switched to N2 and when no more oxygen was detected by and exit lines, respectively.
the analyzer, a batch of DSS was fed through a pipe that ends near
the bed surface. When the CO, CO2, CH4 and H2 concentrations 2.3. Operating conditions
measured by the gas analyzer were nearly zero, devolatilization
was considered to be complete. Then the flow of CO2 or H2O was Table 4 summarizes the most important experimental parame-
turned on, while the flow of N2 was adjusted to set the desired ters and operating conditions. Tests were conducted at three differ-
composition of the gasification mixture (CO2–N2 or H2O–N2) and ent temperatures: 800, 850 and 900 °C and three partial pressures
fluidization velocity. Gasification conditions were maintained until of CO2 (CO2 reactivity measurements) and H2O (H2O reactivity
the CO and CO2 concentrations at the exit were too low to allow measurements) in the fluidizing gas: 0.10, 0.20 and 0.30 bar. The
accurate measurements. This occurred at different conversions of pressure inside the reactor during the experiments was below
char, x = 0.60–0.85, depending on the operating conditions. After 1.05 bar. The gas velocity, during both devolatilization and gasifi-
gasification, the gas feed was switched to air in order to burn the cation of char, was set to three times the minimum fluidizing
remaining char. velocity. With this value, entrainment was avoided, no slug flow
The effect of gas mixing was taken into account to correct the
data of gas concentration measured during the char reactivity
Table 4
tests. Two blank tests were performed to assess the effects of gas Main experimental parameters.

Rig
Table 1 Bed diameter 51 mm
Proximate and elemental analyses of DSS and composition of DSS Freeboard diameter 81 mm
char. Material
Bed material Bauxite,
DSS DSS char
250 lm < dp < 500 lm
LHV, MJ/kg 11.18 – Mass of bauxite in the bed 300 g
HHV, MJ/kg 12.25 – DSS batch size 3–8 g
DSS particle sizes tested* 1.2 mm, 2.4 mm,
Composition, wt.%
3.4 mm and 4.5 mm
Moisture 8.65 –
Operating parameters
Dry basis
Temperature 800–900 °C
Ash 43.15 84.14
Total pressure 1 bar
Volatiles 51.75 –
Minimum fluidization velocity (bauxite) 0.20 m/s
Fixed carbon 5.10 –
Nominal gas velocity 3 umf
C 30.88 13.17
Fluidizing gas during devolatilization N2
H 4.36 0.63
Partial pressure of CO2/H2O in the feed gas 0.10–0.30 bar
N 4.76 1.05
(mixtures of CO2–N2 or H2O–N2) during char
S 1.24 1.01
gasification
O* 15.61 0
*
* Averages of the ranges shown in the particle size distribution shown in Table 3.
Calculated by difference.
S. Nilsson et al. / Fuel 92 (2012) 346–353 349

was detected and the char particles were assumed to be suffi- the char reactivity test in the FB. These tests are called ex situ char
ciently well mixed with the bed material. Four particle size ranges tests, to distinguish them from the other tests performed in this
were studied (in mm); 1–1.4 (average 1.2), 2–2.8 (average 2.4), study, where devolatilization and char gasification were made
2.8–4 (average 3.4) and 4–5 (average 4.5). Together these ranges sequentially, i.e. in situ char tests. A comparison of the reactivity
include more than 98% of the as-received DSS (see Table 3). Batch tests using in situ and ex situ char is shown in Table 5, where
sizes between 3 g and 8 g were employed. The batch size was ad- the reactivity at x = 0.2 is shown at two temperatures. The reactiv-
justed in order to ensure that the concentrations of the product ity is found to be much higher when using in situ generated char.
gases at the exit were kept below 1%. The results in Table 5 show that the temperature has a stronger
influence on the reactivity of ex situ char. An analysis of the data
2.4. Data treatment shown indicates a higher activation energy for the reaction of car-
bon with steam using ex situ char compared to that using in situ
During char gasification tests with CO2–N2 mixtures, the follow- char. The results show that the thermal history of the char is
ing reaction occurs: important for the gasification reactivity and that char generated
in situ should be employed to simulate FBG. This fact should also
C þ CO2 ¢ 2CO ð4Þ be kept in mind when comparing the results of this work to other
and the amount of reacted carbon in the char was calculated research where the char is generated in external devices, for in-
from the measured CO in the exit gas. During the char gasification stance, laboratory apparatus like ovens or TGA. In laboratory de-
tests with H2O–N2 mixtures, the following reactions occur: vices the char is cooled down before the reactivity test. In
addition, the heating rate and temperature under which the char
C þ H2 O ¢ CO þ H2 ð5Þ is generated may also differ from that in an FB. Consequently,
the results may not be directly applicable to FBG. All the results
CO þ H2 O ¢ CO2 þ H2 ð6Þ presented below have been conducted using in situ char in an at-
tempt to reproduce the reactivity that would exist in an FBG.
therefore, the amount of reacted carbon in the char was calcu-
The variation of size and shape of DSS particles during devola-
lated from the measurements of CO and CO2 in the exit gas. The gas
tilization has been studied elsewhere [27]. The results showed that
concentrations measured were corrected taking into account the
the particles maintained their shape and that shrinkage was not
mixing of gas in the exit line, characterized by the blank test de-
important. Similar observations have been made by other authors
scribed above.
[28]. Specific surface area has been measured for chars obtained
At the beginning of each test there is a transient period before
from DSS of different particles sizes and using two different char
the composition of the gas surrounding the char particle reaches
preparation methods; (i) in an FB at 800 °C and (ii) in an oven
the steady state value. Once this value has been reached, it remains
applying a low heating rate (5 °C/min) and a final temperature of
constant throughout the test. Using the time constant obtained in
800 °C. It was found that the particle size did not influence the spe-
the blank test for the mixing in the gas feed line it can be calculated
cific surface area of the char, but the char preparation method did
that it takes approximately 6 s for the concentration of CO2 or H2O
to some extent: the BET surface area was 41.1 and 35.4 m2/g for
in the reactor to reach 95% of the steady state value. In addition,
char generated in FB and in oven at low heating rate, respectively.
there is a transition time for the reactive gas in the emulsion of
These values are low compared to those measured for biomass
the FB to reach the internal surface of the char particle. This time
chars like wood, having BET surface area over 300 m2/g. The results
is estimated by the intra particle gas diffusion: s ¼ r 2p =De , De being
also indicate that the properties of char from DSS depend on the
the effective diffusivity and rp the char particle radius. Taking
char preparation method. Around 50% of the surface area of DSS
De = 1.05  104 m2/s [26] and rp = 2.5 mm (the maximum particle
char was present in micropores, a low proportion compared to
size of DSS used, assuming that the char particle size equals that of
those measured for biomass and coal, where micropores represent
DSS) gives s below 0.1 s, which is negligible compared to the delay
practically 100% of the total surface area.
caused by gas mixing. Then a transient period of 6 s was assumed.
Char gasification experiments were carried out at 850 °C and
Taking into account that the conversion achieved up to this time in
900 °C using various particle sizes. The reactivity as a function of
the test with the highest reactivity (900 °C using H2O with the
conversion is shown in Fig. 2, reaction with CO2 (using CO2–N2
smallest DSS particles) was almost 0.05, the conversion rates mea-
mixtures) and Fig. 3, reaction with H2O (using H2O–N2 mixtures).
sured during the initial period with x < 0.05 were considered not
The results indicate that the reactivity is only affected by particle
reliable enough so they were not taken into account in the results
size in the tests using H2O at 900 °C, for which the reactivity is low-
presented below.
er when using courser particles, due to mass transport effects. The
influence of particle size is more significant at low conversion as
3. Results and discussion the reactivity is high (Fig. 3b), whereas it becomes smaller at high
conversion. In order to obtain the reactivity free of diffusion effects,
3.1. Assessment of the effect of char preparation method and particle only the tests using 1.2 mm DSS particles were taken to fit the
size measurements. This was the smallest size that could be tested
without risk of entrainment. Since diffusion effects could still be
Prior to the determination of the effects of temperature, gas present in the tests using H2O–N2 mixtures at 900 °C, a theoretical
composition and degree of conversion on the char reactivity, a
number of char gasification tests are analyzed to assess: (i) the
influence of the method of char preparation on the reactivity ob- Table 5
Comparison between the reactivity at x = 0.2, r20  104 (s1) obtained with in situ and
tained, and (ii) the diffusion effects on the reactivity due to mass
ex situ generated char for the reactions with CO2 and H2O at two different
transport limitations, the latter in order to ensure the determina- temperatures.
tion of the intrinsic reactivity.
Temperature, °C In situ char Ex situ char
Various tests were performed using char that was generated in
the same way as that described above (devolatilization of DSS in CO2 800 4.9 2.2
H2O 800 13.3 4.2
the FB with nitrogen at the same temperature) but the resulting
900 67.1 35.2
char was cooled down to room temperature before conducting
350 S. Nilsson et al. / Fuel 92 (2012) 346–353

2.5 3.2. Determination of intrinsic reactivity of char with CO2 and H2O
1.2 mm
2.4 mm In Figs. 4 and 5 the effects of temperature and gas reactant con-
2 4.5 mm centration on r30, the reactivity at reference conversion, x = 0.3, are
studied. In Fig. 4, r30 is presented at different temperatures using
900 ºC fixed partial pressures of CO2 and H2O (pCO2 =H2 O = 0.20 bar),
1.5
r · 103 (s-1)

whereas Fig. 5 shows a comparison between r30 measured at


800 °C using different values of pCO2 and pH2 O . It is shown that
850 ºC the reaction with H2O is roughly three times faster than with
1
CO2 for all temperatures tested. This suggests that both reactions
have similar activation energies.
0.5 In order to verify the ability of the nth-order reaction model to
represent the intrinsic reactivity, i.e. that Eq. (3) can be used to fit
the measurements, the order of reaction for both reactions, with
0 CO2 and H2O, was calculated at a constant temperature at different
0 0.2 0.4 0.6 0.8 conversions. Experiments were carried out at 800 °C varying par-
Conversion, x tial pressures of CO2 and H2O in the feed gas. Assuming nth-order
kinetics with respect to CO2 and H2O, the reaction order was calcu-
Fig. 2. Char reactivity as a function of conversion in a mixture of CO2 and N2 with lated at different conversions, and the results are shown in Table 6.
pCO2 ¼ 0:20 bar, measured for different particle sizes at 850 °C and 900 °C. For the reaction with H2O, the reaction order remains practically
constant throughout the whole range of conversion tested and is
analysis of the influence of diffusion effects for different particle
sizes is presented in Appendix A. It is confirmed that diffusion ef-
fects are negligible when using 1.2 mm particles so the results ob- 6
CO2
tained for this particle size can be used to estimate the intrinsic
H2O
reactivity. 5

4
(a) 10
r30· 103 ( s-1)

1.2 mm
2.4 mm 3
8 3.4 mm
2

6
r ·10 3 (s-1)

0
4
800 850 900
Temperature (ºC)
2
Fig. 4. Comparison of reference char reactivity (at x = 0.3), r30, in CO2–N2 and H2O–
N2 mixtures as a function of temperature at pCO2 ¼ 0:20 bar and pH2 O ¼ 0:20 bar.

0
0 0.2 0.4 0.6 0.8
1.4
CO2
Conversion, x
H2O
1.2
(b) 10 1.2 mm
2.4 mm 1
3.4 mm
8
r30· 103 (s-1)

0.8

6 0.6
r ·10 3 (s-1)

0.4
4
0.2

2
0
0.1 0.2 0.3

0 Partial pressure of gas reactant


0 0.2 0.4 0.6 0.8 (CO2 or H2O) (bar)
Conversion, x
Fig. 5. Reactivity of reference (at x = 0.3), r30, at 800 °C in CO2–N2 and H2O–N2
Fig. 3. Char reactivity as a function of conversion in a mixture of H2O and N2 with mixtures as a function of partial pressure of reactant in the mixture, i.e. CO2 (pCO2 )
pH2 O ¼ 0:20 bar, measured for different particle sizes at: (a) 850 °C; (b) 900 °C. and H2O (pH2 O ).
S. Nilsson et al. / Fuel 92 (2012) 346–353 351

Table 6 parameter, b, in Eq. (7) is not essential to represent the data, but
Reaction order, n, for the reactions of char with CO2 and H2O it was included in order to ensure (r/r30) > 0 at initial time, i.e.
at different conversions, x.
x = 0. To estimate the best values of the parameters a, b and c,
x n (CO2) n (H2O) the procedure described below was applied.
0.1 0.320 0.329 For gasification modeling, it is interesting to have an expression
0.2 0.372 0.316 allowing the accurate calculation of the time required to reach cer-
0.3 0.416 0.338 tain conversions of char. This time can be calculated from the reac-
0.4 0.391 0.318
0.5 0.450 0.356
tivity by integration according to:
0.6 0.452 0.339 Z x
0.7 0.463 0.339
1
t ðxÞ ¼ dd ð8Þ
0 rðdÞ
The parameters a, b and c in Eq. (7) were calculated by mini-
approximately 0.33. For CO2, the change in reaction order with mizing the sum of the accumulated errors of 1/r up to a given
increasing conversion is more significant than for H2O, but the var- conversion for the three temperatures studied and for fixed par-
iation is still small, so an average constant value between 0.4 and tial pressures of CO2 or H2O in the feed gas, pCO2 or
0.45 can be assumed. These values of reaction order are in agree- pH2 O ¼ 0:20 bar. The values of the parameters obtained for the
ment with those obtained by Nowicki et al. [18], who measured reactions with CO2 and H2O are shown in Table 8. A comparison
values of n = 0.39, and n = 0.3 for the reactivities of sewage sludge between the calculated and experimental r/r30 curves is shown in
char with CO2 and H2O, respectively. It has been concluded that the Fig. 7. The r/r30 curves obtained at different temperatures show
nth-order model is a reasonable model to fit the measurements by the same trends with increasing conversion and only some differ-
using Eq. (3), for both CO2 and H2O. The Arrhenius plots for r30 for ence is observed for extreme x values. As can be seen in Fig. 7a
CO2 and H2O are shown in Fig. 6. The corresponding values of k0 and b, for x > 0.15, the r/r30 curves are similar at different temper-
and Ea obtained from the fitting are shown in Table 7. It is con- atures. Moreover, the curves obtained using CO2–N2 mixture and
firmed that the activation energy is similar for both reactions. H2O–N2 mixture are almost identical. This suggests the use of the
To determine the variation of reactivity with conversion, a mod- same expression for F(x) to represent the reactivity with both CO2
el for F(X) (see Eq. (2)) was sought. Various models established in and H2O, simplifying the model of char conversion in an FBG.
the literature were initially applied in an attempt to fit the mea- Therefore, a new set of values of parameters a, b and c to give
surements, such as the Modified Random Pore Model [29] and oth- the best fit for all the r/r30 curves (using both CO2–N2 mixture
ers [1]. However, the data did not fit satisfactorily so an alternative, and H2O–N2 mixture) was calculated. These values are also
empirical expression was used: shown in Table 8, marked as Average.
In summary, an expression for the intrinsic reactivity of DSS
r   char in CO2–N2 mixtures and H2O–N2 mixtures in an FB can be ob-
FðxÞ ¼ ¼ ð1  xÞðax þ bÞ exp cx1=2 ð7Þ
r 30 tained by combining Eqs. (2) and (3), using r30 as reference reactiv-
ity, leading to:
This expression was chosen because it gave good fit of measure-
 
ments for both CO2 and H2O and because it was able to reproduce Ea n
the maximum of the reactivity at low conversions, x < 0.20. The r ¼ k0 exp pCO2 =H2 O FðxÞ ð9Þ
RT

with F(x) given by Eq. (7) with empirical parameters a, b and c from
-4 Table 8, and frequency factor, k0, activation energy Ea, and order of
reaction, n, given in Table 7.
This reactivity was used to calculate the conversion as a func-
-5
tion of time at different temperatures. Two sets of parameters for
H2O representing F(x) were used. The first set of parameters uses the
Ln (r30 (s-1))

-6 values of a, b and c calculated separately for each reaction (see Ta-


ble 8), whereas the second set corresponds to a, b and c as valid for
CO2 both reactions, i.e. using a, b and c marked as Average in Table 8.
-7 Fig. 8 shows the conversion vs. time curves calculated with the
two models of F(x) together with the experimental curves, at the
three temperatures studied with pCO2 or pH2 O ¼ 0:20 bar. Since
-8
the results at x < 0.05 were not reliable, t = 0 was set for x = 0.05.
It is shown that the model exhibits good agreement with the mea-
-9 surements, except at 800 °C for x > 0.6. The use of a common model
8.5 8.7 8.9 9.1 9.3 9.5 of F(x) to represent both the reactivities with CO2 and H2O only af-
1/T · 10 4 (K-1) fects the results slightly, so it is a reasonable simplification for
modeling purposes.
Fig. 6. Arrhenius plot for the reactivity of reference (at x = 0.3), r30, in CO2–N2 and Finally, in a real FBG the gas in the emulsion is a complex mix-
H2O–N2 mixtures with, respectively, pCO2 ¼ 0:20 bar and pH2 O ¼ 0:20 bar.
ture of CO2, H2O, N2, CO and H2, as well as other minor compounds.

Table 8
Table 7 Values of parameters a, b and c in Eq. (7) calculated for each reaction separately and
Values of kinetic parameters: activation energy, Ea, frequency factor, k0, and order of calculated to represent both the reactions with CO2 and H2O (marked average).
reaction, n, for the char reactivity with CO2 and H2O.
a b c
n Ea, kJ/mol k0, barn s1
CO2 51.3 2.9 4.6
CO2 0.41 163.5 6.33  104 H2O 11.5 3.6 3.0
H2O 0.33 171.0 3.90  105 Average 30.8 3.6 4.0
352 S. Nilsson et al. / Fuel 92 (2012) 346–353

(a) 2.5 900 ºC (a) 1

850 ºC 0.9
800 ºC
Normalized reactivity, r/r 30

2 0.8
Calculated 900 ºC 850 ºC
0.7
800 ºC

Conversion, x
1.5 0.6

0.5 Experimental
1 0.4 Calculated
0.3 Calculated Average
0.5 0.2

0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 500 1000 1500 2000 2500 3000
Conversion, x Time (s)

(b) 2.5 900 ºC (b) 1


850 ºC 0.9
800 ºC
Normalized reactivity, r/r 30

2
0.8
Calculated 900 ºC 850 ºC
0.7
800 ºC
1.5 Conversion, x 0.6

0.5
Experimental
1 0.4
Calculated
0.3
Calculated Average
0.5 0.2

0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 200 400 600 800 1000
Conversion, x Time (s)
Fig. 7. Experimental and theoretical (using Eq. (7)) normalized reactivity r/r30 as a Fig. 8. Comparison between experimental and theoretical (using Eqs. (9) and (7))
function of conversion at different temperatures. (a) Reactivity in a mixture of CO2– curves of char conversion versus time. (a) CO2 reactivity tests with pCO2 ¼ 0:20 bar;
N2; (b) reactivity in a mixture of. H2O–N2. and (b) H2O reactivity tests with pH2 O ¼ 0:20 bar.

In this work we have obtained the reactivity of DSS char in a mix-


ture of CO2–N2 and in a mixture of H2O–N2. The reactivity of char both reactions, approximately 170 kJ/mol). The order of reaction
in a mixture containing simultaneously CO2 and H2O is not neces- is approximately 0.33 for H2O and 0.41 for CO2, remaining practi-
sarily the sum of the two pure-gas reaction rates. This question will cally constant with conversion. Comparing the reactivities deter-
be dealt with in a future work. mined to those obtained in tests using ex situ char (the char was
generated in the FB but was cooled down before conducting the
gasification test) demonstrated that the preparation method has
4. Conclusions a significant influence on char reactivity. It has been concluded
that in situ char reactivity tests should be used for simulating FBGs.
The gasification reactivity of char from dried sewage sludge The reactivity of char in a mixture containing simultaneously CO2
(DSS) was measured in a laboratory scale fluidized bed (FB) in and H2O will be dealt with in a future work.
the temperature range of 800–900 °C. The char was generated by
devolatilizing the DSS with nitrogen at the bed temperature and Acknowledgements
subsequently gasifying the resulting char with mixtures of CO2
and N2 (CO2 reactivity tests) and H2O and N2 (H2O reactivity tests). The authors acknowledge the Junta de Andalucía for its financial
Kinetic expressions for calculating the intrinsic reactivity of DSS support in the project FLETGAS. They also wish to express their appre-
char as a function of temperature, partial pressure of reactant in ciation to Elisa López, José Guillermo Claro, Verónica Hidalgo and Juan
the mixture (CO2 or H2O) and degree of conversion were obtained: Carlos Vázquez for their help in the experimental work. The help of Ped-
they are represented by Eq. (9) with F(x) given by Eq. (7) taking the ro Ollero in the management of the project is recognized.
empirical parameters a, b and c from Table 8, and frequency factor,
k0, activation energy, Ea, and order of reaction, n, from Table 7. The Appendix A. Assessment of the influence of mass transfer
variation of reactivity with conversion was similar for tests using limitations on the reactivities with H2O at 900 °C
CO2–N2 mixtures and H2O–N2 mixtures, so a single function F(x)
can be used (the one marked as Average in Table 8). The reactivity As mentioned above, it was not possible to carry out measure-
with H2O was roughly three times faster than with CO2 at all tem- ments with DSS particles smaller than 1.2 mm. To check if the reac-
peratures (similar values of activation energy were calculated for tivities measured for this particle size in the tests at 900 °C were
S. Nilsson et al. / Fuel 92 (2012) 346–353 353

Table 9 [4] Liu H, Luo C, Kaneko M, Kato S, Kojima T. Unification of gasification kinetics of
Effectiveness factors calculated for different particle sizes for the reaction with H2O at char in CO2 at elevated temperatures with modified random pore model.
900 °C, x = 0.30 and pH2 O ¼ 0:20 bar (external, ge; internal, gi and global, g). Energy Fuels 2003;17:961–70.
[5] Matsouka K, Kajiwara D, Kuramoto K, Sharma A, Suzuki Y. Factors affecting
dp, mm ge gi g steam gasification rate of low rank coal char in pressurized fluidized bed. Fuel
Process Technol 2009;90:895–900.
1.2 0.98 0.96 0.95
[6] Liu H, Zhu H, Kaneko M, Kato S, Kojima T. High-temperature gasification
2.4 0.95 0.83 0.79 reactivity with steam of coal chars derived under various pyrolysis conditions
3.4 0.92 0.70 0.66 in a fluidized bed. Energy Fuels 2010;24:68–75.
[7] Chitsora CT, Mühlen HJ, van Heek KH, Jüntgen H. The influence of pyrolysis
conditions on the reactivity of char in H2O. Fuel Process Technol
1987;15:17–29.
affected by mass transfer limitations, the effectiveness factor was [8] Sears JT, Muralidhara HS, Wen CY. Reactivity correlation for the coal char–CO2
calculated taking into account the resistances to mass transport reaction. Ind Eng Chem Process Des Dev 1980;19:358–64.
[9] Mühlen HJ. Finely dispersed calcium in hard and brown coals: its influence on
in the gas film (external) and within the particle (intraparticle or pressure and burn-off dependencies of steam and carbon dioxide gasification.
internal). Fuel Process Technol 1990;24:291–7.
The observed reactivity can be expressed as r = grkin, rkin being [10] Kojima T, Assavadakorn P, Furusawa T. Measurement and evaluation of
gasification kinetics of sawdust char with steam in an experimental fluidized
the intrinsic reactivity (without mass transfer limitations) and g bed. Fuel Process Technol 1993;36:201–7.
the effectiveness factor with g 6 1. The effectiveness factor is cal- [11] DeGroot WF, Richards GN. Influence of pyrolysis conditions and ion-
culated as the product of the external and internal effectiveness exchanged catalysts on the gasification of cottonwood chars by carbon
dioxide. Fuel 1988;67:352–60.
factors, ge and gi, taking into account, respectively, the resistance
[12] DeGroot WF, Shafizadeh F. Kinetics of gasification of Douglas Fir and
to mass transport in the gas film and within the particle, g = gegi. Cottonwood chars by carbon dioxide. Fuel 1984;63:210–6.
The method to calculate ge and gi presented in [1] was applied, tak- [13] Moilanen A, Mühlen HJ. Characterization of gasification reactivity of peat char
in pressurized conditions. Fuel 1996;75:1279–85.
ing the effective diffusivity from [26] and estimating the Sherwood
[14] Scott SA, Davidson JF, Dennis JS, Fennell PS, Hayhurst AN. The rate of
number according to Hayhurst and Parmar [30] in order to get con- gasification by CO2 of chars from waste. Proc Combust Inst 2005;30:2151–9.
servative values, the equivalent length of the particle was assumed [15] Marquez-Montesinos F, Cordero T, Rodríguez-Mirasol J, Rodríguez JJ. CO2 and
to be equal to its diameter. steam gasification of a grapefruit skin char. Fuel 2002;81:423–9.
[16] Gea G, Sánchez JL, Murillo MB, Arauzo J. Kinetics of gasification of alkaline
The effectiveness factors, g, ge and gi calculated for x = 0.30 for black liquor from wheat straw. Influence of CO and CO2 concentration on the
the three particle sizes studied at 900 °C are shown in Table 9. The gasification rate. Ind Eng Chem Res 2004;43:3233–41.
global effectiveness factor calculated for the smallest particle size [17] Gea G, Sánchez JL, Murillo MB, Arauzo J. Kinetics of gasification of alkaline
black liquor from wheat straw. 2. Evolution of CO2 reactivity with the solid
is very close to 1, so the measured reactivity can be assumed to conversion and influence of temperature on the gasification rate. Ind Eng
be free of diffusion effects, i.e. the intrinsic reactivity. This is cor- Chem Res 2005;44:6583–90.
roborated by the fact that the Arrhenius plot gave a straight line [18] Nowicki L, Antecka A, Bedyk T, Stolarek P, Ledakowicz S. The kinetics of
gasification of char derived from sewage sludge. J Therm Anal Calorim
with R2 > 0.99 (see Fig. 6). The effectiveness factors calculated for 2011;104:693–700.
dp = 2.4 mm and dp = 3.4 mm were also in good agreement with [19] Zhu X, Song B, Kim D, Kang S, Lee S, Jeon S, et al. Kinetic study on catalytic
the experimental results, which makes the method applied consis- gasification of a modified sludge fuel. Particuology 2008;6:258–64.
[20] Schumacher W, Mühlen HJ, van Heek KH, Jüntgen H. Kinetics of K-catalysed
tent. Therefore, it can be concluded that: (i) diffusion effects are steam and CO2 gasification in the presence of product gases. Fuel
negligible when using 1.2 mm particles and therefore the results 1986;65:1360–3.
obtained using this particle size can be used to estimate the intrin- [21] Senneca O, Russo P, Salatino P, Masi S. The relevance of thermal annealing to
the evolution of coal char gasification reactivity. Carbon 1997;35:141–51.
sic reactivity, and (ii) intraparticle transport of gas influences the
[22] Gómez-Barea A, Leckner B, Evaluation of char reaction rate in a fluidised bed
reactivity when using particle sizes of 2.4 mm and higher. gasifier: from reactivity determination to reactor simulation. In: Proceedings
It is possible that for very low conversions (when the rate of of the 17th European biomass conference and exhibition; 2009. p. 683–9.
conversion is high), mass transfer limitations could have some [23] Ollero P, Serrera A, Arjona R, Alcantarilla S. The CO2 gasification kinetics of
olive residue. Biomass Bioenergy 2002;24:151–61.
influence. This could explain the difference in the r/r30 curves with [24] Jüntgen H. Reactivities of carbon to steam and hydrogen and applications to
H2O for x < 0.10 at 900 °C compared to the other temperatures. In technical gasification processes – a review. Carbon 1981;19:167–73.
any case, this effect does not have a significant influence on the re- [25] Di Blasi C. Combustion and gasification rates of lignocellulosic chars. Prog Prog
Energy Combust 2009;35:121–40.
sults since the r/r30 curves for different temperatures are practi- [26] Dennis JS, Lambert RJ, Milne AJ, Scott SA, Hayhurst AN. The kinetics of
cally the same for x > 0.10 and Eq. (9) fits the measurements combustion of chars derived from sewage sludge. Fuel 2005;84:117–26.
with H2O at 900 °C very well (see Fig. 8). [27] Gómez-Barea A, Nilsson S, Vidal-Barrero F, Campoy M. Devolatilization of
biomass and waste in fluidized bed. Fuel Process Technol 2010;91:1624–33.
[28] Scott SA, Davidson JF, Dennis JS, Hayhurst AN. The devolatilisation of particles
References of a complex fuel (dried sewage sludge) in a fluidised bed. Chem Eng Sci
2007;62:584–98.
[1] Gómez-Barea A, Leckner B. Modeling of biomass gasification in fluidized bed. [29] Zhang Y, Hara S, Kajitani S, Ashizawa M. Modeling of catalytic gasification
Prog Energy Combust 2010;36:444–509. kinetics of coal char and carbon. Fuel 2010;89:152–7.
[2] Miura K, Hashimoto K, Silveston PL. Factors affecting the reactivity of coal [30] Hayhurst AN, Parmar MS. Measurement of the mass transfer coefficient and
chars during gasification, and indices representing reactivity. Fuel Sherwood number for carbon spheres burning in a bubbling fluidized bed.
1989;68:1461–75. Combust Flame 2002;130:361–75.
[3] Linares-Solano A, Mahajan OP, Walker Jr PL. Reactivity of heat-treated coals in
steam. Fuel 1979;58:327–32.

You might also like