Está en la página 1de 7

Fuel 158 (2015) 263269

Contents lists available at ScienceDirect

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

Establishment of kinetic parameters of particle reaction


from a well-stirred uidized bed reactor
H.D. Haustein , T. Kreitzberg, B. Gvert, A. Massmeyer, R. Kneer
Institute of Heat and Mass Transfer (WSA), RWTH Aachen University, Augustinerbach 6, 52056 Aachen, Germany

a r t i c l e

i n f o

Article history:
Received 13 January 2015
Received in revised form 12 May 2015
Accepted 18 May 2015
Available online 28 May 2015
Keywords:
Kinetic parameters
Reaction rate
Fluidized bed
Char particles
Boudouard
FTIR

a b s t r a c t
A novel method is presented for experimental study of gas-particle reactions, based on realizing a
well-stirred reactor as a small scale uidized bed. This reactor is evaluated against the drop tube reactor
and the thermo-gravimetric analyser. It shows to enable high heat up rates (104 K/s), long timescale
observation (up to several hours), operation with small fuel particles (100 lm) and accurate control
of reaction conditions. Char reaction rates are established from real-time gas product analysis by FTIR
spectroscopy, through a detailed data-analysis procedure. This procedure employs a particle
surface-evolution model and accounts for sampling system signal attenuation. The validity of the
well-stirred conditions is established, and the method is employed for char combustion and gasication.
Highly consistent results for char gasication over a wide range of conditions (T = 8001100 C,
C CO2 = 1976%), are used to demonstrate the establishment of kinetic parameters for an n-th order
approach. Activation energy and order of reaction are found and compare well with the literature.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
A well-stirred reactor provides spatially homogenous, controlled conditions for process and reaction analysis. Therefore, it
might serve as a basis for model development and validation
[1,2]. For example, it can be used to experimentally establish inherent reaction rates of char combustion, required for reliable
reactive-ow CFD simulation [3]. However, practical realization
of such an ideal system is difcult. The present study attempts to
do this using a small-scale Fluidized Bed Reactor (FBR), with focus
on the establishment of kinetic parameters of gas-particle reaction,
under controlled conditions.
An FBR is a commonly used type of reactor, though generally
operated on a larger scale and in continuous-feed mode, for processes such as gasication, bio-oil production or coal and char
combustion (e.g., [46]). Here, this reactor type was chosen and
adapted for kinetic studies, due to the possibility of conducting
long timescale measurements up to reaction completion.
Additionally, it is shown that the signicant mixing of the reactor,
as well as its operation with small fuel batches, promotes

Corresponding author. Tel.: +972 (0)3 640 6515.


E-mail addresses: hermanh@post.tau.ac.il (H.D. Haustein), kreitzberg@
wsa.rwth-aachen.de (T. Kreitzberg), goevert@wsa.rwth-aachen.de (B. Gvert),
massmeyer@wsa.rwth-aachen.de (A. Massmeyer), kneer@wsa.rwth-aachen.de
(R. Kneer).
http://dx.doi.org/10.1016/j.fuel.2015.05.038
0016-2361/ 2015 Elsevier Ltd. All rights reserved.

homogenous and controlled conditions for more reliable kinetic


measurements. This allows for the evaluation of the reaction over
its entirety with well-dened conditions greatly simplifying the
analysis and generality of results. Nevertheless, the FBR also has
some short-comings, which are addressed a few paragraphs
further down. Dedicating an FBR to chemical kinetic parameters
study, places
it
among
other well-known methods:
thermo-gravimetric analyzer (TGA, [7]), drop tube reactor (DTR,
[8]) or its variant an entrained ow reactor (EFR, see Shaddix
and co-workers [9,10]). By comparison, the present method has
several advantages as it combines high heat up rates and high temperatures (characteristic to the DTR) with the strong signal and
ability to observe long timescales (characteristic to the TGA).
Furthermore, accurate control over gas composition and temperature allows reactions to be conducted at well-dened, uniform
conditions. These inherent characteristics also enable examination
of application-relevant conditions, e.g. high heat up rates, intermediate residence times and small fuel particles.
On the other hand, this method also has some limitations: First,
it is not as suitable as the DTR for rapid reactions (time resolution
of ms), and it does not allow direct measurement of the fuel conversion in contrast to the TGA (mass-loss with resolution of lg).
Secondly, the mixing bed does not permit fuel-particle quenching
and extraction, for examination of morphology and composition
evolution, as possible in an EFR or DTR. Finally, the modeling of
the transport processes in an FBR for conversion of results to a

264

H.D. Haustein et al. / Fuel 158 (2015) 263269

generalized/intrinsic form, is generally more complex than for


other methods (Kunii and Levenspiel [11]). A concluding comparison of the FBR to these other methods will be conducted in
Section 3.3.
The uidized bed approach for performing char kinetic studies
was rst explored by Fennel et al. [6]. That study was instrumental
for the methods and the approach employed in the present study.
Indeed, many similarities to that study exist here (similar particle
sizes, sampling rates and relative ow rates 10 times the minimum uidization), though in order to extend the knowledge
obtained there key differences are introduced: (i) Bed diameter is
8 times smaller resulting in a total reactor volume about 200 times
smaller which should increase reactor homogeneity. Larger fuel
batch sizes are used for increased signal to noise ratio and lower
sample size uncertainty; (ii) an extensive uidization and homogeneity study is conducted here to validate the desired conditions;
(iii) rather than the problematic, noise amplifying de-convolution
(reverse) employed there, an iterative convolution (forward, noise
reducing) and tting process is used; (iv) by examining a slower,
endothermic, gasication process errors related to sampling delay
and sampling rate are reduced; (v) although changes in CO2 levels
could not be directly measured for carbon mass closure (whereas
stoichiometry is assumed and only CO is measured here), due to
the advantages listed above the present study has much higher
repeatability and much lower uncertainty. Finally, the chars examined there are of the traditional kind (lignite, bituminous), while
the present study examines an alternative emerging source
woodchar (biogenic fuel).
The reaction process examined here is related to coal combustion and gasication. The complexity of this process can be reduced
by splitting it in two: pyrolysis and char reaction, corresponding to
shorter and longer timescales (see [12,13]). In the present study,
only the slower process of char reaction is examined experimentally, which is characterized by three different regimes [14]:
Regime (I) For smaller particles/low temperatures the reaction is
limited only by the inherent chemical reaction rate (kinetically
controlled) and takes place throughout the accessible particle surface; Regime (II) At increased temperatures (or particle size) it is
controlled by both kinetics and intra-particle pore diffusion and
takes place closer to the outer surface of the particle; Regime
(III) At high temperatures, pressures and particle sizes the reaction
is limited by the transport of reactant-gas to it (diffusion controlled) and reaction takes places only on its outer surface.
In the current study, char reaction rates were established experimentally through real-time FTIR gas sampling and comprehensive
data analysis, under Regime I conditions. Since the presented
method is quite novel, the analysis method is described in detail.
Furthermore, the rst part of the results focuses on the validation
of the desired operating conditions (well stirred). This is followed
by a parametric study of char gasication reaction rates under various conditions. The method and its validation are conducted for
both gasication and combustion, although the full parametric
study and nding of kinetic parameters is done for gasication
alone in this work. In the light of these results, the methods advantages and limitations are re-evaluated and discussed.

2. Experimental method
The experimental setup consists of a small-scale bed of inert
particles, which is uidized by a rate and composition controlled
gas-mixture. Into this, small batches of well-characterized fuel
are introduced, while the products are continuously analyzed from
the exhaust gas. The analysis procedure accounts for the evolution
of the char particle surface during the reaction and the dispersion

caused by the sampling system, to calculate a characteristic reaction rate from measurement data.
2.1. Experimental setup
The experimental system employed here is the same that has
previously been used by the authors and is only briey described
at this point [15,16]. The FBR is located inside a controlled electric
oven, allowing temperatures up to 1280 C to be imposed with high
stability (e.g.: 1000 2 C). The desired composition of the gas mixture can be set from a base gas (air, CO2, N2 or Ar), which can be
enriched with a reactant (O2 or CO2) by independent
temperature-corrected mass ow controllers (MFC), in order to set
ow rate at reactor conditions. This gas mixture heats up as it ows
down the annular gap to the gas distributor (sintered silica glass,
pore diameter range 40100 lm) which uniformly distributes the
ow to uidize an inert bed of round sand-like alumina (Al2O3 diameter dp = 112 30 lm and sphericity /s = 0.80 0.18, established by
laser diffraction analysis and microscopy). A small portion of this gas
mixture is used to purge and mildly pressurize the char injection
(fuel feed) system. For each run a small batch (<25 mg) of pulverized
fuel is dropped onto the bed, where it heats up and reacts with the
uidizing gas. The heating rate has been approximated analytically
taking radiative and convective heat transfer into account. The emissivity of the bed has been calculated adapting the model of Palconok
[17]. The approximation gives values on the order of 104 K/s similar
to the values found by Yu et al. [18]. After complete reaction of the
char, the remaining ash becomes an inert part of the bed material
which is periodically exchanged. The pressure loss over the distributor and uidized bed is measured by differential pressure gauge,
while the bed temperature is measured with an immersed
ceramic-shielded type S thermocouple. The bed has a diameter of
D = 34 mm, with a non-uidized bed height of Hd = 30 mm, and a
typical uidized height uctuating around Hf = 70 mm. The gaseous
reaction products are captured just above the uidized bed. Driven
by a slight reactor overpressure (typically 10 mbar) the exhaust
gas is fed into the gas analyzer through a sampling line and a lter;
afterwards it exits out to a safety venting system. The entire sampling system is heated to 180 C to prevent unwanted tar condensation. A Gasmet DX-2000 FTIR spectrometer, measuring in the mid-IR
range (wave numbers of 6004200 cm1) was employed for
real-time gas analysis, sampling at 0.5 Hz with an accuracy of 2%
of the measurement range after initial calibration.
For additional validation experiments (Section 3.1), some of the
tubes in the FBR were exchanged to measure the temperature and
gas composition at various heights within the bed (details in
Fig. 1). An identically scaled, transparent cold uidized bed was
used for observation of bed uid-dynamics. Therein the pressure
drop across the distributor and bed height was measured as a function of ow rate, under standard air conditions. This system
demonstrated that the char particles are thoroughly mixed into
the bed in less than 2 s.
2.2. Data analysis
The char reactions examined here can be well-described by the
carbon conversion curve, or burnout (the mass fraction of solid
carbon that has reacted). While obtaining the curve from analysis
of exhaust gas-analysis is straightforward, establishing a characteristic reaction rate for the entire conversion requires a more complex analysis procedure: In general, an appropriate char surface
evolution model is used to generate a predicted curve and the reaction rate (control parameter) is found by iterative comparison to
the experimental one.
As the particle is consumed its surface is constantly changing,
and eventually decreases towards complete burnout. By using a

265

H.D. Haustein et al. / Fuel 158 (2015) 263269

Purging gas
Fuel feed
(1)Thermocouple
(2) Sampled gas

Sealings
Fluidizing gas

Fluidized
bed
Distributor

El. heating
Fig. 1. Interior of the uidized bed reactor.

suitable model for particle surface evolution, a characteristic reaction rate covering the entire process can be determined. Several
surface evolution models exist in literature: the uniform reaction
model (URM), stated to be appropriate for Regime I/II conditions
and highly porous chars [11]; the shrinking core model (SCM),
which has recently been shown to be suitable for high-ash/
low-porosity char [19], its offspring the grain model for agglomerated char [20] or the more complex random-pore model [21]. In
this work, due to examination of a highly-porous fuel the URM
was considered and tested for validity. This model (given in Eqs.
(1)) represents a limiting case reaction throughout the volume
of the particle with constant size).

In Eq. (3) Erfc represents the complementary error-function and


parameters a and b represent the sampling system specics. Their
values, a = 4.60 and b = 0.376, were found within 10% by repeated
FTIR measurement of the systems response to a stepwise inlet
input. Fig. 3 shows the systems response to a sudden increase in
CO2 concentration and the tted convolution function for the determination of a and b. The transfer function can be used to determine

These equations are based on the assumption of the initial condition


X(t = 0) = 0, where X is the carbon conversion, t is time and r is the
characteristic reaction rate.
For the investigated conditions, reactions lasted from 15 s up to
several hours. The progression of the reaction was found through
the evolving gas concentrations, measured by the FTIR, from which
the carbon conversion rate was then deduced. Based on a mass balance and the dominant chemical reactions (2C(s) + O2 ? 2CO and
C(s) + O2 ? CO2 or C(s) + CO2 ? 2CO when the CO2 to O2 ratio is
high), the measured changes in the concentration of the carbon
reaction products CO and CO2 relate to the carbon conversion
rate, according to Eq. (2).



_ O2
_i m
@X
MC m
C CO t DC CO2 t


@t mC;0 M i M O2
1  C CO t=2






1
a  2bt
a 2bt
p
p
exp2abErfc
Erfc
2
2 t
2 t

0.1

0.8

0.08

0.6

0.06

0.4

0.04
X (URM)
X (URM, conv.)
X
dX/dt

0.2

0
0

_ indicates the mass ow rate,


where M is the molecular weight, m
and subscripts i and C represent the carrier gas (N2, CO2, etc.) and
carbon (char), respectively. Furthermore, subscript C,0 denotes the
initial carbon-fuel mass (found from the total carbon captured in
the exhaust by FTIR, as explained later). The Parameter C refers to

10

20

30

40

50

60

dX/dt [1/s]

URM : Xt 1  ert

Ft

Carbon Conversion X [-]

(1) Sampled gas


(2) closed

concentration, i.e. the molar fraction of a gas species instantaneously measured by FTIR, and D indicates the difference between
outlet and inlet value.
As a mass balance based on the samples weight did usually not
obtain closure 50% to 80% of the introduced carbon was captured
in the exhaust gas the data was normalized to obtain complete
burnout (X = 1). This is equivalent to the assumption that a percentage of the char does not undergo a reaction in the uidized
bed. This loss may occur by particles sticking to the fuel feed
system or by elutriation of ne particles in the early stages of the
reaction. In any case, complete conversion of the char in the bed
was veried by continued measurement until recovery of steady
levels of CO2 and CO (<50 ppm). This procedure of normalizing
the calculated burnout curves to the detected carbon in the
exhaust has been done by other research groups, (e.g. Lou et al.
[4], who recovered similar fractions) and doesnt affect the measured reaction rate.
The analysis method is demonstrated for the case of a rapid
reaction (combustion at 800 C in air) in Fig. 2. This gure shows
the carbon conversion rate dX/dt, which is determined by means
of Eq. (2). The resulting carbon conversion X is obtained as the integral over time thereof. Before tting a chosen surface evolution
model (e.g. Eq. (1)) to this experimentally found carbon conversion
curve, it is necessary to take dispersion along the sampling line
into account. These effects are related to the long piping (2 m)
and the ne ltration (<2 lm) required by the gas analyzer. To
allow for these issues the carbon conversion, predicted by the surface evolution model, is convoluted by a transfer function F(t) (Eq.
(3)), which is derived from the TaylorAris convectiondiffusion
transport equation. The procedure of (de)convoluting data is
described in detail by Abad et al. and is outlined briey in the following [22]:

0.02

0
70

Time t [s]
Fig. 2. Experimental rapid char reaction in air T = 800 C, CO2 = 21%. Carbon
conversion rate as calculated from Eq. (2) using FTIR measurements and corresponding carbon conversion and its comparison to URM prediction with and
without convolution. Circles stand for the experimentally found carbon conversion
rates dX/dt. Its integral over time, the carbon conversion, is represented by triangles.
The dashed line gives the carbon conversion prediction of the uniform reaction
model, whereas the continuous line expresses its convolution.

H.D. Haustein et al. / Fuel 158 (2015) 263269

dp = 140 20 lm (>80% mass) a compromise between


application-like smaller fuel particles and the lower limits of operation in the current system (entrainment at minimal gas analysis
ow-rate). This diameter was matched to the bed particles
(Archimedes number matching) to promote mixing and prevent
buoyancy-driven separation. As the analysis in Table 1 shows, the
char consists mainly of carbon, with ash and volatiles accounting
for less than 8%.

1
0.8
0.6
0.4
unit-step
step-response
fitted step-response

0
0

10

20

30

40

3. Results
50

3.1. Validation of the method

Time t [s]
Fig. 3. Response of the FBR-System to a sudden increase of CO2 concentration from
0 to 20 Vol.-% in the feed-gas.

the temporal evolution of the systems response R(t) to a unit-step


perturbation S (Eq. (4)):

Rt S  Ft

Considering the predicted carbon conversion curve X(t) as a series of unit-step perturbations at discrete values of t, convolution of
this curve can be carried out by using a linear combination of Eq.
(4):

Rt k St1  Ft k

k1
X
Sti1  t i  Ftk  ti

i1

By substituting S in Eq. (5) with the predicted carbon conversion X, R(tk) gives its convolution at the specic time tk.
After the convolution has been performed, the resulting burnout curve is iteratively t with a least-square regression to the
measured data by changing the parameter r until convergence
has been achieved. The tting was done to the 1080% part of
the carbon conversion curve, to avoid the inuence of reaction
startup and delayed completion. Reaction rates found according
to this procedure, are representative of the majority of the conversion, and not just of a single part of the reaction, as is often the case
in kinetic studies.
The need for the above described convolution procedure is
especially apparent at short timescales on the order of the typical
system delay. By including the convolution procedure, the working
range of this method is extended.
Fig. 2 shows the different curves obtained by the above
described analysis procedure.
2.3. Fuel characterization
For the detailed kinetic-parameter establishment it is imperative to characterize the fuel used. The fuel chosen was a
well-pyrolyzed wood-based char (biomass). The pyrolyzation has
been carried out in the uidized bed reactor under pure nitrogen
atmosphere with a residence time of 30 min at 900 C. Fuel particles were ground and sieved to a typical diameter of

Table 1
Fuel composition of wood char WC1173 (biomass pyrolyzed at 1173 K for 30 min).
Water

Ash (dry)

Volatiles (dry)

Proximate analysis (wt-%)


0.10

3.88

3.70

Ultimate analysis (dry, wt-%)


92.3
0.42

To verify operation under well-stirred conditions the uidization regimes of the bed must be identied, as well as homogeneity
of temperature and composition inside the uidized bed.
3.1.1. Qualication of the uidization regimes
To examine the uidization regimes, observations were conducted in a transparent (cold) uidized bed with identical dimensions. From analysis of high-speed video (HSV) and pressure
measurements the regimes of uidization were identied. Bed
height was deduced from time-averaged images, were it was
dened arbitrarily as the location of a 50% decrease in brightness.
Fig. 4a) shows the results. With the onset of ow there is a slight
increase in bed height, though no bubbling can be observed just
gas ow between the particles. The rst bubbles appear 60 Nl/h,
accompanied by a clear increase in bed height, after which bed
height increases almost linearly with ow (Fig. 4a) & leftmost
inset).
An additional interesting observation made is the signicant
inuence of system over pressure (due to back-pressure in the
sampling line). This is represented by the curves in Fig. 4a. As the
gure shows, even a low backpressure of 20 mbar extends the
pre-bubbling and bubbling range. Most importantly it shifts the
transition to full-uidization to somewhat higher ow rates. This
aspect must be considered when operating the bed under higher
pressures.

50

(a)

40

Bed Height [mm]

0.2

30
20
10

No back pressure
20 mbar back pressure

0
5

Pressure Drop [mbar]

Normed Concentration [-]

266

(b)

4
3
2

Total
Distributor
Particle Bed

1
0
0

0.44

0.01

2.95

100

200

300

400

500

600

Flow Rate [Nl/h]


Fig. 4. Fluid-dynamics of the uidized bed (cold transparent system): (a) average
increase in bed height; (b) pressure drop across each element.

267

H.D. Haustein et al. / Fuel 158 (2015) 263269

As the pressure readings show (Fig. 4b), subtraction of the distributor pressure drop (measured without a bed) from the total
(measured with a bed) gives the additional drop due to the bed
particles. This bed ow resistance shows a constant increase past
the point of bubbling (60 Nl/h), with a clear drop-off from this
trend (saturation of the pressure drop) when full uidization is
attained (100 Nl/h). The fully uidized state can be seen in the
rightmost inset in Fig. 4a at an extreme ow-rate of 500 Nl/h). At
such high ow rates a large amount of particles can be carried
away by the ow obstructing the ow in the sampling line and
leading to an intermittent type ow (periodic eruption).
Therefore, optimal operation of a uidized bed of these size particles is in the lower fully uidized range (well stirring) at 200
250 Nl/h, the range used in subsequent experiments. The ranges
obtained from the cold ow observations were converted to the
much hotter bed conditions under the assumption that the density
follows the ideal gas temperature dependence and viscositys
dependence is described by the kinetic theory of gases.

3.1.3. Homogeneous bulk gas-composition


Similar measurements were conducted to establish uniformity
of the gas composition: The FTIR pickup pipe was introduced into
the uidized bed and raised from 30 mm to 80 mm in a 1 mm
interval. At each height char particles (15.75 2.05 mg) were
dropped in and combusted with air (21% O2), while the gas was
extracted from the respective location. This procedure was
repeated 23 times at each height. The real-time CO2 and CO concentrations measured by FTIR were converted to a burnout form
(integration of Eq. (2)), as shown in Fig. 6. As the gure indicates,
no clear trend can be observed with height, i.e. the variation of
concentration with height is not larger than the variations encountered from run to run at a xed height. This shows that the signal is
the same regardless of the height within the bed, suggesting that
the gas composition in the bed is indeed closely homogenous.
Once the well-stirred conditions were experimentally validated,
the establishment of reliable kinetic parameters was pursued.
3.2. Method application kinetic parameters of char gasication

3.1.2. Homogeneous bulk temperature


Once uid-dynamic operating conditions were established, the
FBR could be optimally operated at higher temperatures. In order
to verify that the conditions in the bed are known and controlled,
the temperature of the inow gas (heating up in the annulus and
then owing through the distributor) was measured. This was
done as follows: Initially, the ow was allowed to exit the reactor
through the exhaust pipe at the top of the reactor. At a certain
instant the exhaust was closed and the ow was forced past the
thermocouple, which is located in a pipe 30 mm above the distributor. This procedure was repeated several times and in two different modes: under steady state conditions at 800 C and during heat
up (500 C at a heat up rate of 200 K/h). In the rst case, no measureable change in temperature was observed (the interior of the
system is typically 3 K lower than the preset/oven temperature).
While in the latter, a clear increase of the measured heat up rate
was observed indicating that the air ow was hotter than the inner
system (which typically trailed the oven temperature by 30 K).
From these two observations it is safe to conclude that the current
setup does indeed deliver the inow gas at a temperature very
close to the preset value. To further establish the existence of spatially homogeneous conditions in the hot bed, K-type thermocouple measurements were conducted at several heights above the
distributor (cf, Fig. 5).
Thereby, the uidized bed operational height was typically
70 mm (immersed thermocouple). As Fig. 5 shows, the temperature over the height of the bed did not vary more than the typical
uncertainty of the measurement (0.5%) even over a 24 h period.
Thus, the bed can be claimed to be homogenous in temperature.

To demonstrate the applicability of the method for nding


kinetic parameters a specic fuel was chosen wood char
pyrolyzed at 900 C (see fuel details in Section 2.3). This fuel was
chosen due to its high carbon/low ash & volatile content, which
is suitable for a pure gasication reaction (Boudouard reaction,
Cs + CO2 ? 2CO). Gasication experiments were conducted in the
temperature range of To = 8001100 C and at concentrations of
C CO2 = 1976% (rest: N2).
Trends of CO and CO2 concentrations for three consecutive gasication experiments at 900 C are presented in Fig. 7. The three
peaks in this diagram (CO) are typical for the gasication experiments in the uidized bed reactor. Furthermore the gure reveals
that CO2 concentrations stay nearly constant throughout the complete reaction, which is favorable for the extraction of kinetic data.
Assuming the reaction is well described by an n-th order rate
equation and temperature dependency can be expressed with help
of the Arrhenius equation, the following term can be derived for
the reaction rate:
E

r k0 eRT C nCO2

where R represents the universal gas constant and T the temperature. In this equation the pre-exponential factor k0, activation
energy E and reaction order n are the kinetic constants that have
to be determined.
First, the activation energy and pre-exponential factor are found
by variation of the temperature, at constant CO2 concentration.
Next, the order of reaction is deduced by variation of the
reactant-gas concentration at constant temperature, as shown in
Figs. 8 and 9, accordingly. The applied reaction rate for the determination of these parameters is the characteristic rate found by

Carbon Conversion X [-]

1100

Temperature [K]

1050
1000
950
900
T(t )
1

850

T(t +24 h)
1

800

50

100

150

200

250

300

350

400

1.0
0.8

30 mm
40 mm
50 mm
70 mm
80 mm

0.6
0.4
0.2
0

30 mm
50 mm
60 mm
70 mm

30 mm
50 mm
60 mm
80 mm

R = 0.039 s -1
R 0.006 s -1
0

50

100

150

Time t [s]

Height above Distributor [mm]


Fig. 5. Temperature vs. height above distributor, measured in a continuous 24 h
interval.

Fig. 6. Carbon conversion vs. time in air (21% O2) at 800 C. Different symbols
indicate gas-sampling at different heights above the distributor, and comparison to
average URM prediction.

268

H.D. Haustein et al. / Fuel 158 (2015) 263269


4
3.5 x 10

2
50
1.5
1

ln(r) [-]

75

2.5

CO2 [Vol. %]

CO [ppm]

3.6

100
CO2
CO

4.5

25

0.5
0
0

10

15

20

25

30

35

0
40

5
2

Time t [min]

iterative tting (to the 1080% of the carbon conversion curve)


using the Eq. (1), and the convolution procedure described by
Eqs. (3)(5). For the kinetic study an averaged value of ve repetitive measurements is used.
The activation energy was found by varying the temperature at
a xed concentration of 22 2% CO2 at four different temperatures
from 900 to 1100 C (cf, Fig. 8). As the gure shows can be approximated by an exponential law. Using Eq. (6) the activation energy
and pre-exponential factor are calculated to be 205.4 kJ/mol and
6.88106 s1 bar0.6 respectively.
In order to establish the order of reaction 4 measurements at
1000 C and different reactant gas concentrations were performed.
As Fig. 9 shows, the distribution of repeated results was reasonable
(see error bars). The order of reaction was found as n = 0.60, by tting a straight line to four different CO2 concentrations. This value
is similar to the value found by Kajitani et al. [23] for the lowest
volatile sub-bituminous coal char they examined (n = 0.56), though
the activation energy there is somewhat higher (257 kJ/mol).
Better agreement is found with the CO2 gasication literature of
wood-based biomass: Risnes et al. [24] obtained very similar values
of n = 0.59, 205.6 kJ/mol and 5.81106 s1 bar0.59 for activation
energy and pre-exponential factor, while Barrio gave values of
215 kJ/mol and 3.10106 s1 bar0.38 [25].
It is important to conduct the kinetic study under intrinsic reaction conditions (Regime I). To verify that these conditions existed
in our experiments, they were conducted only up to a temperature
where a deviation from the low-temperature trend was observed
(above 1100 C). Furthermore, the Thiele modulus and
Effectiveness Factor expressing the extent of pore-diffusion limitation on an nth-order reaction were evaluated, according to the
form given from Levenspiel [26] (similar to the form used in [6]):

-3
-3.5

0.5

Fig. 9. Reaction rate vs. CO2 concentration, WC1173 at 1000 C.

dp
/
6

s
E
ko eRT C n1
CO2
DCO2 e2



1
1
1

/ tan h3/ 3/

Here / is the Thiele modulus, dp is the particle diameter, k0eE/RT is


the rate constant, C CO2 is the CO2 concentration, n is the order of reaction, DCO2 is the diffusivity constant of CO2 in N2 and e is the particle
void fraction. In Eq. (7), the effective pore diffusivity was introduced
as DCO2  e2 following [6]. An estimate of the Effectiveness Factor at
the maximal reaction rate found (r < 0.1 [1/s]) around 1100 C (with
C CO2 = 0.76 and DCO2 = 1.1  104 [m2/s] from [27]), together with
order of reaction found above and characteristic particle voidage
e = 0.74 (found from Hg-porosimetry), gives a value of / = 1  103,
resulting in g = 0.999, indicating that pore-diffusivity is not a limiting
factor to the reaction, and intrinsic kinetic conditions are maintained.
3.3. Comparison to other methods
(1) With the characteristics of the method claried and validated a more detailed comparison to other existing methods,
beyond the obvious differences presented in the introduction, is possible. As demonstrated in this paper the FBR
delivers spatially homogeneous conditions. In comparison
to that the TGA does not represent the individual particle
well or have homogenous conditions since particles are in
direct contact with each other and the scale substrate.
(2) Currently, FBR results of gasication (a slower reaction) are
well-established, though combustion (a faster reaction)
results are still not sufciently reliable. Here the DTR has
some advantage over the FBR (and the TGA) due to its ability to resolve much shorter time scales.
(3) The present method (and the DTR) with high heat up rates
allows the study of an apparent reaction rate of high volatile
content fuels whereas the TGA mode of operation (slow heat
up) prevents the fuel from carrying almost any volatiles into
the reaction.
4. Discussion & conclusion

-4

ln(r) [-]

ln(CCO ) [-]

Fig. 7. Concentration proles of CO and CO2 for a typical gasication experiment at


900 C and 73% CO2.

-4.5
-5
-5.5
-6
-6.5
7.2

1.5

7.4

7.6

7.8

1/T [1/K]

8.2

8.4

8.6
-4
10

Fig. 8. Dependence of reaction rate on temperature, WC1173 fuel at 20% CO2.

A method was presented for the experimental study of


gas-particle reactions. Spatially homogeneous conditions were
demonstrated by investigation of temperatures and gas concentrations at different heights in the uidized bed. Measurements
showed good uniformity, indicating homogenous conditions. This
can be understood to be partially due to intensive mixing and partially due to the small scale: Even strong reactions will not inuence the bulk temperature or gas composition, as fuel samples
are small compared to the bed. This homogeneity is further supported by the consistency and repeatability of results over several
weeks (including system restarts).

H.D. Haustein et al. / Fuel 158 (2015) 263269

With homogeneity and reliability established, kinetic parameters were obtained for char gasication with high certainty and
in agreement with the literature. Char burnout rates were obtained
in batch-experiments by repeated FTIR spectroscopy of the product
gas. Data analysis procedures and establishment of kinetic parameters were exemplied by gasication of a pyrolyzed wood char.
The presented method having well-controlled conditions and
repeatable results, delivers competitive performance that allows
the development of a reaction kinetics database for support and
validation of simulation. Based on the understanding and results
gained from this study, the FBR can be evaluated as a
well-stirred reactor and by comparison to the mentioned methods.
The presented method also entails shortcomings which are discussed in the following.
The fuel delivery has uncertainty related to it as not all the carbon delivered is recovered in the gas analyzer. Additionally long
timescale operation (>3 h) has proven to be problematic: Lower
fractions of carbon are recovered, possibly because of weak gas
leakage, system noise or continuous particle elutriation from the
bed. Dead volume, piping length in the sampling system and dispersion of the signal limit reliable measurement of reactions
shorter than 15 s. Another limitation is the inherent coupling of
bed uid-dynamics to reactant-gas delivery. At higher ow rates
excessive entrainment occurs, yet sufcient reactant-gas may not
be available. This limit depends primarily on particle size, e.g. in
this study rates were limited to a reaction rate of about 0.07 s1.
Acknowledgements
The authors would like to thank the Helmholtz association for
funding Dr. Herman Haustein and Thobias Kreitzberg within the
frame-work of the Helmholtz Virtual Institute for Gasication
Technology (HVIGasTech). Therein, funding has been provided via
the Initiative and Networking Fund of Helmholtz Association.
Also, the authors would like to express their gratitude to DFG for
funding Benjamin Gvert and Dr. Anna Massmeyer via the
SFB/Transregio 129 Oxyame.
References
[1] Lacroix R, Fournet R, Ziegler-Devin I, Marquaire P. Kinetic modeling of surface
reactions involved in CVI of pyrocarbon obtained by propane pyrolysis. Carbon
2010;48:13244.
[2] Sarathy SM, Westbrook CK, Mehl M, Pitz WJ, Togbe C, Dagaut P, et al.
Comprehensive chemical kinetic modeling of the oxidation of 2-methylalkanes
from C7 to C20. Combust Flame 2011;158:233857.
[3] Geier M, Shaddix CR. Kinetic rate parameters for an extended single-lm char
consumption model proposed for CFD simulations of oxycombustion of
pulverized coal. In: 28th international Pittsburgh coal conference,
Pittsburgh; 2011.

269

[4] Luo C, Watanabe T, Nakamura M, Uemiya S, Kojima T. Development of FBR


measurement of char reactivity to carbon dioxide at elevated temperatures.
Fuel 2001;80:23343.
[5] Barker AF, Hart D, Hayhurst AN. Kinetics of production of nitric oxide during
the pyrolysis of small particles of coal in a hot (electrically heated) bed of sand
uidised by pure nitrogen. J Energy Inst 2008;81:12530.
[6] Fennell PS, Kadchha S, Lee H, Dennis JS, Hayhurst AN. The measurement of the
rate of burning of different coal chars in an electrically heated uidised bed of
sand. Chem Eng Sci 2007;62:60818.
[7] Duan L, Zhao C, Zhou W, Qu C, Chen X. Investigation on coal pyrolysis in CO2
atmosphere. Energy Fuels 2009;23:382630.
[8] Shaddix CR, Hecht ES, Jimenez S, Lee SM. Evaluation of rank effects and gas
temperature on coal char burning rates during oxy-fuel combustion. In: 34th
international conference on coal utilisation and fuel systems, Clearwater;
2009.
[9] Murphy JJ, Shaddix CR. Combustion kinetics of coal chars in oxygen-enriched
environments. Combust Flame 2006;144:71029.
[10] Molina A, Shaddix CR. Ignition and devolatilization of pulverized bituminous
coal particles during oxygen/carbon dioxide coal combustion. Proc Combust
Inst 2007;31:190512.
[11] Kunii D, Levenspiel O. Fluidization engineering. Butterworth Heinemann;
1991.
[12] Smoot LD, Pratt DT. Pulverized-coal combustion and gasication. Springer;
1979.
[13] Smoot LD. Fundamentals of coal combustion: for clean and efcient
use. Elsevier; 1993.
[14] Smith IW, Tyler RJ. The reactivity of a porous brown coal char to oxygen
between 630 and 1812 K. Combust Sci Technol 1974;9:8794.
[15] Christ D, Habermehl M, Frster M, Hatzfeld O, Kneer R. The effect of the
Boudouard-reaction on reaction rates of coal chars in CO2/O2- and N2/O2atmospheres at oxygen contents from 0% to 30%. In: 28th international
pittsburgh coal conference, Pittsburgh; 2011.
[16] Haustein HD, Christ D, Habermehl M, Gvert B, Hatzfeld O, Kneer R. Operation
of a small uidized bed reactor for investigation of particle reaction: pyrolysis,
char combustion and the Boudouard reaction. In: 6th European combustion
meeting ECM 13, Lund;2013.
[17] Palconok GI. Heat and mass transfer to a single particle in uidized
bed. Gteborg Chalmers University of Technology; 1998.
[18] Yu J, Zeng X, Zhang J, Zhong M, Zhang G, Wang Y, et al. Isothermal differential
characteristics of gassolid reaction in micro-uidized bed reactor. Fuel
2013;103:2936.
[19] Everson RC, Neomagus HWJP, Kasaini H, Njapha D. Reaction kinetics of
pulverized coal-chars derived from inertinite-rich coal discards:
characterisation and combustion. Fuel 2006;85:106775.
[20] Szekely J, Evans JW. A structural model for gas-solid reactions with a moving
boundary. Chem Eng Sci 1970;25:1091107.
[21] Bhatia SK, Perlmutter DD. A random pore model for uid-solid reactions: I.
Isothermal, kinetic control. AIChE J 1980;26:37986.
[22] Abad A, Cardona SC, Torregrosa JI, Lopez F, Navarro-Laboulais J. Flow analysis
deconvolution for kinetic information reconstruction. J Math Chem
2005;38:27192.
[23] Kajitani S, Suzuki N, Ashizawa M, Hara S. CO2 gasication rate analysis of coal
char in entrained ow coal gasier. Fuel 2006;85:1639.
[24] Risnes H, Holst Srensen L, Hustad JE. CO2 reactivity of chars from wheat,
spruce and coal. Progress Thermochem Biomass Convers 2001:6172.
[25] Barrio M, Hustad JE. CO2 gasication of birch char and the effect of CO
inhibition on the calculation of chemical kinetics. Progress Thermochem
Biomass Convers 2001;1:4760.
[26] Levenspiel O. Chemical reaction engineering. Wiley; 1999.
[27] Marrero TR, Mason EA. Gaseous diffusion coefcients. J Phys Chem Ref Data
1972;1(1):62.

También podría gustarte