Sunteți pe pagina 1din 11

542 Ind. Eng. Chem. Res.

1997, 36, 542-552

An Experimental Investigation of Heat-Transfer Limitations in the


Flash Pyrolysis of Cellulose
Mario Lanzetta, Colomba Di Blasi,* and Federico Buonanno
Dipartimento di Ingegneria Chimica, Universitá degli Studi Napoli Federico II,
P.le V. Tecchio, 80125 Napoli, Italy

A new experimental system is presented to investigate the fast pyrolysis of solid fuels, in the
absence of heat- and mass-transfer limitations. It consists of an electrically heated furnace,
where a thin layer of powdered solid is exposed, on both sides, to radiative heating. A PID
temperature controller is programmed for two different working conditions: the usual constant
furnace temperature (A) and a constant sample temperature (B). Cellulose pyrolysis is
investigated in the temperature range 523-699 K. It is shown that significant heat-transfer
limitations cannot be avoided with the modality A, unless very slow heating rates, as in the
classical TGA systems, are applied. In the modality B (global heating rates 19-56 K/s), the
independence of the char yields from the sample thickness, for values of this below a critical
value, indicates negligible spatial temperature gradients and activity of intraparticle secondary
reactions of primary vapors. External heat-transfer limitations, due mainly to endothermic
reaction energetics, are also avoided through proper variation in the intensity of the external
radiative heat flux. Consequently, conversion occurs under exactly determined temperature
conditions. A cold helium flow carries away from the reaction environment volatile products so
that the activity of extra-bed secondary reactions is hindered as well. Cellulose weight loss and
temperature curves are applied to evaluate the global degradation kinetics and to study the
influences of heat- and mass-transfer limitations.

Introduction and mass transfer and rather narrow ranges of (low)


temperature conditions (Antal (1982, 1985), Varhegyi
Pyrolysis of lignocellulosic materials is important in et al. (1994), Antal and Varhegyi (1995)). In most cases,
the fields of energy recovery, through thermochemical the sample (reaction) temperature is not known, given
conversion, and fire safety science (Di Blasi (1993)). In the significant differences with the controlled system
general, for convective/radiative heating (Di Blasi (1994, and the lags in the measuring devices (Narayan and
1996a,b)), pyrolysis can occur in a chemically controlled Antal (1996)). Intra- and extraparticle mass-transfer
regime (negligible intraparticle spatial gradients of limitations also exert a strong influence on the extent
temperature and no temperature differences between and selectivity of secondary reactions of vapors evolved
particle and reacting environment) or in a heat-transfer- from primary solid degradation. Indeed, primary py-
controlled regime. In the heat-transfer-controlled re- rolysis volatiles may undergo cracking and repolymer-
gime, two further cases (Di Blasi (1996c)) can be ization in the hot charred particle and/or the reacting
observed: (1) internal heat-transfer control (thermally environment.
thick regime) with very large spatial gradients along Most of the studies on the thermal decomposition of
the particle, and (2) external heat-transfer control organic polymers have been carried out by means of
(thermally thin regime) with no spatial gradients along thermogravimetric analysis (TGA). Generally, associ-
the particle, though the temperature may continue to ated with this methodology, there are several shortcom-
change with time, depending on the heat-transfer rate ings (Agrawal and McCluskey (1983)): (1) uncertainties
from reactor to particle. Heat-transfer conditions de- in temperature measurement, (2) slow heating rates
termine the selectivity of primary solid degradation (usually below 2 K/s) and (3) difficulties in product
reactions and thus the characteristics of pyrolysis. It recovery, so that the kinetic analyses are based only on
is known (Scott et al. (1988)) that slow solid heating weight loss curves. In particular, to avoid increases in
rates lead to comparable amounts of char, gas, and the mechanical inertia of the sample pan and to obtain
liquid products (slow-conventional pyrolysis), whereas accurate and sensitive weight measurements, the ther-
solid char formation is almost completely hindered mocouple is never in contact with the degrading sample.
under very fast heating rates (flash pyrolysis) and Consequently, the reaction temperature is not exactly
reduced volatile residence times (Antal et al. (1990)). It known.
should also be mentioned that a special case of flash
Tube furnaces, which allow faster heating rates (up
pyrolysis is achieved in the ablative regime (Lede et al.
to 5-10 K/s), have also been proposed (Bradbury et al.
(1985), Di Blasi (1996d)), where heat transfer from
(1979), Thurner and Mann (1981), Agrawal and Mc-
reactor to particle occurs by direct contact.
Cluskey (1983)). In this case, products and weight loss
The knowledge of the chemical kinetics of pyrolytic are recorded at selected times, through experiment
solid degradation is an important step in the optimal repetition, to draw the curves of interest. However, this
design and operation of chemical reactors. However, procedure is much more tedious than that of TGA
kinetic investigations, mainly for the fast heating rates analyses, and again only the furnace temperature is
of flash pyrolysis, suffer from large limitations in heat controlled. Therefore, in some cases, the reaction tem-
perature is significantly lower than that of the heating
* Corresponding author. Telephone: 39-81-7682232. Fax: system, usually taken as the reaction temperature (see,
39-81-2391800. for instance, Bradbury et al. (1979)).
S0888-5885(96)00551-9 CCC: $14.00 © 1997 American Chemical Society
Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997 543
As the rates of solid heating and degradation are
made faster, the characteristic times of measuring
devices become comparatively too slow, so that accurate
continuous monitoring of the variables of interest turns
in a rather hard task. These problems are well pointed
out in the fast TGA system developed by Antal and co-
workers (Tabatabaie-Raissi et al., 1989), where high
radiative heat fluxes are applied to pyrolyze cellulose
under conditions similar to those of solar environments.
Usually, only final product yields, as functions of the
final reaction temperature, are obtained from experi-
mental systems characterized by very fast nominal
heating rates (up to 10 000 K/s). Also, it should be noted
that, given the very high heat-transfer rates, though the
temperature of the (inert) heating system is well
controlled, significant differences between this and that
of the reacting samples and large intraparticle gradients
are expected. These systems include screen heaters (for
instance, Lewellen et al. (1976), Hajaligol et al. (1982)),
shock tubes (Ozturk and Merklin (1995)), and Curie-
point pyrolyzers (Wanzl (1994)). In the first case, a
stainless steel mesh screen, suspended between massive
brass electrodes, is rapidly heated to the desired tem-
perature. In the shock tube, the propagation of a planar
shock, through a low-pressure gas, results in an almost
step increase in temperature and pressure. The Curie-
point technique is based on the electric current produced
by a conductor introduced in a high-frequency magnetic
field.
Finally, bench-scale, fluid-bed reactors are also used
(for instance, Lipska and Parker (1966), Barooah and
Long (1976)) to measure the total condensable and gas
yields and reaction time on the dependence of the
pyrolysis temperature and to evaluate reaction kinetics.
Shortcomings are related to the difficult determination
of the yields of char (mixed with the granular solid heat
carrier) and, again, to significant heat and mass transfer
limitations.
In this study, a new experimental system is presented
to investigate the pyrolysis of lignocellulosic materials
under heating rates significantly faster than those used Figure 1. Schematic of the experimental system: (A) general,
(B) cross view of the heating chamber, (C) sample holder.
in TGA systems, while continuous, accurate monitoring
of sample temperature and weight is still carried out.
The most innovative aspect of the experimental ap- and gas. The system is a combination of the pyrolysis
paratus consists of the control of the sample tempera- tube used in previous experiments (Madorsky et al.
ture, after conditions for negligible spatial gradients of (1956), Bradbury et al. (1979), Thurner and Mann
temperature are established (thermally thin regime). (1981), Agrawal and McCluskey (1983), Cullis et al.
The externally applied heat flux is used as the adjust- (1983)) and the fast TGA thermoanalyzer developed by
able variable. In this way, heat-transfer limitations are Antal and his group (Tabatabaie-Raissi et al., 1989).
eliminated. Furthermore, intra- and extraparticle sec- Parts A-C of Figure 1 schematically represent the main
ondary reaction activity is minimized by reducing the characteristics of the experimental apparatus. It con-
volatile residence times (thin sample size and inert flow sists of (1) a radiant heating chamber, (2) a quartz reac-
gas). Cellulose is chosen as fuel because, notwithstand- tor, (3) a PID temperature controller, (4) an inert gas
ing the numerous investigations already available (see feeding system, (5) an acquisition data set (PC and re-
reviews by Antal (1982, 1985), Antal and Varhegyi lated accessories), (6) a precision balance, and (7) a col-
(1995), Milosavljevic and Suuberg (1995)), a still large lection system for condensable and noncondensable
uncertainty is shown on conditions and kinetics of flash products of solid pyrolysis.
pyrolysis. The heating system is a radiant chamber, manufac-
tured by Research Inc. The heating elements are
Experimental System tubular quartz infrared lamps with a tungsten wire
filament that emits radiant energy in proportion to the
An experimental system has been designed and applied voltage. Four elliptical, polished aluminum,
constructed to investigate solid pyrolysis under reduced water-cooled reflectors (Figure 1B) focus the high-
limitations of heat and mass transport, so that informa- density infrared energy, emitted by lamps, onto a
tion on the chemical kinetics (semiglobal mechanisms cylindrically shaped target area, whose diameter is 6.5
and data) can be obtained. The system allows for the × 10-2 m. The length of the heated zone is 6.02 × 10-2
continuous monitoring of sample and system tempera- m, delimited on both sides by two inert zones of about
ture and weight and for the collection of products, the same length. The radiant flux is uniform over 80%
lumped into three main classes: solid char, liquid tar, of the chamber length. The maximum temperature,
544 Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997

2500 K, is achieved within 3 s, resulting in rather fast of the sample temperature, T, for the computational
reactor heating rates. To avoid interaction between the domain consisting of the half-thickness layer, τ, can be
volatile pyrolysis products and the lamps, a quartz tube described as:
(i.d. 6 × 10-2 m and length 20 × 10-2 m), transparent
to infrared radiation, is located inside the furnace and σS(TL4 - T4)
used as a reaction chamber. A powdered solid sample ∂T
(Fscs + cgFg) ) Qr + (1)
is exposed to thermal radiation by means of a 325 ∂t V
stainless steel mesh screen (4.2 × 10-2 m large and 5.5
× 10-2 m long; Figure 1C), whose sides are wrapped on where Qr and Fc are dependent on temperature and
two steel rods. Accurate temperature measurements, chemical composition. Also, the emissivity may not be
under different heating conditions, have been made to constant, with further effects on the external heat-
define the uniform temperature zone (this corresponds transfer rate. Thus, eq 1 should be formulated and
to an extention of about 2.6 × 10-2 m, in the direction solved in conjunction with chemical species conservation
of the reactor axis, and 2 × 10-2 m, along the other). equations (Di Blasi (1996c)). In the modality A, Qe )
The sample is located at the center of the radiant σSTL4 is constant and the temperature evolution is
chamber and is radiatively heated on both the upper thus the result (eq 1) of the effects of both the external
and bottom sides, this in contact with the sample holder (constant) heating and the variations in the enthalpy
(Figure 1C). Aluminum supports directly connect the and properties caused by chemical reactions.
sample holder to a precision balance (Mettler-Toledo From the mathematical point of view, eq 1 and the
AT400) with an accuracy of order 0.1 mg and a time control equations are applied to determine how Qe
constant of 0.2-0.4 s. should be varied, in order to achieve a chosen sample
The temperature of the reacting solid is measured by temperature evolution in the modality B. According to
a thin (0.1 mm bead) chromel-alumel thermocouple in the classical theory of chemical process control (Seborg
direct physical contact with the powdered sample layer et al. (1989)), eq 1 is linearized, expressed in terms of
(Figure 1C). As pyrolysis takes place, the thermocouple deviation variables, transformed in the Laplace domain
bead moves according to shrinking rate, and, depending and coupled to a PID controller. In the experiments,
on the final solid residual yield, at the completion of the the first step deals with a trial and error procedure
process, it may touch the surface of the sample holder. applied to the empty sample holder for a first estimate
The measured temperature coincides with the sample of the PID controller parameters. Then, prior to each
temperature only if the conversion process occurs in a run, these are adjusted for the presence of the sample
thermally thin regime, that is, if a uniform profile is to be pyrolyzed, again with some trial and error tests,
established along the sample thickness. Therefore, so that the maximum heating rate can be achieved, with
conditions for conversion in a thermally thin regime no deviation from the chosen reaction temperature.
should be first determined. The procedure applied in A continuous flow of helium is applied to establish
this study is based on the well-known dependence of an inert environment and to reduce the residence time
the residual solid yield on the reaction temperature of vapors inside the reaction chamber. It is distributed
at the entrance of the quartz reactor, at a distance of
(primary char) and the intraparticle residence times of
about 6 × 10-2 m from the sample holder. This is an
primary volatile products (secondary char). Indeed, as
important point. It is generally retained that, under
the temperatures of the primary degradation are in-
conditions of radiative heating (Tabatabaie-Raissi et al.
creased (Shafizadeh et al. (1979), Bradbury et al. (1979))
(1989)), secondary reactions are promptly quenched
or the residence times of primary vapors inside the
because, as soon as volatile products are released from
reacting particle are made shorter (Antal et al. (1990)),
the hot reacting solid, they encounter a zone where the
the char yields decrease. Thus, for chosen heating
temperature is rather low (apart from a narrow gas
conditions, the sample thickness is decreased until the
layer, adjacent to the solid surface, which is heated by
solid residual yield becomes constant.
conduction). However, it is important to reduce the
The radiant heater is made to work according to two residence time of volatile pyrolysis products in the
modalities, here indicated as A and B, which correspond radiant chamber, because it is also well known (Di Blasi
to the emission of a constant flux and to the attainment et al. (1991)) that they can absorb some of the incident
of a certain temperature with an assigned rate, respec- radiation and thus attain temperatures sufficiently high
tively. The modality A, which is the easier one, also for secondary reactions to become active. The continu-
corresponds to a constant temperature of the emitters, ous flow gives rise to very low concentrations of volatiles
after short transients whose duration is successively in the reaction chamber (for the worst case, that is, the
reduced as the intensity of the radiation is increased. highest temperatures, it is estimated that the ratio
The heating modality B can be implemented to impose between volatiles and inert is about 1/1000 mol), thus
a certain time evolution of a temperature, which can keeping gas-phase absorption of radiation to a mini-
be that of the emitters or the specimen, by varying the mum. Furthermore, helium is preferred to argon, given
intensity of the radiant flux. In this study, the con- the higher diffusion coefficient (a factor of about 3) of
trolled variable is the temperature of the sample the high molecular weight volatiles (Varhegyi et al.
surface, after sample sizes for conversion in a thermally (1994)). The flow of inert gas, at ambient temperature,
thin regime are determined. The manipulated variable is also used to cool the residual char, left after complete
is the electric voltage applied to the furnace or, in other solid devolatilization.
words, the intensity of the radiative heat flux Qe. Usually, two runs are required to measure the weight
In the absence of spatial gradients along the sample loss. In the first, the weight loss curve of the degrading
layer, the dynamics of cellulose degradation, for the sample is obtained. The second run is carried out to
heating modality A, are well described by means of the record the so-called “blank” curve, which takes into
mathematical model for a thermally thin regime re- account the effects of spurious weight changes due to
ported by Di Blasi (1996c) (assumptions are listed in temperature effects in the gas phase surrounding the
the same reference). In particular, the time evolution sample while degrading (Tabatabaie-Raissi et al. (1989)).
Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997 545
This curve, obtained with the sample holder left in the
position of the first test and with the final char yield, is
then used to correct the weight loss curve. These cor-
rections are, at the higher temperatures, on the order
of 50% with respect to the measured sample weight. In
order to verify the correctness of the procedure, after
each run, the final char yield is also weighed and com-
pared with that previously determined. Given the rela-
tively large sample mass (20-8 mg) and the very thin
thermocouple, the close contact of this with the sample
does not affect the mechanical inertia of the system and
simultaneous accurate measurements of temperature
and weight are obtained. An acquisition data system,
based on a PC and the software LabView (National Figure 2. Solid mass fraction as function of time (τ ) 50 µm) for
Instruments), allows for continuous monitoring (0.4-1 different system temperatures (heating modality A).
s) and storage of temperature and weight loss values.
In this study no extensive investigation on the
dependence of product yields on reaction conditions has
been carried out, as the analysis is focused mainly on
the effects of heat-transfer limitations. However, a
product collection system is being applied, consisting of
a condenser (a 10 × 10-2 m long tube filled in with glass
wool) and a plastic bag for gaseous products.
Results
Thermal degradation of thin cellulosic layers is car-
ried out according to the heating modalities A and B,
by varying the intensity of the radiative heat flux and
the final temperature of the sample, respectively. The
influence of intrabed temperature gradients on the
conversion process (apparent kinetics) is investigated,
by varying the sample thickness. Conditions leading
to a true isothermal process, controlled by (intrinsic)
chemical kinetics, are determined. In all cases, com-
mercial cellulose with lengths of the fibers in the range
20-150 µm (Sigma Aldrich), predried for about 10 h at
a temperature of 388 K, is degraded. Cellulose is chosen
as fuel, given its importance in the field of biomass
thermochemical conversion and the still uncertain
behavior under flash pyrolysis conditions. The helium
flow, at ambient temperature, is equal to 8 × 10-4 m3/
min, which, given a cross section of the quartz reactor
of 28 × 10-4 m2, gives rise to velocities of about 30 ×
10-2 m/min. No variations in the degradation process
are seen with larger mass flow rates (factors of 2-3). Figure 3. Time-temperature curves for cellulose (solid lines) and
Cellulose Degradation under Constant Radia- char (dashed lines) (A) for two different system temperatures (Tf
tive Heat Flux (Heating Modality A). Tests have ) 577 and 633 K) and τ ) 50 µm (heating modality A) and (B) for
been conducted, according to the heating modality A, two different sample half-thicknesses (τ ) 100 and 44 µm) and Tf
for about 10 mg of cellulose powder, uniformly distrib- ) 633 K (heating modality A).
uted along the uniform temperature zone of the sample
holder, to give τ ) 50 µm. As will be shown later, this Examples of the weight loss curves (the solid residual,
sample thickness corresponds, for moderate reaction W, including char and cellulose, expressed as a fraction
temperatures, to conversion in a thermally thin regime. of the initial sample mass, M0) are reported, for system
The steady temperature value, Tf, here indicated as the temperatures in the range 543-622 K, in Figure 2. As
“system temperature”, is used to characterize the dif- expected, preheating times, before significant weight
ferent tests and has been varied in the range 543-699 loss, become successively shorter and the amounts of
K, through different values of Qe. For the sample solid, left as charred residual, decrease as the intensity
thickness given above, the cellulose heating rate (re- of the radiative heat flux (or the system temperature)
ferred to the sample surface and before significant is increased.
weight loss) varies from about 12 to 54 K/s as the system Associated with weight loss curves, the temperature
temperature is varied. As will appear from the experi- of cellulose (recorded in the course of degradation) and
ments, external heat-transfer limitations can hardly be char mass fraction (recorded together with the blank
avoided with this heating modality, so that it is not curve) are available. These can be used to get informa-
applicable to study the intrinsic kinetics of degradation. tion on the thermal history of the sample. Examples of
However, some results have been obtained which can the (char and cellulose) temperature curves are plotted
be of interest from the practical point of view, that is, in Figure 3A,B, for two different system temperatures
in relation to pyrolysis units, where biomass particles (Tf ) 577 and 633 K) and τ ) 50 µm (A) and for two
are exposed to a certain reactor temperature. Also, different sample sizes (τ ) 100 and 40 µm) and Tf )
comparison with previous analyses can be carried out. 633 K (B). In all cases, a region exists where the sample
546 Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997

temperatures are significantly lower than those of the


char. The difference becomes successively larger as the
system temperature or the sample thickness is in-
creased. The lower cellulose temperature, measured
during the process transients, is due to reaction ener-
getics and, in principle, larger medium thermal capacity.
Concerning this last point, it should be noted that the
reduction in the effective sample capacity is rather
large, given that, even at the lowest temperature, a
reduction in the sample weight of about 75% is mea-
sured. However, given that, in all cases, for tempera-
tures below 500 K no significant difference is seen
between cellulose and char, it can be assumed that the
role played by thermal capacity is less important than
Figure 4. Maximum temperature difference between cellulose
that of the endothermicity of pyrolysis reactions (tar and char, ∆Tm, and cellulose temperature for ∆T ) 50%∆Tm, T1,
formation). and for ∆T ) ∆Tm, T2, as functions of the system temperature (τ
As expected, the temperature difference becomes ) 50 µm, heating modality A).
successively smaller as time increases, because the
reaction process approaches completion and the two values of this below 600 K. However, for larger values,
media tend to be the same. Also, a plateau in the the rate of increase is successively reduced and a
cellulose time-temperature curve is clearly seen, as- tendency toward a constant value of about 600 K
sociated with the onset of high rates of solid degrada- appears. Apart from low Tf values (e580 K), ∆Tm
tion, for Tf g 600 K. It becomes successively more continuously increase with Tf, at a rate that becomes
evident as the system temperature or the sample faster for values of this above 600 K, as a consequence
thickness is increased. In fact, given the exponential of the trend previously observed in T2. It should be
increase of the reaction rate with temperature, the heat noted that the temperature difference may become as
requirements by the endothermic degradation also high as 30 K.
become more stringent, as Tf or the amount of solid to Results presented in Figure 4 indicate that, for
be pyrolyzed (sample thickness) are increased. assigned external thermal conditions (radiation), such
as in chemical reactors, the effective reaction temper-
Under mild heating conditions, (Tf < 580 K), both ature can hardly be brought above a certain critical level
cellulose and char slowly approach the steady final of about 600 K. Increasing the sample heating rate,
conditions and a period (whose duration is, at least, 1/3 through reactor temperature, reduces the duration of
of the total conversion time) is seen where the two the first stage of degradation, which occurs below 600
temperatures become equal. The period where the two K and on average the reaction temperature increases.
temperatures are equal is reached only for conversion However, given the endothermicity of the degradation
levels of about 90% for 580 K < Tf e 600 K. Finally, process and the successively faster reaction rates, the
for Tf > 600 K, cellulose always degrades at tempera- increase in the effective reaction temperature of cel-
tures significantly lower than those of the system, lulose is successively reduced and hardly goes above 600
though they remain almost constant for most of the K, with this heating modality. According to eq 1 and
devolatilization. It should be noted that this behavior extensive computations (Di Blasi (1996c), Narayan and
has been observed also for very thin sample layers, Antal (1996)), an almost constant temperature is at-
indicating that external heat-transfer limitations are tained at the fast heating rates because of a balance
very high and cannot be overcome under constant between the requirements of endothermic pyrolysis and
heating conditions and temperatures above 600 K, the external heat supply. Also, the occurrence of
which give rise to rather fast sample heating rates. This pyrolysis during the time-temperature plateau may
behavior was already observed in previous experimental explain why the thermal conversion of cellulosic materi-
studies (Antal et al. (1980), Bradbury et al. (1979)) and als has been interpreted as a fusion process. In par-
is the thermal lag due to the endothermicity of cellulose ticular, for contact (ablative) pyrolysis of wood (Lede at
degradation (high temperatures), well described in the al. (1985)), a “wood fusion temperature”, equal to 738
theories proposed by Di Blasi (1996c) and Narayan and K, has been introduced. This value is significantly
Antal (1996). Convective outflow of volatiles also affects higher than the limit observed in this study for cellulose,
the sample heating rate. As the devolatilization rate but the different chemical behavior between the two
(and the flow velocity) increases with temperature, it solids and the difference in the external heat-transfer
has been observed that such a contribution becomes rates (much larger in the case of ablative pyrolysis) may
successively more important as more severe conversion justify the difference.
conditions are considered (Di Blasi (1994, 1996c)). The significant differences between cellulose and
The maximum temperature difference (between char system temperature for a large part (low temperatures)
and reacting cellulose), ∆Tm, and two sample temper- or for the whole duration of the degradation process
atures, T1 and T2, corresponding to ∆T ) 50%∆Tm and (high temperatures and/or heating rates) make ques-
∆T ) ∆Tm, respectively, are reported in Figure 4 as tionable the use of the data (weight loss and product
functions of the system temperature. The temperature yields) obtained according to the heating modality A for
T1, which can be taken as representative of the onset the formulation of kinetic models and the estimation of
of significant rates of solid degradation, increases almost kinetic data, since conversion occurs under heat-transfer
linearly with the system temperature (apart from the control. Heat-transfer limitations for modality A can
very low values). On the other hand, T2, which can be be reduced only for slow heating rates, such as those
taken as representative of the temperature of the usually applied in the slow TGA systems which, of
completion of the degradation process, also shows an course, are not directly applicable for flash pyrolysis.
almost linear increase with the system temperature for Also, very thin sample thickness (or small mass) ap-
Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997 547

Figure 5. Solid mass fraction for solid residence times, ts, equal Figure 6. Temperature-time curves of the cellulose sample
to 300 s (Tr ) 573 K) and 10 s (Tr ) 623 K) as a function of the during the initial stage of the process for the heating modality B
sample half-thickness (heating modality B). (τ ) 50 µm).

pears to cause a reduction in the heat-transfer limita- Indeed, the char yield does not depend on the sample
tions, but this approach is again effective only for thickness only when solid degradation occurs at the
relatively slow heating rates. same temperature, independently of the spatial position,
Cellulose Degradation under Controlled Sample that is, no spatial temperature gradients exist. More-
Temperature (Heating Modality B). The analysis over, these conditions also indicate that intrabed sec-
of the thermal degradation data, obtained through ondary charring reactions do not occur to a significant
modality A, has shown that it is possible to avoid heat- extent. For cellulose and for the chosen heating condi-
transfer limitations only for very thin samples and slow tions, the transition from a thermally thick to a ther-
heating rates (low system temperatures). In the heating mally thin regime occurs for sample half-thicknesses
modality B the temperature of the sample is the result slightly increasing with the temperature, that is, τ )
of a control process. If the proper size of the sample is 70 µm (Tr ) 573 K) and τ ) 60 µm (Tr ) 623 K),
chosen to avoid intrabed gradients, a proper selection respectively. Therefore, conversion of samples with
of control parameters allows degradation to be carried half-thicknesses below 60 µm and reaction temperatures
out under exactly known temperature conditions. There- below 623 K can be assumed to occur in a thermally
fore, some preliminary tests have been conducted to thin regime.
determine the conditions giving rise to a conversion in Examples of the time history of the cellulose temper-
a thermally thin regime for modality B. It should be ature, for the initial stage of the process, are shown in
pointed out that, in relation to intrabed heat- and mass- Figure 6. These have been obtained for different set
transfer limitations, the sample thickness, not the points (reaction temperatures) and τ ) 50 µm (about
sample mass, is the key parameter. Consequently, an 10 mg of cellulose powder). After the first stage of linear
important point is to determine, before all, the size of increase (heating rates from 40 to 78 K/s for the
the sample thickness which does not give rise to temperature range considered), a constant (reaction)
significant spatial gradients of temperature (primary temperature, Tr, is more slowly attained (the time
degradation at different temperatures) and volatile needed to attain this value is indicated as heating time,
residence times (activity of secondary reactions). As th) and maintained for the whole duration of the process,
previously illustrated, the char yields are taken as a with no overshoot in practice. The global heating rates,
measure of the importance of these processes. In for the attainment of the reaction temperature, vary
particular, conditions of char yields not dependent on from about 19 to 56 K/s. For the whole range of reaction
the sample thickness correspond to negligible effects on temperatures tested, the endothermicity of degradation
primary degradation of intrabed transport phenomena reactions is well compensated by increased external heat
and secondary reactions. Given that intrabed space supply. Consequently, no plateau is seen in the time-
gradients and secondary reaction rates are expected to temperature curve, even when fast degradation rates
increase as heating conditions are made more severe, are achieved. Therefore, compared to the curves of
two reaction temperatures have been considered, 573 Figure 3, weight loss curves obtained with the heating
and 623 K. modality B show shorter degradation times (Figure 7)
Tests have been conducted, according to the heating for comparable values of Tf (system temperature of the
modality B (examples of the heating curves will be given modality A) and Tr (the reaction temperature of modal-
in the following), by decreasing the thickness of the ity B).
cellulose layer from about 200 µm. The solid residence The weight loss curves of Figure 7 are obtained under
times are 300 (low temperature) and 10 s (high tem- isothermal sample conditions, apart from the very
perature). After this time, the power is turned off, and beginning of the process. Figure 8 reports the ratio
the sample is promptly moved into a cold zone under a between the conversion time, tc, and the heating time,
continuous helium flow. The solid residual, expressed th, and the solid mass fraction at t ) th as functions of
as a fraction of the initial sample mass, is reported as the reaction temperature. For Tr e 607 K, the conver-
a function of the sample half-thickness in Figure 5. As sion time is significantly longer than the sample heating
expected, it initially decreases with the sample size, time (factors in the range 1000-16), and the conversion
because of the increase in the average reaction temper- level achieved in relation to the heating period is also
ature and the decrease in the volatile residence times, small (0.001-22%) (the factors become 5 and 40%,
and then tends to a constant value. respectively, for Tr ) 623 K).
The attainment of a constant value of the char yield Effects of Intrabed Resistances. In order to
corresponds to conversion in a thermally thin regime. understand the influence of intrabed heat- and mass-
548 Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997

The trend shown by Figure 9 is in qualitative agree-


ment with the previous literature on the pyrolysis of
cellulose. In other words, the solid residual increases
as the effective reaction temperature decreases (from
modality B to A) or the cellulose layer is made thicker
for the same heating modality (B). Extrapolation of the
measured data for values of the char yield below 1%
shows that cellulose degradation, under radiative heat-
ing, occurs with no char formation for Tr ) 625 K or Tr
) 630 K as the sample layer thickness is increased
(modality B) or for Tf ) 640 K (modality A).
At first glance, it appears that, for temperatures in
the range 523-560 K, the measured char yields are
Figure 7. Solid mass fraction as a function of time for different comparable with those given by Lipska and Parker
sample temperatures (τ ) 50 µm, heating modality B). (1966), though, in this case, the particle sizes (762 µm)
are much larger than the sample thicknesses used in
the present experiments. This is understandable be-
cause at slow heating rates (low temperatures) a kinetic
regime can be more easily established (Di Blasi (1996c))
or, in other words, the degradation process is so slow
that heat transfer limitations do not impact the results
(Varhegyi et al. (1994)). In general, in all cases, the char
yields measured are lower that those reported in the
literature. For instance, at high temperatures, a com-
parison with the data by Shafizadeh et al. (1979) shows
values lower by factors from 3 to 8. Also, it is worth
observing that the average sample heating rates,
achieved in this study, should be faster than those
typical of fluid-bed flash pyrolysis (Scott et al. (1988)),
given that, in this case, negligible char yields have been
Figure 8. Ratio of the conversion to the heating time, tc/th, and
found only for Tf g 900 K (data from this study are not
solid mass fraction for t ) th, as functions of the reaction
temperature (τ ) 50 µm, heating modality B). included in Figure 9, because they are available only
for Tf g 700 K). On the other hand, negligible char
yields have been reported by Lewellen et al. (1976) for
very fast heating rates (400-10 000 K/s). It can be
understood that these results are due to significant
intra- and extraparticle (bed) heat- and mass-transfer
effects, given the much larger sample sizes used in
previous studies, namely, 595 µm (Scott et al. (1988)),
200 µm (Shafizadeh et al. (1979)), and 420-700 µm
(Barooah and Long (1976)), the different external heat-
transfer mechanisms, or the different solid and volatile
residence times.
Therefore, the analysis of the weight loss curves,
measured in this study, can give information on the
kinetic mechanism and data of cellulose flash pyrolysis
Figure 9. Final char mass fraction as a function of temperature (modality B, τ ) 50 µm). Another consideration can be
for the heating modality B (τ ) 50 and 100 µm) and the heating made in relation to the effective reaction temperature
modality A (τ ) 50 µm). For comparison purposes, literature data of modality A (plausibly in the range of temperatures
are also enclosed. T1-T2, introduced in Figure 4) and the final solid
residual yields (Figure 9). It appears that, with the
transfer resistances on the conversion process, the final heating modality A and for the range of heating condi-
char yields (the solid residual left on the sample holder) tions considered in this study, a very small part of the
are examined on the dependence of the heating condi- reaction, if any, occurs at sample temperatures above
tions, through Figure 9. The temperature range con- 600 K. On the other hand, for system temperatures
sidered is 523-630 K, where the upper limit is dictated above 630 K the char yields are of the order of 1%.
by the accuracy of the weighing instrument, that cannot Therefore, in terms of primary degradation, significant
detect char yields below 1% of the initial cellulose mass. char formation can be avoided if a temperature of 600
For comparison purposes and in relation to the temper- K is achieved within a relatively short time. However,
ature range of interest, not only data obtained for τ ) it should be stressed that the limit temperature of 600
50 µm (modality A and B) and τ ) 100 µm (modality B) K is typical of the experimental conditions achieved in
but also some literature data on isothermal cellulose this study (very low heat- and mass-transfer resis-
degradation are enclosed, obtained through fluid-bed tances).
reactors (Lipska and Parker (1966), Barooah and Long Global Kinetics. For the heating modality B, the
(1976)), a vacuum tube furnace (Shafizadeh et al. process can be divided into two stages: the first is
(1979)), and a vacuum TGA system (Ramiah (1970)) dynamic, that is, the system or sample temperature
(these data are obtained under known external temper- increases to the desired isothermal point, and the second
ature conditions (modality A)). is the true isothermal stage, that is, the temperature is
Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997 549
kept at the constant value attained before. In general, K2M0 -Kt
the isothermal approach of the pyrolysis process can be MT ) MT∞ - e (9)
K
considered valid if (1) there is a negligible weight loss
in the first stage and (2) the duration of the first K1νCM0 -Kt
unsteady stage is a small fraction of the whole time MC ) MC∞ - e (10)
needed for the thermal degradation of the sample. In K
fact, in general, one of the disadvantages of the static The final tar yield, MT∞, can be obtained from eq 9
approach is that there is an excessive weight loss before written for t ) 0 as:
pyrolysis temperature is reached and, of course, the
higher the temperature, the greater the weight loss K2M0
(Ramiah (1970)). MT∞ (11)
It has been shown (Figures 5 and 8, τ ) 50 µm and K
heating modality B) that, for large part of the temper-
Furthermore, the total solid mass (char and cellulose),
ature range investigated, the above assumptions are
W, can be obtained from eqs 8 and 10 as:
verified. Thus, a pure kinetic, isothermal theory is

( )
proposed. Of course, this theory is not applicable for K1νC -Kt
the heating modality A. The kinetics of primary cel- W ) W∞ + M0 1 - e (12)
lulose degradation are described through a semiglobal K
mechanism:
where W∞ is the final solid residual and can be obtained
K1 νCCHAR + νGGAS
from eq 10 written for t ) 0:

CELLULOSE K2 (a1–a2) K1νC


W∞ ) M0 (13)
TAR K

This is a modified version of the well-known Broido- An alternative formulation of eq 12 is thus


Shafizadeh scheme (Broido and Weinstein (1971), Brad-
bury et al. (1979)) for cellulose pyrolysis and takes into
W - W∞
P) ) e-Kt (12′)
account the competitive formations of tar (high temper- M0 - W∞
ature) and linked char and gas (low temperature). The
formation of the “active” cellulose with a reduced degree Figure 10 shows the plots of ln(P) versus time, for the
of polymerization, preceding devolatilization, has been weight loss curves reported in Figure 7 (heating modal-
recently shown not to be controlling (Varhegyi et al. ity A, τ ) 50 µm). It should be noted that the final char
(1994)), and thus it has been dropped from the kinetic yields for Tr ) 614 and 623 K are those extrapolated
model. Moreover, though different orders at different from Figure 9, given their very low values and the
stages of the reaction have been suggested (Milosavljevic limited accuracy of the weighing system. The slope of
and Suuberg (1995)), the assumption of first-order the straight lines is the kinetic constant, K, for the
kinetics is usually retained valid in engineering ap- different temperatures. Clearly, in accordance with
plications (Agrawal (1988)). Varhegyi et al. (1994), the process of weight loss can be
The kinetic equations for the pyrolysis scheme (a1- described by a simple first-order reaction.
a2) are expressed as: The usual Arrhenius plot (Figure 11) and a least-
squares analysis give the activation energy and the
∂MS preexponential factor of the global degradation kinetics.
) -KMS (2) The solid line describes the process in the temperature
∂t
range 523-623 K by one set of kinetic data: E ) 214.5
∂MT kJ/mol, A ) 1.2 × 1017 s-1, with a standard deviation
) K2MS (3) equal to 6%. Again, these data are very close to those
∂t
reported by Varhegyi et al. (1994), who analyzed the
∂MC slow dynamics of the thermal degradation of small
) νCK1MS (4) (0.5-2 mg) cellulose samples. It is interesting to note
∂t that, notwithstanding the highly different experimental
systems and heating conditions, the absence of signifi-
First-order rates are expressed as: cant heat- and mass-transfer limitations leads to com-
parable values of the kinetic data. As expected, the
Ki ) Aie-Ei/(RT), i ) 1, 2 (5) presence of significant intraparticle effects and unmet
heat demands (Antal and Varhegyi (1995)) gives rise
and, from mass conservation: to a process described by a lower activation energy (for
543 K e Tr e 623 K, E ) 191.8 kJ/mol for τ ) 100 µm
K ) K1 + K2 (6) against 200.6 kJ/mol for τ ) 50 µm).
The kinetic data for the rates of tar and the linked
Integration of eqs 2-4 (the subscript ∞ is used to char and gas formation can also be estimated by means
indicate the final yields) with conditions (Agrawal of eqs 8 and 15, if the stoichiometric coefficient νC is
(1988)): known. Alternatively eqs 8 and 11 together with the
measurements of the final tar yields can be used.
MS(0) ) M0, MT(∞) ) MT∞, MC(∞) ) MC∞ (7) Though volatile products have not been collected and
thus the final tar and gas yields (and νC) are not known,
gives an estimation of the kinetic data of the two competitive
reactions has been made on the basis of literature values
MS ) M0e-Kt (8) for νC. The results obtained for νC ) 0.35 (Bradbury et
550 Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997

Figure 10. Plots of ln(P) versus time (reduction in the scale as Figure 12. Experimental (symbols) and simulated (dashed lines,
in Figure 7) for several reaction temperatures (τ ) 50 µm, heating one set of kinetic data) solid mass fraction versus time for τ ) 50
modality B). µm (heating modality B).

experimental weight loss curves shows very good agree-


ment, confirming the validity of the kinetic scheme and
data proposed (Figure 12). On the contrary, the evalu-
ation of the difference between the measured solid mass
fraction and the predictions, based on the two sets of
kinetic constants, shows rather large values at both low
and high temperatures.
The results presented in this study show that for both
the heating modalities A and B, that is, with and
without significant heat-transfer limitations, cellulose
degradation hardly occurs for temperatures above 600
K. In the first case, the increase in the external heat-
transfer rate does not change significantly the temper-
ature of the reacting cellulose, which is cooled by the
Figure 11. Arrhenius plots for the global degradation rate, K, endothermic evolution of volatiles, so that the process
the reaction rates of tar (K2) and linked char and gas (K1) is controlled by heat-transfer (Bradbury et al. (1979),
formation, obtained with the heating modality B for τ ) 50 µm.
Antal and Varhegyi (1995)). In the second case, for high
temperatures, the process becomes very fast and com-
al. (1979)) are also enclosed in Figure 11. The activation plete sample devolatilization is seen for Tr ) 625 K. It
energy of tar formation is significantly higher (225.1 kJ/ is likely that heat- and mass-transfer limitations are
mol, standard deviation 5%) than that of char and gas responsible for the decrease in the activation energy as
formation (158.3 kJ/mol, standard deviation 9%), con- the temperature is increased and for weight loss re-
firming, in agreement with the Broido-Shafizadeh ported at temperatures much higher than it actually
mechanism, that cellulose degradation occurs according occurs (Antal and Varhegyi (1995)). Consequently, the
two competitive pathways, the low temperatures leading temperatures reported for negligible char formation are
to char formation and the high temperatures to tar also higher.
formation. The appearance of a single smooth curve of Some considerations are also needed about the initial
weight loss indicates that the transition from the char- processes which alter the substrate but cause only minor
dominating to the tar-dominating reaction is gradual changes in the weight, described through the formation
(Agrawal (1988)). Again, kinetic data for tar and char of “anhydrocellulose” (Broido and Weinstein (1971)) or
formation are in the range of those previously deter- “active cellulose” (Bradbury et al. (1979)) (loss of water
mined (Bradbury et al. (1979), Agrawal (1988), Arse- and reduction in the degree of polymerization). The
neau (1988), Varhegyi et al. (1994)). descriptions of these processes do not appear necessary
In a recent study (Milosavljevic and Suuberg (1995)) for the prediction of the global degradation kinetics in
a rough classification of global cellulose kinetics has the kinetic regime, as observed by Varhegyi et al. (1994)
been made on the dependence of the thermal conditions. and confirmed by this study. However, changes in the
For rapid heating to temperatures above 600 K, activa- physical properties (i.e., melting) can be important in
tion energies are seen to be in the range 140-155 kJ/ practical situations, where conversion is the result of a
mol, while slow heating to temperatures below 600 K strong interaction between chemical and physical pro-
appears to give rise to larger values (about 218 kJ/mol). cesses such as in the case of ablative pyrolysis. Assum-
Therefore, kinetic constants have also been estimated, ing that the rate of active cellulose formation is propor-
taking the reaction temperature of 600 K as the bound- tional to 1/ti, where ti is the time needed to attain the
ary between two different kinetics (curves are not maximum rate of weight loss (Bradbury et al. (1979)),
reported to avoid crowding). Again, this interpretation the following kinetic data are obtained: E ) 250.5 kJ/
gives data in the same range of those reported in the mol, A ) 1.8 × 1021 s-1. These are again comparable
above reference, that is, E ) 220.5 kJ/mol (A ) 3.47 × to those reported by Bradbury et al. (1979).
1017 s-1) for 523 K e Tr e 600 K and E ) 143.2 kJ/mol
(A ) 3.77 × 1010 s-1) for 600 K e Tr e 623 K. However, Conclusions
standard deviations, equal to 13% and and 10%, respec-
tively, are higher. On the other hand, a comparison A new experimental system has been developed to
between the model predictions through only one set of investigate the kinetics of the thermal degradation of
data (E ) 214.5 kJ/mol, A ) 1.2 × 1017 s-1) and the solids under reduced heat- and mass-transfer limita-
Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997 551
tions. The system, which is based on radiative heat  ) emissivity
transfer, allows for the continuous monitoring of the F ) density
sample temperature and weight, while the heating rates σ ) Stephan-Boltzmann constant
are much faster than those usually encountered in τ ) bed half-thickness
thermogravimetric analyses. ν ) stoichiometric coefficient
Experiments have been carried out for cellulose, taken
as representative of biomass fuels, with two different Subscripts
heating modalities. The first modality (A) consists of C ) char
the sample degradation made to occur under known g ) gas/vapor phase
system temperatures, as in most of the previous studies. S ) cellulose
It is shown that external heat-transfer limitations, T ) tar
which become progressively more important as the
reaction conditions are made more severe, cannot be Literature Cited
avoided for system temperatures above 600 K. Fur-
thermore, for this and higher values, it is confirmed that Agrawal, R. K. Kinetics of reactions involved in pyrolysis of
most of the pyrolytic degradation occurs for effective cellulose I. The three reaction model. Can. J. Chem. Eng. 1988,
66, 403.
reaction temperatures below 600 K, with negligible char
Agrawal, R. K.; McCluskey, R. J. The low pressure pyrolysis of
formation. As a consequence of reduced mass transport newsprint. J. Appl. Polym. Sci. 1983, 27, 367.
limitations, the char yields are significantly lower than Antal, M. J. Biomass Pyrolysis: a Review of the Literature. Part
those reported in the literature for temperatures in the I. Carbohydrate Pyrolysis. In Advances in Solar Energy; Boer
range 543-630 K. K. W., Duffie, J. A., Eds.; American Solar Energy Society:
The second heating modality (B) represents the most Boulder, CO, 1982; Vol. 1, p 61.
innovative aspect of the study. It is based on the control Antal, M. J. Biomass Pyrolysis: a Review of the Literature. Part
II. Lignocellulose Pyrolysis. Advances in Solar Energy; Boer,
of the sample temperature, using the applied radiative K. W., Duffie, J. A., Eds.; American Solar Energy Society:
heat flux as the manipulated variable, in order to Boulder, CO, 1985; Vol. 2, p 175.
eliminate external heat-transfer limitations. The pro- Antal, M. J.; Friedman, H. L.; Rogers, F. E. Kinetics of cellulose
cess is two-stage. In the first, the sample is rapidly pyrolysis in nitrogen and steam. Combust. Sci. Technol. 1980,
brought to the chosen sample temperature (set point), 21, 141.
while in the second the temperature is maintained at Antal, M. J.; Varhegyi, G. Cellulose Pyrolysis Kinetics: the Cur-
the chosen value. The global heating rate has been rent State of Knowledge. Ind. Eng. Chem. Res. 1995, 34, 703.
Antal, M. J.; Mok, W. S. L.; Varhegyi, G.; Szekely, T. Review of
varied from about 19 to 56 K/s for cellulose setpoints in Methods for Improving the Yield of Charcoal from Biomass.
the range 523-623 K. Sample half-thicknesses below Energy Fuels 1990, 4, 221.
60 µm guarantee the absence of intrabed temperature Arseneau, D. F. Competitive reactions in the thermal degradation
gradients and activity of secondary charring reactions. of cellulose. Can. J. Chem. 1971, 49, 632.
Therefore, the process is made to occur under known Barooah, J. N.; Long, V. D. Rates of thermal decomposition of some
thermal conditions and reduced heat- and mass-transfer carbonaceous materials in a fluidized bed. Fuel 1976, 55, 116.
limitations. Bradbury, A. G. W.; Sakai, Y.; Shafizadeh, F. A Kinetic Model for
Pyrolysis of Cellulose. J. Appl. Polym. Sci. 1979, 23, 3271.
Cellulose weight loss curves have been interpreted in Broido, A.; Weinstein, M. Kinetics of Solid-Phase Cellulose Py-
terms of a single-step process (consisting of two com- rolysis. In Proceedings of the 3rd International Conference on
petitive reactions leading to tar and linked gas and char Thermal Analysis; Wiedemann, H. G., Ed.; Birkhauser Verlag:
formation), and the estimated kinetic data are in Basel, Switzerland, 1971; p 285.
agreement with previous studies conducted under ki- Cullis, C. F.; Hirschler, M. M.; Townsend, R. P.; Visanuvimol, V.
netic control. The pyrolysis of cellulose under conditions of rapid heating.
Combust. Flame 1983, 49, 235.
Di Blasi, C. Modeling and Simulation of Combustion Processes of
Acknowledgment Charring and Non-Charring Solid Fuels. Prog. Energ. Combust.
Sci. 1993, 19, 71.
The research was funded in part by the European Di Blasi, C. Numerical simulation of cellulose pyrolysis. Biomass
Commission in the framework of the Non Nuclear Bioenergy 1994, 7, 87.
Energy Programme (JOULE III), Contract JOR3-CT95- Di Blasi, C. Influences of model assumptions on the predictions
0081. of cellulose pyrolysis in the heat transfer controlled regime. Fuel
1996a, 75, 58.
Nomenclature Di Blasi, C. Heat, momentum and mass transfer through a
shrinking biomass particle exposed to thermal radiation. Chem.
A ) preexponential factor Eng. Sci. 1996b, 51, 1121.
c ) specific heat Di Blasi, C. Kinetic and heat transfer control in the slow and flash
pyrolysis of solids. Ind. Eng. Chem. Res. 1996c, 35, 37.
E ) activation energy
Di Blasi, C. Heat transfer mechanisms and multi-step kinetics in
M ) mass the ablative pyrolysis of cellulose. Chem. Eng. Sci. 1996d, 51,
MC∞ ) final solid char mass 2211.
M0 ) initial cellulose mass Di Blasi, C.; Crescitelli, S.; Russo, G.; Cinque, G. Numerical model
Qe ) radiative heat flux of ignition processes of polymeric materials including gas phase
Qr ) enthalpy variation due to chemical reactions absorption of radiation. Combust. Flame 1991, 83, 333.
S ) heat-exposed surface Hajaligol, M. R.; Howard, J. B.; Longwell, J. P.; Peters, W. A.
Tf ) system temperature Product Compositions and Kinetics for Rapid Pyrolysis of
TL ) lamp temperature Cellulose. Ind. Eng. Chem. Process Des. Dev. 1982, 21, 457.
Tr ) (final) sample temperature Lede, J.; Panagopoulos, J.; Li, H. Z.; Villermaux, J. Fast Pyrolysis
of Wood: Direct Measurement and Study of Ablation Rate. Fuel
tc ) conversion time 1985, 64, 1514.
th ) heating time Lewellen, P. C.; Peters, W. A.; Howard, J. B. Cellulose Pyrolysis
ts ) solid residence time Kinetics and Char Formation Mechanism. 16th Symposium
V ) sample volume (International) on Combustion; The Combustion Institute: Pitts-
W ) solid residual (including char and cellulose) burgh, PA, 1976; p 1471.
552 Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997

Lipska, A. E.; Parker, W. J. Kinetics of the pyrolysis of cellulose Shafizadeh, F.; Furneaux, R. H.; Cochran, T. G.; Scholl, J. P.;
in the temperature range 250-300 °C. J. Appl. Polym. Sci. Sakai, Y. Production of Levoglucosan and Glucose from Pyroly-
1966, 10, 1439. sis of Cellulosic Materials. J. Appl. Polym. Sci. 1979, 23, 3525.
Madorsky, S. L.; Hart, V. E.; Straus, S. Pyrolysis of cellulose in a Tabatabaie-Raissi, A.; Mok, W. S. L.; Antal, M. J. Cellulose
vacuum. J. Res. Natl. Bur. Stand. 1956, 56, 343. pyrolysis kinetics in a simulated solar environment. Ind. Eng.
Milosavljevic, I.; Suuberg, E. M. Cellulose thermal decomposition Chem. Res. 1989, 28, 856.
kinetics: global mass loss kinetics. Ind. Eng. Chem. Res. 1995, Thurner, F.; Mann, U. Kinetic investigation of wood pyrolysis. Ind.
34, 1081. Eng. Chem. Process Des. Dev. 1981, 20, 482.
Narayan, R.; Antal, M. J. Thermal lag, fusion, and the compensa- Varhegyi, G.; Jakab, E.; Antal, M. J. Is the Broido-Shafizadeh
tion effect during biomass pyrolysis. Ind. Eng. Chem. Res. 1996, model for cellulose true? Energy Fuels 1994, 8, 1345.
35, 1711. Wanzl, W. Techniques for studying and modeling coal pyrolysis
Ozturk, Z.; Merklin, J. Rapid pyrolysis of cellulose with reac- and their relevance to biomass and wastes. Biomass Bioenergy
tive hydrogen gas in a single-pulse shock tube. Fuel 1995, 74, 1994, 7, 131.
1658. Received for review September 9, 1996
Ramiah, M. V. Thermogravimetric and differential thermal analy- Revised manuscript received November 19, 1996
sis of cellulose, hemicellulose, and lignin. J. Appl. Polym. Sci. Accepted November 20, 1996X
1970, 14, 1323.
Scott, D. S.; Piskorz, J.; Bergougnou, M. A.; Graham, R.; Overend, IE960551R
R. P. The role of temperature in the fast pyrolysis of cellulose
and wood. Ind. Eng. Chem. Res. 1988, 27, 8.
Seborg, D. E.; Edgar, T. F.; Millichamp, D. A. Process Dynamics X
Abstract published in Advance ACS Abstracts, January 15,
and Control. Wiley Ser. Chem. Eng. 1989. 1997.

S-ar putea să vă placă și