Vous êtes sur la page 1sur 11

Journal of the Air & Waste Management Association

ISSN: 1096-2247 (Print) 2162-2906 (Online) Journal homepage: http://www.tandfonline.com/loi/uawm20

Combustion and Gasification Properties of Plastics


Particles

Ron Zevenhoven , Magnus Karlsson , Mikko Hupa & Martin Frankenhaeuser

To cite this article: Ron Zevenhoven , Magnus Karlsson , Mikko Hupa & Martin Frankenhaeuser
(1997) Combustion and Gasification Properties of Plastics Particles, Journal of the Air & Waste
Management Association, 47:8, 861-870, DOI: 10.1080/10473289.1997.10464461

To link to this article: https://doi.org/10.1080/10473289.1997.10464461

Published online: 27 Dec 2011.

Submit your article to this journal

Article views: 2127

Citing articles: 20 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=uawm20
TECHNICAL PAPER Zevenhoven, Karlsson, Hupa,
ISSN 1047-3289 J. Air & and Frankenhaeuser
Waste Manage. Assoc. 47:861-870
Copyright 1997 Air & Waste Management Association

Combustion and Gasification Properties of Plastics Particles


Ron Zevenhoven, Magnus Karlsson, and Mikko Hupa
Åbo Akademi University, Department of Chemical Engineering, Turku, Finland

Martin Frankenhaeuser
Borealis Polymers Oy, Porvoo, Finland

ABSTRACT 99% hydrocarbon. The gasification rate of PVC char (at 1


The combustion and gasification behavior of the most bar and 25 bar) was of the same order as that of char from
common plastics is studied and compared with conven- coal. Peat-char and wood-char were gasified an order of
tional fuels such as coal, peat, and wood. The aim is to magnitude faster.
give background data for finding the optimum conditions
for co-combustion or co-gasification of a conventional fuel INTRODUCTION
with a certain amount of plastic-derived fuel. Atmospheric In western Europe 6-10% of the municipal waste is com-
or pressurized fluidized bed co-combustion of conven- posed of plastics (9.3 million tons in 1992), the largest
tional fuels and plastics are considered to be promising part (72%) of which is disposed of by landfill. In Finland
future options. The plastics investigated were the annual yield of plastics in municipal waste is 6%,
poly(ethylene) (PE), poly(propylene) (PP), poly(styrene) which equates to 90,000 tons.1 Increased environmental
(PS), and poly(vinyl chloride) (PVC). Some of the samples consideration and stricter regulation during recent years
had a print or color. The reference fuels were Polish bitu- is resulting in efforts to reduce landfill disposal of plastic
minous coal, Finnish peat, and Finnish pine wood. waste and increase recycling or recovery.
PE, PP, and PS were found to burn like oil. The par- Recovery of resources contained in waste plastics can
ticles shrank to a droplet and burned completely during be separated into either product recycling (e.g., plastic
the pyrolysis stage, leaving no char. Printing and color- bottles), material recycling (in the form of plastic particles),
ing left a small portion of ash. PVC was the only plastic feedstock recycling (depolymerisation intermediates or
that produced a carbonaceous residue, and its timescales monomers), or energy recovery. Energy recovery is stan-
for heating, devolatilization, and char burning were of dard procedure in municipal waste incineration; older
the same order as those for peat and wood, and much incinerators are facing operational problems associated
shorter for the other plastics studied. An important result with the increase of plastic content in MSW (municipal
is that char from PVC contains less than 1% chlorine, solid waste). The reason for this is the high calorific value
of plastics (of the order of 40 MJ/kg), which reduces
throughput in a heat restricted conventional waste incin-
IMPLICATIONS
erator.2 Therefore, there is increasing interest in methods
Landfill disposal of the plastics in municipal solid waste
(MSW) is no longer in line with increasing environmental in which plastic waste is separated from the municipal
consideration and stricter regulations. Apart from material waste and added as a prepared fuel (up to 30%) to the
recovery of plastics in the form of product recycling, mate- solid fuels that are burned in power plants for thermal
rial recycling, and feedstock recycling, the energy recovery power and electricity production.
from plastics in waste is of increasing interest. However, The composition of municipal waste in western Eu-
recovery requires combustion methods that are more ad-
rope (in 1992) is given in Figure 1 and can be separated
vanced than conventional incineration because the high
calorific value of plastics or MSW with increased plastic into three parts: an organic fraction that can be
content can only be exploited efficiently in systems that are composted, a combustible fraction mainly composed of
specially designed and optimized for that purpose. Basic plastics and paper, and a non-combustible fraction of
data on the combustion and gasification behavior of sev- metals, minerals, etc. Depending on the separation
eral types of plastics are given here, using laboratory scale
method, the combustible fraction is referred to as either
testing. A comparison is made with conventional fuels. This
information may support the design of future incinerators in refuse derived fuel (RDF) or packaging derived fuel (PDF).
which plastics are co-fired with conventional fuels. RDF is municipal waste from which organic waste and
non-combustible material have been separated, and PDF

Volume 47 August 1997 Journal of the Air & Waste Management Association 861
Zevenhoven, Karlsson, Hupa, and Frankenhaeuser

Figure 2. The distribution of plastics consumption in western Europe,


Figure 1. The composition of municipal waste in western Europe, 1992.1
1992. 1

is obtained from the separated collection of combustible of PVC and PS to form soot. Kaminsky3 found up to 56%
waste, which is mainly paper and plastic packaging. Be- HCl and 9% solid char residue resulting from pyrolysis of
cause of their calorific value, both RDF and PDF must be PVC in a fluidized bed at 740 °C. Under the same condi-
considered as high-grade fuels, although the technolo- tions, PS gave 65% styrene and 0.6% solid residue; for PE
gies for optimal large-scale combustion or gasification still and PP the solid residues were 1.8 and 1.6% of the initial
need basic research directed toward thermal efficiency of mass, respectively. Darivakis et al.4 pyrolysed PE and PS
the system, operational problems, and emission control samples at a heating rate of 1,000 K/s and report a com-
of pollutants. plete pyrolysis without solid residue for final temperatures
The present work provides results of a laboratory study above 700 °C.
in which the most commonly encountered plastics are Panagiotou and Levendis10 investigated the combus-
characterized with respect to their behavior during py- tion properties of PS, PVC, PE, and PP. Spherical particles
rolysis, gasification, and combustion compared with ranging in size from 53 to 300 µm were burnt at 930-
conventional fuels such as wood, peat, and coal. Consid- 1,230 °C in 21, 50, and 100% oxygen in an atmospheric
ering the distribution of the consumption of base plastics drop-tube furnace. Time required for combustion was
in western Europe in 1992, (Figure 2), it was decided that found comparable to that of an equally sized droplet of
poly(ethylene), poly(propylene), poly(styrene), and light oil. PVC and PS burned with a luminous yellow
poly(vinyl chloride) would be considered. Throughout this flame, producing a lot of soot. PVC gave the shortest time
paper, these compounds will be referred to by their ab- for combustion, while PE and PP appeared to burn more
breviations PE, PP, PS, and PVC, respectively. slowly and with a less clear flame.
The goal of most of the work on the thermal decom- McGhee et al.11 gave a very simple model for the py-
position of plastics reported in literature thus far has been rolysis of municipal waste on the basis of co-pyrolysis tests
to retrieve monomers or other valuable products through with PVC and cellulose-based material. It was concluded
thermal processes in various types of reactors. In that work, that co-pyrolysis gives an increased amount of solid char
the temperature levels and heating rates were generally residue with a lower reactivity as compared to the char
lower than in technical combustion systems.3-6 The pro- residue of the separate materials. Pyrolysis of RDF at tem-
duction of liquid products by thermal decomposition of peratures between 500 and 900 °C was recently reported
polymers in a fluidized bed at temperatures up to 850 °C by Cozzani et al.12,13 It was found that interactions be-
was extensively studied in the United States.7 tween RDF components (plastics, paper, and cardboard,
Mechanisms for thermal decomposition of plastics wood-like materials) can be neglected with respect to solid-
and pyrolysis product compositions have been given by phase reactions. It was also stated that methane and light
Cullis and Hirschler.8 The pyrolysis of PE was found to hydrocarbons are formed in gas-phase reactions at suffi-
give alkanes and alkenes, while PP gave a very small ciently high temperatures.
amount of its monomer during pyrolysis. PS yields more Frankenhaeuser et al.14,15 investigated the co-combus-
than 50% styrene during pyrolysis, together with its tion of PDF and RDF with coal, peat, and wood in a full-
dimers and trimers. The pyrolysis of PVC gives HCl and scale circulating fluidized bed combustor (65 MWth). Tests
aromatic hydrocarbons (e.g., benzene), plus a consider- showed no significant harmful side effects with respect
able amount of char. Elomaa9 also reported the tendency to pollution as compared to firing with conventional fuel.

862 Journal of the Air & Waste Management Association Volume 47 August 1997
Zevenhoven, Karlsson, Hupa, and Frankenhaeuser

Figure 3. The experimental set up for the video recordings.

Emissions (less than 500 mg SO2/m3STP and less than 200


mg NO2/ m3STP) were of the same level as obtained from
conventional fuels although HCl emissions increased with
increasing chloride content in the fuel mixture. Also, the
emissions of poly-chlorinated di-benso-p-dioxines and di- Figure 4. The pyrolysis reactor.

benso-furanes (PCDD/F) did not correlate with the chlo-


ride content of the fuel mixture, but were related more to co-combustion or co-gasification of plastics waste with
the combustion process conditions. Fuel sulfur clearly peat, wood, or coal. More specifically, atmospheric or pres-
appeared to have a reducing effect on PCDD/F emissions. surized, bubbling or circulating, fluidized bed combus-
This was also found with limestone added to the bed for tion or gasification are considered to be the major large-
the capture of SO2 and HCl (Kojo16), thus giving increased scale applications, which explains why experimental con-
PCDD/F concentrations in the ash. No increased levels in ditions chosen here correspond to these systems: tempera-
15
corrosion were found. tures 750-950 °C, pressure up to 25 bar, and heating rates
The co-gasification of municipal waste and wood in on the order of 100-1,000 K/s.
an atmospheric fluidized bed was investigated by Czernik
et al.17 Wood waste and wood waste containing 10% PE EXPERIMENTAL PROCEDURES
and RDF were gasified at 50 kg/h. The facility operated Atmospheric Pressure Laboratory Furnace with
satisfactorily at 700-850 °C, producing 1.7-2.4 m3STP dry Video Recording
gas/kg dry solids at a calorific value of 5.2-8.2 MJ/ m3STP. An atmospheric pressure, electrically heated laboratory
Recently, Wey and Chang18 analyzed the incineration furnace (Multitherm® N7) suitable for temperatures up to
kinetics of PP, PE, HDPE, and ABS (acrylonitrile-butadi- 1,260 °C, with an internal volume of 7 liters was used
ene-styrene) particles (8-20 mm) at temperatures between (Figure 3). PtRh/Pt thermocouples were used for furnace
550 and 800 °C. At this temperature level the time
to ignition dominated. Because of the relatively low Table 1. The plastics and other fuel samples.
temperatures, accumulating combustible gases
LDPE (low density PE) sheet 0.7 mm thick density 924 kg/m3 666 g/m2
tended to exhibit explosive behavior. The combus- 3
HDPE (high density PE) film 16 µm thick density 952 kg/m 15.3 g/m2
tion process was modeled using Spalding’s approach HDPE printed as 2. with 10-µm color print on one side 25.0 g/m2
to droplet combustion,19 leading to values for the PP sheet 1 mm thick density 910 kg/m 3
912 g/m2
Spalding transfer number, B. This model will also PP white 3% wt color* sheet 0.45 mm thick density 950 kg/m 3
429 g/m2
be discussed and used in this paper. PS white 3% wt color* sheet 0.52 mm thick density 1,070 kg/m 3
557 g/m2
3
Objectives for the work reported here were two- PVC sheet 0.3 mm thick density 1,370 kg/m 417 g/m2
fold. The first objective was to verify if the test Polish bituminous coal particle size 0.5-0.7 mm
methods for fuel characterization used in our labo- Finnish peat particle size up to 1 mm
ratory were suitable for plastic fuels. A second objec- Finnish pine wood fibers 3 mm long and 0.5 mm thick
tive was to obtain quantitative data to be used in
assessing the possibilities for and consequences of * 50% TiO2 and 50% organics

Volume 47 August 1997 Journal of the Air & Waste Management Association 863
Zevenhoven, Karlsson, Hupa, and Frankenhaeuser

(a) Table 2. Proximate analysis of the fuels (% wt in dry samples, Cfix determined from
char, not by difference).

Volatiles Ash Cfix Moisture HHV


(MJ/kg)

LDPE 99.9 <0.05 0.10 46.0


HDPE 99.9 <0.05 0.16 43.1
HDPE printed 97.4 2.33 0.34 0.21 41.9
PP 99.9 <0.05 0.13 46.5
PP white 97.8 2.21 0.10 46.7
PS white 97.0 2.89 0.11 40.6
PVC 92.3 <0.05 7.5 0.18 21.2
Coal 28.3 18.4 47.9 2.35 26.2
(b) Peat 72.1 2.35 23.2 27.9 24.1
Wood 85.8 0.45 12.6 6.15 19.1

Pressurized Thermogravimetric Reactor for


Pyrolysis Yield and Char Reactivity Measurement
A pressurized thermogravimetric reactor (P-TGR) for fur-
nace temperatures and pressures up to 1,100 °C and 100
bar was used (Figure 5(a)). Four gases plus steam and a
purge gas could be fed to the furnace. The inner diam-
eter of the reactor was 17 mm. The sample could be low-
ered from a water-cooled and purged top section into
Figure 5. (a) pressurized thermobalance: experimental set up, (b)
thermobalance sample holder. the reactor by an electrical spindle within 10 seconds.
The mass of the filled sample holder (Figure 5(b)) was
temperature measurement and control. The gas flow measured by a Sartorius 4406 MP8 balance (maximum
through the furnace was 100 or 200 litersSTP/h. Samples of mass 1,100 mg, resolution 0.01 mg). The flow through
one piece of material or a specific amount of it were fed the reactor was 200 litersSTP/h. In this facility char gasifi-
to the furnace from the top in a sample holder. Through cation reactivities were measured as a function of tem-
a quartz window the progress of the conversion process perature and pressure. The char was produced by lower-
was recorded on VHS video tape. From these recordings ing the sample (approximately 200 mg) into the heated
the times to ignition, the duration of the pyrolysis (i.e., (pressurized) reactor section with a gas composition of
volatiles burning), and the duration of char burning was 95% N2 and 5% CO. After approximately 400 seconds a
determined as a function of furnace temperature (750, stable weight signal was obtained; i.e., the end of the
850, and 950 °C), oxygen content (0, 2, 7, and 12% O2), devolatilization stage. The gas composition was then
and sample mass (30 to 70 mg).

Atmospheric Pressure Laboratory Reactor for


Char Production
An atmospheric pressure laboratory reactor designed at
our laboratory was composed of a quartz tube (length 600
mm, diameter 25 mm) and an electrically heated furnace
allowing for temperatures up to 1,000 °C (Figure 4). The
gas flow used here was 100 litersSTP/h, which corresponds
to a flow velocity of approximately 0.2 m/s. In this device
the samples (30-80 mg) were lowered into a nitrogen purge
and removed after 15 or 30 seconds. The char yield was
determined from the mass change. Furthermore, some of
the chars were submitted to chemical analysis. Tests were
Figure 6. Duration of the combustion stages at 850 °C, 7% oxygen.
done at reactor temperatures 750, 850, and 950 °C.

864 Journal of the Air & Waste Management Association Volume 47 August 1997
Zevenhoven, Karlsson, Hupa, and Frankenhaeuser

RESULTS AND DISCUSSION


Table 3. Chemical analysis of the plastics and other fuels (% wt in dry samples, oxygen measured, not by difference).
Measurement of
C H N O* S Cl Ti Sn** Total Conversion Rates from
Video Recordings
LDPE 85.7 14.3 0.16 100 A first set of video recordings
HDPE 85.6 14.2 0.30 100
was made to measure the rate
HDPE printed 82.4 13.6 0.34 1.28 97.6
of pre-ignition processes, py-
PP 85.5 14.3 0.19 100
rolysis rate, and char burnout
PP white 83.6 14.1 0.18 1.31 99.2
PS white 88.9 8.3 0.16 1.69 99.1
rate for all fuels at various
PVC 40.1 5.1 0.65 53.8 0.20 99.9 temperatures and oxygen con-
Coal 66.6 4.0 1.07 8.85 1.22 81.7 centratio ns. The start and
Peat 57.8 6.1 1.79 31.5 0.19 97.4 completion of the pyrolysis
Wood 48.9 6.0 0.17 43.8 0.06 99.0 stage were determined by the
moment when ignition of the
* Excluding oxygen from TiO2. **Measured since Sn is a stabilizer in PVC. volatiles was visible and the
moment at which this flame
instantaneously changed to 80% N2, 18% CO2, and 2% extinguished, respectively. Figure 6 shows the dura-
CO. The average gasification rate was then extracted tion of the stages for 50 mg samples at 850 °C in 7%
from the weight vs. time curve of the gasification stage, O2 in N2. An example of the images that were recorded
as will be discussed. is given in Figure 7 for the LDPE, PVC, and the wood
samples, respectively. The results show that there is very
Temperature
The temperatures referred to are furnace temperatures. In
all cases the process is practically isothermal because of the
small heat effect of the chemical reactions that is exchanged
with a large amount of (flowing) gas. In the most reactive
case (i.e., the combustion of plastics at 12% oxygen), the
maximum possible temperature deviation that can be
reached is less than 10 °C.

Plastics and Other Fuel Samples


In this study seven types of plastic materials and three types
of conventional fuel (coal, peat, and wood) were used (Table
1). The results of standard tests (done at an external labora-
tory) for proximate analysis and chemical analysis are given
in Table 2 and 3, respectively.

Definitions Used in Data Processing


• Char. Solid residue (here including inorganic non-
combustibles) after devolatilization in chemically
inert atmosphere.
• Soot. Carbonaceous material released from the sample
and visible as condensed particles.
• Time to ignition. Recorded time between inserting fuel
particle into the furnace and first appearance of a
visible flame.
• Pyrolysis duration. Time during which a flame is
visible.
• Char burnout time. Time between extinction of the
flame and termination of the particle combustion
process.
• Time for combustion. Total of time to ignition, Figure 7. Examples of recorded video images for LDPE, PVC, and
pyrolysis duration, and char burnout time. wood; temperature 850 °C; oxygen 7%. Note different times.

Volume 47 August 1997 Journal of the Air & Waste Management Association 865
Zevenhoven, Karlsson, Hupa, and Frankenhaeuser

little difference between the plastic’s samples, with the showing a faster ignition and shorter pyrolysis than the
exception of PVC, which is the only plastic that produces other plastics. This faster ignition can be explained by
a “char.” This “char” may, in fact, be a re-condensed soot the “activation energy for thermal degradation,”21 which
particle, formed from methyl-, ethyl-, or vinyl-substituted is lower for PVC (85-140 kJ/mol) than for the other plas-
mono- and poly-nuclear aromatics, according to tics studied here (200-300 kJ/mol). Overall, the combus-
Knümann and Bockhorn.20 It is also concluded that the tion of PVC is similar to that of wood and peat from a
shape of the plastic’s samples had little effect on the be- visual point of view.
havior recorded. The initial surface of the HDPE sheet was The effect of the oxygen concentration was as ex-
significantly larger than that of the other materials that pected: a higher oxygen concentration speeds up com-
were 10 to 40 times thicker. In several cases, a re-shaping bustion as shown in Figure 9. There was no effect on the
of the plastics sample into a molten spherical particle was rate of pre-ignition processes nor on the pyrolysis rate,
seen during the heating and the early stages of the py- indicating that intra-particle mass (and/or heat) transfer
rolysis. is limiting. Measurements at 2% O2 gave an unstable
The effect of temperature was studied in a second flame and a very slow char burning. Tests in pure nitro-
set of experiments. Figure 8 shows the rates of the differ- gen did not produce any visible phenomena except the
ent stages in combustion at 750, 850, and 950 °C, re- PVC, which showed slight swelling.
spectively, in 7% O2 in nitrogen. From this, again it is Varying the sample size gives information on the
shown that PVC is the only plastic that produces a char, effect of intra-particle heat and mass transfer limitations.
An increasing sample mass, from 10 to 70 mg, (Figure
10) shows a significant effect on the pyrolysis rate and
char burnout rate. This shows that intra-particle and
external heat transfer limitations can both be neglected.
Pyrolysis and char burnout rate, however, roughly de-
crease linearly with mass to the power 2/3; e.g., increas-
ing the mass from 10 to 70 mg leads to a three to four
times lower char burnout rate. This indicates conversion
rates depending on particle external surface; i.e., oxygen
diffusion limitation.
The combustion behavior of the particles that did
not produce a char is similar to fuel droplet combus-
tion. According to an approach by Spalding, 18,19 this
can be modeled by the following rate equation for the
simultaneous pyrolysis and gas-phase combustion of a
spherical droplet:

Figure 8. Rates for the conversion stages at 750, 850, and 950 °C in Figure 9. Char burnout rate and overall rate at 850 °C in 2, 7, or 12%
7% oxygen in nitrogen. oxygen.

866 Journal of the Air & Waste Management Association Volume 47 August 1997
Zevenhoven, Karlsson, Hupa, and Frankenhaeuser

(a) (b)

Figure 11. Spalding transfer number at 7% oxygen vs. temperature


for the plastics that did not produce a char.

ρs
d0
dp
1n(1 + B) = ∫ Shdd p (2)
2ρg Dg t pyrolysis o

where tpyrolysis is the duration of the pyrolysis process during


which a flame is visible. The Spalding number gives infor-
mation on the mass fraction of fuel at the burning particle’s
surface yf,s as compared to that in the bulk of the droplet yf,b
and in the bulk gas phase yf,∞. With yf,b = 1 in the particle
and yf,∞ = 0 in the bulk gas, the value for B equals :
y f,s − y f ,∞ y f,s
Figure 10. (a) average pyrolysis rate vs. sample size at 850 °C, 7% B= = (3)
y f,b − y f,s 1 − y f,s
oxygen, (b) average char burnout rate vs. sample size at 850 °C, 7%
oxygen. A value B = 1 implies that at the particle’s surface, the fuel
mass fraction equals 50%. The values for B were calculated for
dm ρsd p dd p ρg DgSh 1n(1 + B) those fuels that did not produce a char. For an initial particle
− =− = (1) mass of 30 mg (which corresponds to approximately 4 mm
Sdt 6 dt 2d p particle size when molten to sphere) and 7% oxygen in nitro-
gen at 750, 850, and 950 °C, the values for B are given in Figure
h = 2 + 06
SH . Re 1 2 Sc 1 3
11. Each point corresponds to an average value from 10 plastic
where particles. The value for B appeared to be sensitive to the initial
particle size and oxygen concentration. The comparison of
m = particle mass: kg our values for B to those reported by Wey and Chang,18 shows
dp = particle diameter: m similar values for PP and PE even though the lines plotted after
S = particle surface: m2 Wey and Chang were determined for the temperature range
t = time: s 550-800 °C.
ρs = droplet density: kg/m3
ρg = gas density: kg/m3 The Measurement of Char Yields,
Dg = diffusivity of O2 in N2: m2/s Char Composition
Sh = Sherwood number The result of the devolatilization during 15 s of 50 mg
Re = Reynolds number samples in a nitrogen purge flow at 850 °C is given in
Sc = Schmidt number Figure 12. The distinction between char and ash was made
B = Spalding transfer number using the ash content from the proximate analysis of the
fuels (Table 2). Only the reference fuels produce a residue
Rewriting this equation gives a relation from which the B that contained char and ash. The PVC plastic produces a “char”
can be found:18 without ash, while only the printed plastics produce an ash

Volume 47 August 1997 Journal of the Air & Waste Management Association 867
Zevenhoven, Karlsson, Hupa, and Frankenhaeuser

Table 4. Elemental analysis of the solid residues from devolatilization in N2 (% wt and two pressures (1 bar, 25 bar) in 95% N2, 5% CO, with the
in dry samples, oxygen measured, not by difference). LDPE, HDPE printed, PVC, PS, and peat samples. From earlier
tests, data on wood and coal at 1 bar and 20 bar were available.
C H N O Cl Sn Table 5 gives the amount of char residue found, with the re-
sults of the atmospheric tests in the entrained gas flow reactor
PVC 750 °C 92.2 6.65 0.79* 0.31* <0.4*
(see previous section) included. A clear effect of pressure is seen:
PVC 850 °C 92.5 2.46 1.98* 0.14* <0.4*
increased pressure hinders the devolatilization. Furthermore,
PVC 950 °C 94.4 1.68 2.00* 0.21* <0.4*
HDPE printed 850 °C 76.4 11.3 0.35 1.54
no significant amounts of gasifiable material were found for
HDPE printed 950 °C 20.4* 0.96* 1.02* 6.51* the plastics except for the PVC. Therefore, gasification tests
coal 750 °C 71.3 2.83 7.43 were only done with the PVC and peat.
coal 850 °C 66.9 2.00 4.64 The gas phase during the gasification was 18%
coal 950 °C 73.6 1.55 2.56 CO2, 2% CO, and 80% N2; temperature and pressure were
peat 750 °C 71.6 3.10 16.3 the same as during the pyrolysis phase. Figures 14(a)-(c)
peat 850 °C 78.3 2.19 9.76 give time conversion curves for the chars, and Figure
peat 950 °C 80.6 1.36 8.49 14(d) gives the rate conversion curve corresponding to
wood 750 °C 80.6 2.96 12.5 the data in Figure 14(b).
wood 850 °C 87.2 2.26 6.40
The average gasification rate, r* (unit %/min), is found
wood 950 °C 88.5 1.34 5.86
by fitting a 2-parameter model equation to the time-char con-
version data:
* All results are averaged from two tests, except * for which too little material was
available.
X char (t ) = 1 − exp (− at bb )
without char. The amount of ash found from the colored PP
1
and PS corresponded to the mass of inorganic color agent. 1−
1 dX char  1  b
There is an effect of temperature on the char yield for rate r= = ba 1/b  1n  (4)
1 − Xchar dt  1− X 
the fuels that produce a char, (Figure 13) for a 50 mg sample
in nitrogen during 15 s at 750, 850, and 950 °C. The char where Xchar(t) is calculated from the time-mass data:
yield also was measured after 30 s and by varying the sample
size between 30 and 80 mg. No effect of either variation m initial − m(t)
was found. The results from elemental analysis of the chars X char (t) = (5)
m initial − m ash
produced from the 15 s devolatilization tests are given in
Table 4. An interesting result is that the “char” from PVC mash is the mass of ash in the sample, calculated as sample
does not contain significant amounts of chlorine. mass × proximate ash content. The average char gasifica-
tion rate is then determined by the parameters a and b:
Measurement of Char Yield and Char Reactivity
in a Pressurized Thermobalance r ∗ = a 1 / b b Γ(2 − 1 / b) (6)
Char yield measurements were carried out in the pressurized
thermobalance at two temperature levels (850 and 950 °C)

Table 5. Char + ash yield measured after pyrolysis in a pressurized thermobalance.

Thermobalance Tube reactor

1 bar 850 °C 25 bar 850 °C 1 bar 950 °C 25 bar 950 °C 1 bar 850 °C
char char char char char
% wt % wt % wt % wt % wt

PVC 5.9 15.5 5.1 13.6 4.6


LDPE 0.2 0.4
HDPE printed 2.3 2.6
PS 2.1
coal 69.4 72.4 (20 bar) 71.6
peat 24.5 37.4 25.6 30.5 28.2
wood 13.3 23.0 (20 bar) 10.9 Figure 12. Char + ash yield for all fuels after devolitization in nitrogen
during 15 s at 850 °C. (Ash fraction from proximate analysis.)

868 Journal of the Air & Waste Management Association Volume 47 August 1997
Zevenhoven, Karlsson, Hupa, and Frankenhaeuser

Figure 13. Char + ash yields after devolitization in nitrogen for coal,
peat, wood, and PVC: 15 s, 750, 850, 950 °C.

with the gamma function defined as Γ(x + 1) = x!. It is


seen that b = 1 gives r* = a. The average gasification
reactivities that were determined are given in Table 6.
These results show that the gasification rate for the
char from PVC is more than 10 times slower than the
gasification rate of a char from wood and peat, and
roughly of the same order as the gasification reactiv-
ity of coal char. Also, increased pressure has a (weak)
enhancing effect on the char gasification rate.

CONCLUSIONS
Having studied the combustion properties of seven
types of plastics (based on four different polymers) with
three reference fuels, it can be concluded that the ex-
perimental facilities used are well suited for the broad
range of analyses presented here. Valuable informa-
tion on the burning behavior of plastics was obtained.
Pyrolysis and simultaneous burnout in the gas phase
was rate controlling, except for those fuels that give a
combustible solid residue after pyrolysis. It was found
that PE, PP, and PS without a color agent or printing
burned like oil droplets, without leaving a char. The ashes
found related directly to color agents and inks. Due to
the higher temperatures and smaller particles as com-
pared to Wey and Chang,18 pre-ignition processes were
not limiting in our tests. Pyrolysis and char burnout

Table 6. Average gasification rates measured in the pressurized thermobalance.

Average (%/min)
gasification rate

1 bar 850 °C 25 bar 850 °C 1 bar 950 °C 25 bar 950 °C


PVC 0.26 0.40 2.62 4.38
Coal 0.02 0.23 (20 bar)
Peat 7.91 11.6 47.6 64.6 Figure 14. (a) CO2 gasification of chars at 850 °C 1 bar, (b) CO2
Wood 3.64 5.47 (20 bar) gasification of chars at 850 °C 20/25 bar, (c) CO2 gasification of chars
from peat and PVC at 950 °C 1 and 25 bar, (d) CO2 gasification rate of
chars at 850 °C 20/25 bar.

Volume 47 August 1997 Journal of the Air & Waste Management Association 869
Zevenhoven, Karlsson, Hupa, and Frankenhaeuser

rate, however, roughly decrease linearly with mass to Finland, for funding this project. Kristoffer Sandelin is ac-
the power 2/3; e.g., increasing the mass from 10 to 70 knowledged for his experimental work on the pressurized
mg leads to a three to four times lower char burnout thermobalance. Ron Zevenhoven acknowledges the Com-
rate. This indicates conversion rates depending on mission of the European Communities for support via Hu-
particle external surface; i.e., oxygen diffusion limita- man Capital and Mobility post-doc fellowship.
tion. A droplet combustion model was applied to the
plastics that did not produce a residue after pyrolysis, REFERENCES
1. Information System on Plastics Wastes Management in Western Europe,
giving values for the Spalding transfer number B, which European Overview, 1992 Data, Vol. 1; Assoc. for Plastics Manuf. Eu-
correspond with predictions by Wey and Chang. 18 rope (APME), 1994.
2. Mosbacher, R. Kunststoffe-Plastics, 8/89 1989,16-21.
Only PVC produced a carbonaceous residue dur- 3. Kaminsky, W. ACS Symposium Series, 513 (Emerging Technol. Plast.),
ing devolatilization, which supports the findings of 1992, 60-72.
4. Darivakis, S.D.; Howard, J.B.; Peters, W.A. Comb. Sci. and Tech. 1990,
Kaminsky 3 and Cullis and Hirschler. 8 PS, however, 74, 267-281.
5. Scott, D.S.; Czernik, S.R.; Radlein D. Energy from Biomass and Wastes
yielded no solid residue after pyrolysis, although a vast 1993, 14, 1009-1017.
amount of soot was released. The PVC particles showed 6. Redepenning, K.-H., Von VGB Kraftwerkstechnik 1994, 74, 688-696.
7. “Thermal Recycling of Plastics”; Final report to the American Plastics
a faster ignition than a similar sized particle of another Council; Energy & Environmental Research Center; University of
North Dakota, 1994.
plastic. Plastic particle shape had no effect that can be 8. Cullis, C.F.; Hirschler, M.M. The Combustion of Organic Polymers;
explained by a re-formation into a molten droplet dur- Clarendron Press: Oxford 1981; pp. 117, 66-69
9. Elomaa M., Ph.D. Thesis, University of Helsinki, Helsinki, Finland, 1991.
ing heating and early stages of the pyrolysis as was 10. Panagiotou, T.; Levendis, Y. Combustion and Flame 1994, 99, 53-74.
seen in the video recordings. The carbonaceous resi- 11. McGhee, B.; Norton, F.; Snape, C.E.; Hall, P.J. Fuel 1995, 74, 28-31.
12. Cozzani, V.; Nicolella, C.; Petarca, L.; Rovatti, M.; Tognotti, L. Ind.
due from PVC (after atmospheric devolatilization) con- Eng. Chem. Res. 1995, 34, 2006-2020.
13. Cozzani V.; Petarca, L.; Tognotti, L. FUEL 1995, 74(6) 903-912.
tained less than 1% chlorine, 99% hydrocarbon. Dur- 14. Frankenhaeuser, M. In LIEKKI Combustion Research Program, Technical
ing atmospheric and pressurized pyrolysis and gasifi- Review 1988-1992; Hupa, M.; Matinlinna, J., Eds.; Turku/Åbo, Finland,
1993, pp. 901-912.
cation tests with PVC and coal, peat, and wood, a sig- 15. Frankenhaeuser, M.; Manninen, H.; Peltola, K.; Hiltunen, M. In LIEKKI
2 Annual Report 1995; Hupa, M.; Matinlinna, J., Eds.; Turku/Åbo, Fin-
nificant effect of pressure on char yield after the py- land, 1995 pp. 1225-1242.
rolysis and gasification rate was found. The gasifica- 16. Kojo, I.V. Presented at the 10th IFRF Members Conference,
Noordwijkerhout, The Netherlands, May 1992.
tion rate of PVC “char” appears to be of the same or- 17. Czernik, S.; Koeberle, P.G.; Jollez, P.; Bilodeau, J.F.; Chornet, E. In Ad-
der as that of char from coal. Peat-char and wood-char vances in Thermochemical Biomass Conversion; Bridgwater, A.V., Ed.;
Blackie Academic & Professional: Glasgow, U.K., 1994; Vol. 1.
were gasified 1 order of magnitude faster. 18. Wey, M.Y.; Chang; C.L. Polymer Degradation and Stability 1995, 48, 25-33.
19. Kuo, K.K. Principles of Combustion; John Wiley and Sons: New York,
While coal and peat will be converted roughly for 1986, Chapter 6.
half by volatiles combustion and for half by char com- 20. Knümann, R.; Bockhorn, H. Comb. Sci. and Technol. 1994, 101, 285-299.
21. Polymer Handbook, 2nd ed.; Branderup, J.; Immergut, E.H., Eds.; Wiley
bustion, plastics-derived fuels will be burnt almost com- and Sons: New York, 1975; pp. II-467-478.
22. Karlsson, M. M.Sc. Thesis, Åbo Akademi University, Turku/Åbo, Fin-
pletely in the gas phase. Co-firing with plastics-derived land, Jan. 1995.
fuels will significantly increase the amount of volatiles
in the freeboard of a bubbling fluidized bed furnace. This
puts an upper limit to how much PDF or RDF can be fed
to a furnace, depending on furnace size, the mixing in
the furnace, and the location of heat-exchanging sur-
face. For new furnace designs, however, no technical
upper limit exists. Co-firing with low-volatile coals will
allow for higher mass ratios of PDF/RDF. PVC has the About the Authors
Ron Zevenhoven (corresponding author; fax 358-2-2654780)
disadvantage, as compared to the other plastics, of the
and Mikko Hupa are associate professor and professor, re-
heavy HCl emissions and a lower calorific value. PVC
spectively, at the Combustion Chemistry Research Group of
produces a “char” that could catalyze the reduction of the Department of Chemical Engineering, Åbo Akademi Uni-
other gaseous emissions. versity, Lemminkäisenkatu 14-18 B, FIN-20520 Turku, Fin-
land. Magnus Karlsson is process engineer at ABB Carbon
ACKNOWLEDGMENTS Ab, SE-61282 Finspång, Sweden. Martin Frankenhaeuser
works as project manager for energy recovery, plastics, and
This text was produced from the M.Sc. (chemical en-
environment at Borealis Polymers Oy, P.O. Box 330, FIN-
gineering) thesis work of Magnus Karlsson 22 at Åbo 06101 Porvoo, Finland.
Akademi University. The authors want to thank Bo-
realis Polymers Oy, Neste Oy Chemicals, and VAPO Oy,

870 Journal of the Air & Waste Management Association Volume 47 August 1997

Vous aimerez peut-être aussi