Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Int. J. Environmental Technology and Management, Vol. 7, Nos.

1/2, 2007 21

Automobile Shredder Residue (ASR) destruction


in a plasma gasification reactor

Marco G. Tellini* and Paolo Céntola


Politecnico di Milano,
Dipartimento Chimica, Materiali ed Ingegneria Chimica,
P.za Leonardo da Vinci, 32, 20133 Milano, Italy
E-mail: marco.tellini@polimi.it E-mail: paolo.centola@polimi.it
*Corresponding author

James A. Batdorf and William J. Quapp


Integrated Environmental Technologies, LLC
1935 Butler Loop, Richland, WA 99352, USA
E-mail: jbatdorf@inentec.com E-mail: wjquapp@inentec.com

Abstract: Test results on Automobile Shredder Residue (ASR), or car fluff,


demonstrated destruction efficiency and safe conversion to synthesis gas and a
glass residue, in a plasma gasification system. The synthesis gas consists
primarily of hydrogen and carbon monoxide in the range between 20 and
22 vol-% respectively, or 45 and 55 vol-% dry basis, when corrected for
nitrogen. In dry reforming operation, carbon dioxide conversion approached
90%. The system is designed to work with oxygen in autothermal conditions,
reducing thus the electric power requirement for the plasma reactor.
The vitrified residue leach rate makes the product suitable for construction
works.

Keywords: Auto Shredder Residue; ASR; car fluff; plastic scrap; plasma
gasification; vitrification; leaching; syngas; dry reforming.

Reference to this paper should be made as follows: Tellini, M.G., Céntola, P.,
Batdorf, J.A. and Quapp, W.J. (2007) ‘Automobile Shredder Residue (ASR)
destruction in a plasma gasification reactor’, Int. J. Environmental Technology
and Management, Vol. 7, Nos. 1/2, pp.21–38.

Biographical notes: Marco G. Tellini is presently engaged in research on high


temperature waste treatments and holds lectures in Environmental Chemical
Engineering and Petroleum Process Technologies at the Politecnico di Milano.
He graduated and received his PhD in Chemical Engineering from the
same University and has an MBA in Marketing and Finance from
Fairleigh Dickinson University, NJ, USA. He worked in various international
engineering and chemical companies as manager and as a licensed Professional
Engineer.

Paolo Céntola is a Professor of Environmental Chemical Engineering at the


Politecnico di Milano. He graduated in Chemical Engineering from the same
University and conducted several consultations for chemical industries, with
extensive research and development in the area of organic syntheses and
environmental engineering. He initiated and directs the Laboratory for
Instrumental Olfactometry at the Politecnico di Milano.

Copyright © 2007 Inderscience Enterprises Ltd.


22 M.G. Tellini, P. Céntola, J.A. Batdorf and W.J. Quapp

James A. Batdorf is a Director of Research and Development for Integrated


Environmental Technologies (IET), responsible for plasma processing research,
by-product utilisation and energy recovery. He holds BS, MS and PhD Degrees
in Chemical Engineering from the University of Idaho. Prior to working at IET,
he conducted basic science research in plasma technology at the Idaho National
Engineering Laboratory (INEL) and Science Applications International
Corporation, followed by process development, control and modelling in the
synthesis of materials and waste processing.

William J. Quapp is a Director of Project Development at Integrated


Environmental Technologies, responsible for waste processing economics,
technical analyses and preparation of environmental permits. He has a BS in
Science from San Diego State University, a MS in Mechanical Engineering
from the San Jose State University and has 40 years of engineering and project
management experience with prime companies that include General Electric,
Westinghouse, Lockheed, and EG&G. He has over 15 years of experience in
nuclear and hazardous waste management with over 70 publications in the
fields of waste management, system engineering, depleted uranium technology,
nuclear engineering, and reactor safety.

1 Introduction

Automobile recycling involves a shredding operation that allows separation of most


metals from non-metallic residual materials. The residue has come to be known as
‘Auto Shredder Residue’ (ASR) or car fluff. Current waste management practice in most
countries involves ASR simple landfills, but this disposal practice represents a growing
environmental concern due to the presence of toxic materials and other compounds that
may eventually contaminate water supplies. ASR has a substantial quantity of organics
with a carbon content that can exceed 50% on a weight basis.
This paper presents the results of ASR processing in a plasma reactor to produce
synthesis gas (syngas). ASR was expected to be easily processed with plasma gasification
technology since the high temperature and chemically reducing atmosphere ensures the
total destruction of contaminants while converting organic materials to syngas. Inorganic
materials are converted into the glass or metal phase, which is suitable for reuse as road
and construction fill material or is otherwise safely disposable, since the leach rate of
toxic compounds from the glass proves to be below environmental regulations. Based on
results from previous testing, the reducing chemical environment kept in this
gasification system produces minimal dioxins and air pollutants, at nearly non-existent
NOx formation, which would rather occur in an oxidising atmosphere plasma.
These tests were conducted to evaluate the processing characteristics of ASR in a
series of scoping tests and compare steam vs. carbon dioxide as the oxygen donor
for gasification. The use of carbon dioxide offers the option to consume CO2 emissions
from another source (any combustion) and reduce the net greenhouse climate impact
from that source.
Automobile Shredder Residue (ASR) destruction 23

2 Fluff characteristics

Although ASR is also derived from domestic appliances like refrigerators, the greatest
quantity comes from shredded automobiles already stripped of most of their reusable
metals. Vehicle demolition and recycling generates ASR with variable characteristics: the
changes in auto making affect ASR relative quantities and composition as it appears from
the substantial differences in published data.
ASR contains plastics, rubber, paper, wood, fabrics, sand, dirt, glass, teflon,
ferrous and non-ferrous metal pieces and, possibly, automotive fluids and refrigerants.
Subclasses of plastic-base materials and foams are polyurethanes, PP, PE, ABS, nylons,
PVC, paints and acrylic varnishes. Sheet moulded composites and resins, polyesters and
so called ‘engineered plastics’ are also found, since more and more structural plastics are
used in the auto industry. The main constituents and chemical composition of ASR are
given in Table 1, reporting data from the following referenced sources: DBJ (2003),
WWF (2000) and ANPA (2002). The variability of data is often attributed to the age of
vehicles. With the technology and materials of modern cars, for instance, PVC is
moderately used, while in old vehicles PVC could represent 10% of the fluff. The exact
opposite trend is reported for polyethylene and polypropylene.

Table 1 ASR average constituents and composition (wt-%)

Data source Preliminary estimate DBJ ANPA WWF Analysed sample


LHV (kJ/kg) – – – – 25640
Specific weight (kg/l) – – – – 0.60
Glass and inert 14 16.1 0.14 44 ÷ 25 –
Wood 6 – 0.50 – –
Paper 1 – 0.84 – –
Fabric 18 17.1 24.90 – –
Plastics 40 – 13.86 38 ÷ 60 –
Resin 10 39.3 – – –
Rubber 11 8.7 22 15 ÷ 10 –
Of which
Carbon 46.89 – 47 ÷ 49 40 54.40
Hydrogen 6.25 – 6 ÷ 6.15 5 6.48
Oxygen 8.72 – – – –
Nitrogen 1.96 – 1.6 ÷ 2.4 1 2.01
Chlorine 3.61 – 2.87 3.5 1.97
Fluorine – – – – 0.13
Sulphur 0.43 – 0.27 ÷ 0.33 0.95 3.47
Inert 13.68 – – – –
Water 10.46 – – – –
24 M.G. Tellini, P. Céntola, J.A. Batdorf and W.J. Quapp

Table 1 ASR average constituents and composition (wt-%) (continued)

Data source Preliminary estimate DBJ ANPA WWF Analysed sample


Aluminium – 5.1 – – 0.42
Arsenic – – 0.0006 – 0.00038
Cadmium – – 0.0017 0.0008 0.0012
Calcium – – – – 1.61
Chromium – – 0.00004 0.00005 0.021
Copper 2 4.4 0.62 ÷ 1.9 0.56 ÷ 3 0.36
Iron 3 8 2.5 1.5 ÷ 3 3.30
Lead 2 0.2 0.178 0.28 0.14
Magnesium – – – – 0.20
Mercury – – 0.0002 – <0.00005
Selenium – – 0.0001 – <0.00005
Silicon – – – – 11.25
Zinc 1 1.1 0.35 0.7 ÷ 2 0.33

Annual quantities of ASR were reported in the range of 4 million tons in 1999, about
22% of the weight of ferrous scrap recovered from the estimated 18 million recycled
automobiles in the USA (Daniels, 1999). ASR quantities in Europe are estimated to
exceed 2 million metric tons per year and about 800000 tons per year is the amount to be
disposed of in Japan (DBJ, 2003; Daniels, 1999). These large quantities present various
problems of disposal management. The chlorine content of the ASR generally prohibits
incineration due to concerns regarding dioxin production and emissions. Landfill is
currently the most widespread method of disposal and the selective recovery of valuable
materials is practically impossible due to the heterogeneous constituents.
ASR restrictions in Europe require progressive diversion of the ASR from
landfill disposal, with the goal of reducing the landfill fraction to 5 wt-% by 2015
(EC Directive No. 53, 2000). In addition to environmentally driven issues, the average
landfill costs, in the range of $160 per ton in Germany and $80 per ton in Italy, just to
provide two current instances, are increasing. Finding a landfill destination will also
become harder and these costs are a definite potential offset for any ASR industrial
conversion that will also need to be determined in view of applying a carbon tax on
CO2 emissions (Tellini et al., 2004).

3 Thermochemical evaluation

A series of pre-test calculations were performed to evaluate syngas formation at high


temperature, starting from model constituents reacting with oxygen, water and carbon
dioxide. The reactions considered are shown in Table 2, where ‘i’ is an index for different
reactions of constituent cases.
The model’s constituents (organics) were chosen from among a variety of reacting
materials commonly reported in the ASR, namely, polystyrene, polyethylene, PVC and
rubber and were elected to be model constituents. Materials that do not react in the
plasma reactor gasifier to produce syngas, or materials that exist as glass, metals, silica
Automobile Shredder Residue (ASR) destruction 25

and inert, were not modelled. The constituents’ relative mixture ratio was then adjusted to
approach the elemental composition actually analysed in the shredded, mixed,
homogenised ASR sample. Simulated compositions are shown in Table 3.

Table 2 Typical model reactions for ∆Hi (case stoichiometry not shown)

Set 1 Index for the various cases


i=1 CxHy + zS + tiO2 = CO + zH2S + H2
i = 2, 3 CxHy + zS + tiO2 + riH2O = CO + zH2S + H2
i=4 CxHy + zS + tiO2 + siCO2 = CO + zH2S + H2
i = 5, 6, 7 CxHy + zS + tiO2 + siCO2 + riH2O = CO + zH2S + H2

Table 3 Elemental composition used for the thermochemical simulation (wt-%)

Percentage
Mockup species and ASR C H Cl S N factor
PVC 38.00 5.90 56.10 – – 5
PE 85.00 15.00 – – – 9
Isoprene rubber 78.80 10.70 – 10.50 – 30
Polystyrene 92.20 7.80 – – – 56
100
Analysed ASR sample 54.40 6.48 1.97 3.40* 2.01 –
Mix @ 25% for each type** 51.45 6.90 9.82 1.84 – –
Mix @ % factor** 59.38 6.46 1.96 2.21 – –
*Determined as total sulphur.
**Factored with a 30% inert material content.

The calculations were meant to provide preliminary operating ranges and compositions
for the testing. Model reactions were considered at 1300°C and 1100°C to close the
thermal balance for CO and H2 formation in respect of estimated thermal losses and
external energy inputs. The HSC program that utilises an internal data bank taken
from JANAF Thermochemical Tables was used for the thermochemical calculations
(Roine, 1999; Chase et al., 1985). Residual hydrocarbon, carbon soot formation and
simultaneous Boudouard equilibria were also considered for dry reforming with CO2
addition. These topics will be resumed in detail in a later paper.
Each set of model reactions was studied to allow an intrinsic heat and material
balance and the overall result were then factored to match the ultimate feed composition
of the ASR sample to the reactor. Given a fixed electric power input, the quantity of O2
starved combustion was calculated for various feed inputs of H2O and CO2. It is
reasonable to assume that all gas-phase molecules behave ideally, pressure is atmospheric
or slightly negative, kinetics are assumed to be very rapid since the temperatures are well
above the equilibrium temperatures that minimise Gibbs free energy. On this basis, molar
balances and gas compositions were calculated.
The mass feed ratios of O/C, H2O/C and CO2/C were varied between 0.2 and 2.0, so
as to collect experimental data for a combination of oxygen, water and carbon dioxide
feed, at practically constant organic feed rate, with quasi constant DC input of 14 kW to
the plasma electrodes and AC power input lower than 10 kW. We can well assume that
26 M.G. Tellini, P. Céntola, J.A. Batdorf and W.J. Quapp

the AC power was spent as the Joule-effect for heating and maintaining the molten
glass bath temperature, whereas heat losses consumed part of the DC power supply.
The chemical energy from the reactions and partial oxidation (POx) of organics are
therefore tied to pure oxygen feed, reacted CO2 and H2O.
Figure 1 represents the starting base case, calculated for the reactions (1–7) of each
constituent. The identified boxes for the pro-rated mix composition have the following
meaning: A (specific heating value HV of syngas), B (specific heat of reaction), and C
(O/C wt-ratio). Specific values are referred to 1 gram of substance since we used mockup
species (moles of simulated compound may not correspond to the moles of the
assimilated species) and it is useful to refer to the unitary weight of carbon feed.

Figure 1 Operating parameters calculated for the test (simulated mix basis)

While the energy of the products is comprehensively characterised by the specific heating
value of the syngas (3.2–4.9 kcal/gram-syngas, i.e., 13.4–20.5 kJ/g), the heats of reaction
for the mix depend on the constituents and are kept low (negative to +0.2 kcal/g,
≈ 0.84 kJ/g) near the autothermal level, by consuming substance and burning O2 to CO.
By looking at the O/C ratios, we estimated an appreciable consumption of oxygen
Automobile Shredder Residue (ASR) destruction 27

(O/C = 0.1–0.2 or 10–20 wt-% of the organic carbon source). This consumption relates to
the partial oxidation of the substance since we attribute all H2O and CO2 transformation
to CO and H2. Given this amount of partial oxidation we can observe that a smaller
amount of POx is required when no CO2 is fed. At combined or apportioned feed of CO2
and H2O, POx is always needed to sustain the endothermic reactions. For the cases of
Figure 1 the CO grams were between 226 and 323, H2 was between 11 and 23 grams,
starting from the organic carbon feed mix of 97 grams. The subsequent step of the
calculation was done to introduce DC power, compensate for heat losses and sensible
heat, up to 1300°C, and consider the input derived from the consumption of graphite
electrodes.
Table 4 compares the initial base values with the revised calculation: the O/C ratio
creeps to 0.8, 4–5 times higher, having introduced more CO2 (CO2/C about three times
more) and more H2O (H2O/C about 5–6 times more). The additional oxygen has the net
result of increasing POx; in fact, the mass quantity of CO increased to 350 ÷ 450 grams,
i.e., more when CO2 is fed to the system, since dry reforming reactions are more energy
intensive than steam reforming.

Table 4 Calculated variations of operating parameters to include losses

Base case vs. simulation or reactions including kW, losses and sensible heat (HSC)
Increase Increase
Base HSC n-fold Base HSC n-fold
P. Ox. 1 0.381 Same – P. Ox. 226.324 Same –
2 0.211 0.806 3.8 grams 226.324 286.344 1.3
O/C 3 0.091 0.790 8.7 CO 226.324 448.402 2.0
4 0.153 0.886 5.8 322.400 400.217 1.2
5 0.125 0.826 6.6 284.726 405.256 1.4
6 0.132 0.857 6.5 301.672 414.287 1.4
7 0.147 0.854 5.8 304.193 381.488 1.3
P. Ox. 1 0.000 0.000 – P. Ox. 10.996 Same –
2 0.095 0.593 6.2 grams 18.374 22.103 1.2
H2O/C 3 0.163 0.612 3.7 H2 22.768 34.895 1.5
4 0.000 0.000 10.996 15.019 1.4
5 0.059 0.312 5.3 15.280 22.285 1.5
6 0.025 0.160 6.3 13.556 19.967 1.5
7 0.016 0.122 7.5 12.595 16.907 1.3
P. Ox 1 0.000 0.000 – P. Ox. 3.579 Same –
2 0.000 0.000 – kcal/g 4.370 4.272 1.0
CO2/C 3 0.000 0.000 – HV-syn 4.858 4.239 0.9
4 0.342 0.230 3.6 3.244 3.302 1.0
5 0.226 0.631 2.8 3.794 3.738 1.0
6 0.305 0.918 3.0 3.497 3.519 1.0
7 0.307 1.019 3.3 3.420 3.456 1.0

The net DC power input, estimated to be 6 kW for a 6 kg/h fluff feed, equals
3.6 kJ/g-substance and is smaller by a factor of 2.4 when compared to 8.92 kJ/g, the POx
heat of reaction generation. If a 0.2 O/C factor is applied as POx and we downrate it for
28 M.G. Tellini, P. Céntola, J.A. Batdorf and W.J. Quapp

the 83% carbon content of the mix that was simulated, we conclude that our modelling is
for an operation in which the POx energy contribution is (8.92)(0.166) : 3.6 = 0.41 or say
40% of the pilot plant’s useful DC input.
The H2/CO molar ratio for the various reforming cases was calculated to remain
between 0.4 and 0.7.

4 Test system description

The PEMTM system is patented for technological features like the dual heating system,
the bottom glass and metal drains (US Patent, 1997) and it has been previously reported
for very low emissions while treating hazardous waste, electronic scrap and medical
waste. The description here is limited to key process features illustrated in Figure 2 and a
simplified process flow diagram that is shown in Figure 3. The reactor accomplishes two
operations simultaneously, gasification and vitrification. The system is operated with a
molten glass pool at the bottom of the process vessel. The glass pool is heated by AC
joule heating to a temperature that exceeds 1200°C, while the zone just above the bath, in
the vicinity of the arc electrodes, can easily reach much higher temperatures, depending
on the power input. The system is generally operated with a plenum exit temperature of
1200°C. The DC plasma arc is produced via graphite electrodes, which arc to the molten
glass surface and produce intense heat for the endothermic gasification reactions.
The organic materials are reacted at and above the molten glass surface with steam and a
small amount of oxygen. Plasma heating is important for the application of fluff
gasification. Hundreds of cracking or pyrolysis reactions were studied and modelled at
Politecnico di Milano (Ranzi et al., 2001). Radicals formation and propagation of
precursors into tars and polyaromatic hydrocarbons can actually occur in multiple and
complex mechanisms. Carbon is converted to CO with minimal formation of carbon
black. From the chemical point of view, catalysts are often utilised for cracking and
visbreaking processes. The various, solid and heterogeneous wastes can hardly utilise
catalysts, whereas the plasma environment can easily reach thermodynamic equilibrium
conditions and overcome kinetic hindrance for the thermal gasification to syngas.
Without catalysts or the need to operate the plasma reactor at extreme temperatures, the
advantage is also to avoid technological severity and sophisticated materials for the
construction of a commercial plant.

Figure 2 Conceptual view of the plasma enhanced melterTM


Automobile Shredder Residue (ASR) destruction 29

Figure 3 Simplified process flow diagram for the test system

Inorganic materials are typically dissolved into the glass pool as an oxide. Non-oxidised
metals sink to the bottom of the vessel where they melt and alloy together to form a pool
of molten metal at the bottom of the glass pool. After cooling, the glass becomes a
very leach resistant matrix and has been shown to immobilise, effectively, toxic metals
such as lead chromium and cadmium. Iron and nickel are found in a metal phase beneath
the glass.
As regards product gas emissions, the system does not operate at conditions
normally associated with dioxin production (Tuppurainen, 2003). Extensive testing on the
destruction of PCBs using a commercial scale PEMTM system was conducted by
Kawasaki Heavy Industries in Japan in 2003 (Okita et al., 2004). All tests showed dioxin
emissions to the atmosphere well below regulatory standards in the USA, Japan, and the
EU. Chlorine reacts with hydrogen to form HCl which is easily scrubbed from the
syngas. Lastly, the syngas temperatures are well above the dioxin reformation window
until they exit the TRC where they are rapidly quenched to below 200°C in a fraction of a
second.
NOx compounds are not formed because of the reducing environment and very low
O2 partial pressure. Under these conditions, oxygen in the system reacts preferentially
with excess hydrogen and carbon monoxide rather than with nitrogen. A small quantity of
glass, carbon particles, and other materials will exit the plasma reactor with the syngas,
but are readily removed in the baghouse filter. Copper and other metals tested by feeding
electronic scrap remain primarily in the molten glass bath (Quapp and Lamar, 2003).
Tests dealing with mercury contaminated hospital wastes demonstrated that mercury
can be effectively captured and exhibit emission values below the detection limit
(Batdorf et al., 2005).

5 Testing procedure

The key objectives of the test were to demonstrate consistent, controllable, and
reliable feeding of the ASR material, its efficient conversion to syngas, evaluate waste
heat recovery, and demonstrate the leach characteristics of the discharged glass.
The hazardous air pollutants in the syngas were also important but they had already been
measured in previous tests on chlorinated organics and other waste streams. Due to the
expense of repeating such measurements, no confirmatory air pollutant sampling was
conducted during these screening tests.
The ‘as received’ heterogeneous materials, variable in size, quality and physical
properties, were up to 15 ÷ 20 cm wide and too large to feed to the pilot plant.
30 M.G. Tellini, P. Céntola, J.A. Batdorf and W.J. Quapp

Once shredded, an auger, belt or feed device can then be calibrated to obtain constant,
controllable, reproducible feedrate. After shredding and mixing, the ASR particles had an
apparent uniform size of about one-half centimetre and made a relatively homogeneous
mixture.
After bringing the reactor (PEM) and Thermal Residence Chamber (TRC) to above
900°C, fluff and oxygen were fed, and steam, oxygen and carbon dioxide were added as
main process variables, at constant fluff rate.
The solid feedrate and the DC and AC power were established and maintained
constant for the duration of the tests. H2O, CO2, O2 and N2 were manually recorded at
regular intervals. The syngas was continuously measured throughout the test to read the
vol-% of H2, CO, CO2, O2, N2 and CH4.
Utility gases were measured with rotameters calibrated prior to the beginning of the
experiments. Other process variables like temperature and pressure were measured at
various points throughout the system. Most of the process data were recorded by the
process computer operating the main control system.
The off gas volumetric composition was measured by a process gas analyser
(manufactured by NOVA) where the product gas sample is drawn through a series of
heated particulate filters and then dried over a semi-permeable membrane that allows
transport of the water in the sample to a dry nitrogen stream. The resulting clean and dry
syngas is analysed by different detectors. Simultaneous measurement of carbon
monoxide (CO), carbon dioxide (CO2), and methane (CH4) is accomplished with a high
stability infrared detector. Hydrogen (H2) is detected with a methane compensated
thermal conductivity cell. Oxygen (O2) is detected with a long-life electrochemical
‘fuel-cell’ sensor. Calibration gases were used at the start of each test to ensure
measurement accuracy. About 10 vol-% CO2 was allowed to remain in the syngas, which
was achieved by controlling the steam injection with the ultimate aim of measuring a
CO/(CO+CO2) ratio of about 0.7. The procedure is explained for establishing a degree of
water-shift-reaction and simultaneous hydrocarbon Boudouard equilibria to yield and
limit the carbon black formation to about 5% of the total feed rate. The testing feed rate
was held constant at 6–6.3 kg/h for all the selected tests except for the last few runs,
which were operated at twice the feedrate. Given a fixed ASR flowrate, the oxygen donor
reagents (H2O and CO2) were planned to vary for the tests as shown in Table 5.

Table 5 Average feed weight ratios and temperatures

Test no. H2O/C CO2/C PEM (°C) TRC (°C)


T51 – 0.20 1010 990
T52 – 0.43 1070 1050
T53 – 0.12 1050 1080
T54 0.09 – 1090 1085
T55 0.19 – 1080 1110
T56 0.41 – 1095 1130
T61 0.30 – 1080 1045
T62 0.82 – 1062 1065
T63 1.18 – 1080 1090
T64 1.71 – 1075 1110
T65 0.78 1.41 1075 1115
T66 – 0.95 1100 1135
T67 0.93 – 1150 1150
Automobile Shredder Residue (ASR) destruction 31

O2 feed was also varied during the tests so as to obtain near autothermal conditions at the
fixed DC power input of 14 kW, conditions evidenced by the average temperature
readings in the PEM and TRC sections of the pilot plant. Solid glass from the plasma
reactor and ash from the baghouse were removed at the end of each day’s testing, and
represent daily averages.

6 Material and energy balance

Before discussing syngas quality, a preliminary appreciation of quantities of material and


power involved in the process will facilitate the understanding of results. The major
energy and material flows are listed in Table 6, that refer to the stream tags of Figure 4
block diagram.

Table 6 Mass and energy inputs and outputs


32 M.G. Tellini, P. Céntola, J.A. Batdorf and W.J. Quapp

Figure 4 Material and energy balance block diagram

7 Discussion of results

7.1 Power requirements and oxygen sources


Given the DC power feed limitations of the experimental unit, the first operational
requirement was to supply sufficient oxygen to establish a quasi auto-thermal process and
provide the energy for the endothermic gasification reactions. Oxygen feed ratios were
varied with the main criteria to sustain the ASR conversion and syngas generation,
depending on the water or carbon dioxide feed to the reactor. The power consumption is
also affected by thermal losses but since there is functional equivalence between electric
power and oxygen supply, the choice becomes a matter of relative cost. The performance
of the process and the attained temperatures are interchangeable irrespective of the power
source.
Due to power optimisation requirements, oxygen is always used to make the process
quasi autothermal and the cracking of the organic source with O2 is difficult to control:
starving O2 causes elementary carbon deposition and high O2 implies fast kinetics, CO2
formation and overheating. In such a context, the high temperature thermodynamic
comparison of ternary equilibrium systems C-CO2-O2 and C-H2O-O2, where C stands for
any organic source, is less energy intensive for steam reforming, yet it is feasible and
controllable when CO2 is used alone or in combination with water.
When choosing from different sources of ‘spent or reacted’ oxygen, dry reforming
exhibits higher ∆HR than steam reforming. For methane reforming, for instance, both
reactions are endothermic, ∆HR = 205 kJ/mol and ∆HR = 247 kJ/mol for H2O and CO2
respectively. Historically, this thermochemical reason has limited the applications of dry
reforming, which may now become matter of concern if greenhouse reduction or CO2
emissions gain increasing interest. The use of water leads also to three substantial
advantages. The first advantage is that hydrogen is supplied, which is important when
low hydrogen feedstocks are used and hydrogen is the ultimate product. The second
advantage is that a water excess is introduced; this acts as a beneficial operating tool to
control and stabilise the temperature due to the water thermal capacity and through the
parallel reactions of water-gas shift (moderately exothermic) and steam reforming
(endothermic). The third advantage is that the formation of carbon soot is limited, or
easier to control: the CO concentration of the shift reaction limits the parallel Boudouard
carbon equilibrium.
Automobile Shredder Residue (ASR) destruction 33

7.2 Effect of oxidiser


Experimental water to carbon feed ratio or CO2 to carbon feed ratio ranged from
0.2 to 1.8 and the amount of CO and H2 is shown in Figure 5.

Figure 5 Syngas composition for each trial as a function of oxygen donor to carbon ratio

The diameter of the bubbles is proportional to the number of repetitive measurements at


indicated feed ratios, (i.e., 1861 CO measurements for trial no. 7). The maximisation of
the H2 concentration and bubble frequency was obtained for the higher steam to carbon
ratios whereas excessive steam injection did not significantly improve the hydrogen
production though it increased the power input requirement. Providing excess water
beyond the quantity needed for gasification results in no useful output. The mid points
(3–6) have dual feed of H2O and CO2. As the water increases, with lower CO2 feed,
H2 increases and the CO production diminishes.

7.3 Repeatability and process variability


Repeatability for the production of CO and H2 in the selected trials is shown in Figure 6.
Measurements are averaged over thousands of process instruments readings and the
onion ring representation simplifies the reading of bulky data tables.

Figure 6 Repeatability in syngas generation


34 M.G. Tellini, P. Céntola, J.A. Batdorf and W.J. Quapp

While the syngas composition for each trial is represented by the individual skin layer of
the ring diagram, the average result for all trials is shown by the outer exploded ring.
The trials were characterised by stable run-lengths of about 20 minutes, and the operating
variables (O2, H2O, CO2) were ratioed to the carbon feed. Approximately 10 minutes
were allowed between operating conditions changes to stabilise the system at the new
operating points. The CO, H2, CO2 values in the syngas were manually taken every
5 minutes whereas the process instrumentation recorded syngas composition every
2 seconds. The October 6 trials main data are summarised in Table 7.

Table 7 Data averages for tests runs on October 6

Exiting gas (vol-%) Wt-ratio Feed wt-ratios


Trials CO CO2 CH4 H2 CO/(CO + CO2) H2O/C CO2/C O2/C
T61 18.2 10.2 0.4 16.2 0.6 0.3 – 1.7
T62 19.0 9.7 0.7 19.7 0.7 0.8 – 1.7
T63 17.9 11.0 0.5 19.7 0.6 1.2 – 1.8
T64 17.5 11.5 0.4 21.1 0.6 1.7 – 1.8
T65 22.4 13.4 0.4 15.4 0.6 0.8 1.4 1.7
T66 19.6 13.8 – 9.3 0.6 – 0.9 1.7
T67 24.6 7.6 – 19.1 0.8 0.9 – 2.1

Process computer recorded data from one specific test are plotted in Figure 7.
Fluctuations can be explained with the variable composition of the heterogeneous ASR
and by the irregularities in feeding the solid with an auger. Results would be smoothened
out for a commercial plant where the larger solid feedrate is likely to provide a time
averaged homogeneous feed composition. To better see the longer term behaviour that is
masked by short term variations, the gas measurements in Figure 7 are also superimposed
with a 6th degree polynomial interpolation. It then becomes quite evident that the H2
production (blue line, lower % level) develops homothetically with the CO production
(red line, higher % level) syngas peaks. The CO2 (green line) that is present in the syngas
is minimised at the same time that the CO and H2 peaks are maximised. The same
diagrams were developed for all tests and are not exhibited here for brevity: higher H2
syngas concentration (>18 vol-%) is exhibited for steam reforming, while H2 and CO
syngas concentrations (>16 and 22 vol-% respectively) are shown for combined steam
and dry reforming. The optimum split of H2O and CO2 feed will actually require
optimisation in relation to the ultimate quality of the fluff and for the electric power input
available to the plasma reactor.

7.4 Baghouse ash composition and glass leaching test


Baghouse ash was collected to determine the mineral characteristics that could affect the
design of a waste heat boiler. The mineral composition is presented in Table 8. All metals
are expected to be present as oxides. The baghouse ash can be recycled back to the
plasma reactor, so most of the ash will be incorporated into the glass bath and discharged
as molten glass. Recycling will also minimise the carbon black as it will mostly react to
CO. During the test programme, the average composition of the ash was also affected by
the inorganic coating agent, dolomite, which was used to pre-coat the bag filter.
Automobile Shredder Residue (ASR) destruction 35

Figure 7 Test data from October 6, 2004, Run 5

Table 8 Baghouse ash composition (wt-%)

LHV 946 kcal/kg = 3960 kJ/kg


Spec. weight (kg/l) 0.3
Carbon 12.90
Hydrogen 0.22
Oxygen N. a.
Nitrogen Nil
Chlorine 8.20
Fluorine 0.13
Sulphur (volatile) 0.65
Sulphur (as total) 3.47
Aluminium 0.19
Arsenic 0.00225
Cadmium 0.0061
Calcium 0.76
Chromium 0.0064
Copper 0.36
Iron 0.337
Lead 1.066
Magnesium 0.304
Mercury 0.00074
Phosphorus 0.035
Potassium 1.49
Selenium 0.0029
Silicon 11.92
Sodium 17.15
Zinc 2.51
36 M.G. Tellini, P. Céntola, J.A. Batdorf and W.J. Quapp

The discharged glass compares to glass removed in similar test trials. Leaching of the
samples was performed according to the UNI 10802-2002 method, a water leaching
test that conforms to the European norms set for disposable or safe reusable solids.
The comparison of the sample analysis and the permissible regulatory values are given in
Table 9.

Table 9 USEPA TCLP and UNI 10802_2002 metal concentration limits compared
to the concentration found in the analysed leachate from residual glass

US/EPA TCLP UNI 10802_2002 Analysed


Item Metal (including compounds) Test limit (ppm) Test limit (mg/l) leachate (mg/l)
1 pH – – 9.83
2 Tot. dis. solids – 400 100
3 D. org. carbon – 50 <1
4 Phenol index Not applicable 0.1 <0.01
5 Arsenic (As) 5.0 0.05 <0.001
6 Barium (Ba) 100.0 2 0.05
7 Boron (B) Not applicable Not applicable –
8 Cadmium (Cd) 1.0 0.004 <0.001
9 Chromium (Cr) 5.0 0.05 0.001
10 Copper (Cu) Not applicable 0.2 0.003
11 Lead (Pb) 5.0 0.05 0.002
12 Mercury (Hg) 0.2 0.001 <0.001
13 Molybdenum (Mo) Not applicable 0.05 <0.001
14 Nickel (Ni) Not applicable 0.04 0.001
15 Antimonium (Sb) Not applicable 0.006 <0.001
16 Selenium (Se) 1.0 0.01 <0.001
17 Silver (Ag) 5.0 Not applicable –
18 Thallium (Tl) Not applicable Not applicable –
19 Zinc (Zn) Not applicable 0.4 0.01
20 Chlorides (Cl) Not applicable 80 3.4
21 Fluorides (F) Not applicable 1 <0.1
22 Sulphates (SO4) Not applicable 100 <1

8 Conclusions

This paper reported the results from a series of 13 tests that continuously treated ASR in a
plasma gasification reactor with different oxidiser matrices to produce syngas. The tests
demonstrated the effectiveness of the chemically reducing environment of the plasma
reactor to destroy ASR and to produce high quality syngas.
The baghouse ash from the tests was analysed for designing a waste heat boiler to be
placed between the plasma reactor and the quench/flue gas treatment.
Automobile Shredder Residue (ASR) destruction 37

The clean syngas can be used to produce electric power in a gas turbine or
internal combustion engine generator (genset). Alternatively, the syngas can be used for
downstream production of high purity hydrogen, or fuel cells, or CO chemical synthesis.
ASR is a voluminous, leachable, hard to treat waste. Landfills, at least in Europe, are
not going to remain its preferred means of disposal, since regulations are evolving to
prohibit it. There is a series of advantages associated with the gasification treatment of
the ASR, which may well be considered and tailored to the specific reality of the country
where a plant is to be built. Plasma processing will produce two major revenue streams:
car fluff tipping fee and power or syngas sales destined for chemical synthesis.
Other revenue streams may also come from district heating in selected communities
where centralised hot water heating is used, from using the plant vitrified discharges as
construction fillers and by saving the carbon tax on CO2 emissions.

Acknowledgements

This work has been supported by Integrated Environmental Technologies, Richland, WA,
USA for joint investigations in 2004 with the Politecnico di Milano University.

References
ANPA (2002) ‘La caratterizzazione del ASR di frantumazione dei veicoli.’, Agenzia Nazionale per
la Protezione dell’Ambiente, Report 15/2002, Roma, February, pp.33–35, 45.
Batdorf, J.A., Lamar, D. and Quapp, W.J. (2005) ‘Demonstration of mercury capture efficiency
during medical waste processing in a PEMTM’, 24th International Conference on Incineration
and Thermal Treatment Technologies, Galverston, Texas, May 9–13.
Chase, M.W., Davies, C.A., Downey, J.R., Frurip, D.J. and McDonald, R.A. (1985)
JANAF Thermochemical Tables, American Institute of Physics, New York, NY, 3rd ed.,
ISBN 0883184737.
Daniels, E. (1999) Recovering the Plastic in Junked Cars, Argonne National Laboratory,
www.pprc.org/pprc/rpd/fedfund/doe/doe_oit/automobi.html.
Development Bank of Japan (DBJ) (2003) Prospects and Challenges for End-of-Life Vehicle
Recycling, Research Report No. 41, May, p.32.
European Community (EC) Directive No. 53 (2000) ‘2000/53/EC on End-of-Life Vehicles’,
Official Journal of the European Communities, October 21, Vol. L 269, pp.34–42.
Okita, M., Funaki, K., Quapp, W.J., Batdorf, J.A., Merrill, D.A. and Lamar, D.A. (2004)
‘Destruction of electrical equipment and PCB oil in a PEMTM’, IT3 2004 Conference
Proceedings, Phoenix, Arizona, May 10–14.
Quapp, W.J. and Lamar, D. (2003) ‘Waste gasification – test results from plasma destruction of
hazardous, electronic, and medical wastes’, IT3 Conference, Florida, May 12–16.
Ranzi, E., Dente, M., Bozzano, G., Goldaniga, and Faravelli, T. (2001) ‘Lumping procedures in
detailed kinetic modeling of gasification, pyrolysis, partial oxidation and combustion of
hydrocarbon mixtures’, Progress in Energy and Combustion Science, Vol. 27, pp.99–139.
Roine, A. (1999) HSC Chemistry for Windows Version 4.0, Outokumpu Research Oy, Pori,
Finland, June.
Tellini, M., Céntola, P., del Rosso, R. and Gronchi, P. (2004) ‘Hydrogen from waste’, Chemical
Engineering Transactions, AIDIC Hydrogen Conference Proceedings, ISBN 88-900775-3-0,
Pisa, Italy, May, Vol. 4, pp.143–150.
38 M.G. Tellini, P. Céntola, J.A. Batdorf and W.J. Quapp

Tuppurainen, K., Asikainen, A., Ruokojarvi, P. and Ruuskanen, J. (2003) ‘Perspectives on the
formation of polychlorinated dibenzo-p-dioxins and dibenzofurans during Municipal Solid
Waste (MSW) incineration and other combustion processes’, Acc. Chem. Res., Vol. 36, No. 9,
pp.653, 654.
US Patent (1997) Arc Plasma-Melter Electro Conversion System for Waste Treatment and
Resource Recovery, Issued 9/16/97, Licensed from BMI (patent followed by additional patents
and international patent counterparts), No. 5,666,891.
WWF (2000) Relazione Tecnica sui Rifiuti Derivanti Dalla Frantumazione Delle Carcasse di
Autoveicoli., WWF Italia, Piemonte-Val d’Aosta Regional Section, Torino, October, pp.3–8.

You might also like