Vous êtes sur la page 1sur 15

Progress in Computational Fluid Dynamics, Vol. 7, No.

1, 2007

25

Analysis of transport phenomena in a rotary-kiln


hazardous waste incinerator
Yongxiang Yang* and Marc J.A. Pijnenborg
Department of Materials Science and Engineering,
Delft University of Technology, Mekelweg 2,
2628 CD Delft, The Netherlands
E-mail: Y.Yang@TUDelft.NL
*Corresponding author

Markus A. Reuter
Faculty of Engineering, Department of Chemical and Biomolecular Engineering,
University of Melbourne, Melbourne, Australia
E-mail: mreuter@unimelb.edu.au

Joep Verwoerd
AVR, P.O. Box 59144, 3008 PC Rotterdam, The Netherlands
E-mail: Joep.Verwoerd@avr.nl
Abstract: Processing of hazardous waste in a rotary kiln incinerator is a complex process.
Since hazardous waste has often very complicated chemical compositions and physical forms, the
transport phenomena within the incinerator are not well understood, and the incineration process
expects large uncertainties in process chemistry and is difficult for emission control. For better
understanding of the incineration process, various transport phenomena taking place in the rotary
kiln were discussed and analysed in this paper. To get more quantitative understanding, process
simulation was conducted by using Computational Fluid-Dynamics (CFD) to characterise gas
and solid flow and mixing, temperature and species distribution in the incinerator. To include all
the waste streams in a single CFD model is a difficult task, and how to define the different
waste streams with different calorific values and chemical compositions is a challenge to the
CFD modelling. In this study, hazardous waste in various forms is firstly converted to a
hydrocarbon-based virtual fuel mixture based on an overall mass and energy balance.
The combustion of the simplified waste was then simulated with a combustion model.
The distribution of temperature and chemical species is broadly investigated under various
conditions. The predicted temperature distribution has been validated with available
measurement data from an operating rotary kiln waste incinerator, and reasonable agreement
between the predicted and measured data has been reached.
Keywords: hazardous waste incineration; rotary kiln; transport phenomena; Computational
Fluid-Dynamics (CFD); combustion model.
Reference to this paper should be made as follows: Yang, Y., Pijnenborg, M.J.A., Reuter, M.A.
and Verwoerd, J. (2007) Analysis of transport phenomena in a rotary-kiln hazardous waste
incinerator, Progress in Computational Fluid Dynamics, Vol. 7, No. 1, pp.2539.
Biographical notes: Yongxiang Yang received his BSc and MSc Degrees at Northeastern
University in Extractive Metallurgy (China, 1982 and 1988), and the degrees of Licentiate and
Dr. of Tech. at Helsinki University of Technology in Materials Processing (Finland, 1992 and
1996). He has worked as design engineer, researcher and lecturer in metallurgical engineering.
Since 1998 he has been working at Delft University Technology (The Netherlands) as an
Assistant Professor in Process Metallurgy. His main interests include metallurgical and materials
processing, transport phenomena, and applications of CFD in process simulation. He is a
co-author of the book Metrics of Material and Metal Ecology.
Marc J.A. Pijnenborg got his MSc Degree in Resource Engineering in 2003, Delft University of
Technology (TU Delft). He has made his MSc Thesis in modelling combustion and flow
behaviour in hazardous waste incineration process. He is now working outside TU Delft as
Process Engineer.
Markus A. Reuter: Chief Executive Technologist Ausmelt Ltd. Australia. Degrees: B.Eng.
(Chemical Engineering 1981), M.Eng. (1985), PhD (1991), D.Eng. (2006) all University of
Stellenbosch (South Africa) and Dr.Habil. (1995) Aachen University of Technology (Germany).

Copyright 2007 Inderscience Enterprises Ltd.

26

Y. Yang, M.J.A. Pijnenborg, M.A. Reuter and J. Verwoerd


Industry: Manager in the Measurement and Control Division, Mintek, Metallurgist at Anglo
American Corporation (both South Africa) and registered Professional Engineer (South Africa).
Academic: 19962005 Professor at the Delft University of Technology (Netherlands) and since
July 2005 Professor at University Melbourne (Australia). He has ca. 300 publications (journal
and conference proceedings) and has written a book Metrics of Material and Metal Ecology
Elsevier (Amsterdam).
Joep Verwoerd works as a Process Engineer at AVR Rozenburg, the largest MSW and hazardous
waste incineration company in the Netherlands. Before January 2005 he has worked for 32 years
in the plant of hazardous waste incineration at AVR-Chemie, part of the AVR Rozenburg.
After the kilns have stopped operation he is in charge of waste treatment related research. He has
written a book on hazardous waste processing, together with other authors, called gevaarlijk afval
(hazardous waste).

Introduction

Rotary kiln incinerators are widely used in the incineration


of various hazardous wastes such as liquid, sludge, and
solids in bulk or in packages. The benefits lie in the drastic
volume reduction and the substantial energy recovery.
The main objectives of the incineration are the complete
combustion of all the waste materials, and efficient recovery
of the thermal energy from the off-gases after the waste
combustion. Emission control of certain remaining species
in the off-gases such as CO and dioxins is an important
criterion for the operation. The complete destruction of
hazardous compounds depends very much on gas mixing
extent of air and various waste streams, the distribution of
gas temperature and residence time within the kiln
and the Secondary Combustion Chamber (SCC). Due to
large variations of waste types and difficulties in feed
characterisation of physical, chemical and thermal
properties, the complex transport and chemical processes
within the kiln system are not well understood, and thus
the incineration process often anticipates substantial but
unpredictable fluctuations of gas temperatures within the
system. The temperature fluctuations lead to uncertainties in
the process chemistry and difficulties in emission control.
The newly enforced directive from the European Union
(L332) (Directive 2000/76/EC, 2000) with increasingly
strict emission control requires a better understanding of the
incineration process and improved process control for low
emissions and less environmental impact. On one hand, the
operators have to comply with strict European guidelines
and legislation. On the other hand, the waste supply for
incineration is declining and the composition of the waste is
frequently fluctuating. The high calorific wastes can be used
in energy intensive industries to replace primary fuels,
and thus only the most difficult types of waste are delivered
for incineration in rotary kilns. In order to get better
understanding of the overall incineration process, research
has been carried out in close cooperation with AVR, the
largest waste processing company in the Netherlands. CFD
has been used to predict more insights of the gas flow, heat
transfer and waste combustion within the incineration rotary
kiln. CFD is a very convenient and flexible tool to simulate
the flow related transport phenomena for large scale

industrial processes. For rotary kiln waste incineration a


number of modelling attempts concerning flow and heat
transfer in incinerators have been reported (Jenkins and
Molar, 1980; Wolbach and Garman, 1984; Williams et al.,
1988; Leger et al., 1993; Khan et al., 1993; Chen and Lee,
1995; Jakway et al., 1996; Veranth et al., 1996, 1997;
Wardenier and van den Bulck, 1997; Ficarella and Laforgia,
2000), and CFD simulation has been used in a couple cases
for hazardous waste incineration (Leger et al., 1993; Khan
et al., 1993; Jakway et al., 1996; Veranth et al., 1996, 1997;
Wardenier and van den Bulck, 1997; Ficarella and Laforgia,
2000). However, in all the previous modelling work
combustion of different types of wastes in one process has
not been investigated, and a lot of questions need to be
further answered.
The early work from the current project focused more
on the thermal contribution to the temperature distribution,
by using a global combustion model of Spalding
(3-gas model) (Yang et al., 2001, 2002). All the waste was
averaged and modelled as a global fuel. In general, it proved
to be a useful approach, but it is not able to model the
chemical species distribution within the incinerator. In order
to model the distribution of major chemical species in the
waste combustion system, an extended global combustion
model (7-gas model) was applied. Since the diversity of the
waste types and large difference in heating value and
chemical compositions, the simulation program could not
handle directly the multi-fuel system. Therefore conversion
of various waste streams to a general fuel which can be used
in the CFD model becomes an important step. This has been
carried out through the definition of artificial fuel mixture
and waste stream optimisation. Then the subsequent
combustion modelling has been conducted by defining each
individual waste stream by using the artificial fuel mixture.
In this paper, the latest model developments and simulation
results with the 7-gas combustion model are presented.
New post-processing approach was illustrated to present
large data output from CFD simulation in a condensed and
engineering format. Temperature measurements at a number
of accessible locations of the operating incinerator at AVR
in the Netherlands was used to validate the combustion
model, and the further needs to validate the model from

Analysis of transport phenomena in a rotary-kiln hazardous waste incinerator


both thermal and chemical aspects of the waste combustion
process are emphasised.

Figure 2

27

Illustration of waste combustion system: the rotary kiln


and the Secondary Combustion Chamber (SCC)

The rotary kiln incinerator

The rotary kiln waste-incinerator operated before January


2005 at AVR was the target of simulation. The incineration
process consists of the waste combustion and Air Pollution
Control (APC) devices for off-gas cleaning system. Figure 1
illustrates the whole incineration system. The key to the
successful operation is the combustion system, though the
APC system is important to meet the environmental
regulations. It consists of a rotary kiln as the primary
combustion chamber and a vertical shaft as the SCC.
The rotary kiln is 4.2 m in diameter and 11.7 m in length,
mounted at a 12 angle and it rotates at a speed of
0.07 rpm. The SCC is 6.3 m in width, 5.5 m in depth, and
18 m in height. The thermal processing capacity ranges
from 30 MW to 40 MW, and the waste processing rate is
about 7 tons per hour. A wide range of hazardous wastes
with heating value from 5 MJ/kg to 30 MJ/kg is incinerated
in the system. The annual waste processing capacity is about
50,000 tons. Due to the difficult waste supply and high
maintenance cost, the 2 rotary kilns at AVR have ceased
operation since January 2005.
The waste enters the kiln and SCC in a variety of ways,
as is shown in Figure 2: main burner, burner for shredded
solid waste, sludge burner, SCC burner, and a couple of
lances for various liquid or sludge wastes. A load-chute is
used to supply combustion air and previously for supplying
containerised solid waste, and 14 air lances are installed in
SCC to supply additional combustion air. Air is known to
leak into the system through the front and end seals of the
kiln, viewports, and the ash sump. The off-gas leaves the
SCC to the Waste-Heat Boiler (WHB) for energy recovery
and preliminary dust separation. Molten slag formed in the
kiln flows down to the lower end of the rotary kiln and falls
into the ash sump located at the bottom of the SCC, which
also causes water vapour entering the SCC.
Figure 1

Rotary kiln incineration system for hazardous waste


processing

Transport phenomena

Hazardous waste processed in rotary kilns take various


physical forms: the liquid sludge, organic solutions, paint
and paste, containerised solid, shredded solid, and bulk
solids. The waste contains both organic and inorganic
compounds. The incineration of these various types of waste
is essentially oxidation reactions of organic waste with
oxygen and slag formation from the inorganic compounds.
Roughly 30% of waste by weight is converted as molten
slag which is discarded as the non-hazardous solid waste
after water quenching, and 70% of the waste is transformed
into combustion product of CO2 and water vapour. In order
to have a complete combustion, it requires a proper supply
of the oxidant (the combustion air), and a good and
preferably a turbulent mixing of the waste and air.
In addition, sufficiently high temperature is needed to get a
reasonable rate of the combustion reactions, which is
demanded by both thermodynamic and kinetic conditions.
Lastly the residence time of both the waste components
and the oxidation air while they are in contact with each
other is a very critical factor. These three factors are so
called the 3-T principle: (residence) Time, Temperature, and
Turbulent mixing. All these three requirements have to be
met in order to have a complete combustion.
In general, the conditions of 3-T principle could be
fulfilled in most of the part of the combustion chambers.
However, according to Niessen (1995), the local conditions
are not always meeting all the three conditions in the
complicated incineration system, and there may exist
always some cold spots, or poorly mixed packets, or
channelling streams (short circuits) in the combustion
chambers. These are often the cause of the operation failures
for the incomplete destruction and combustion of the

28

Y. Yang, M.J.A. Pijnenborg, M.A. Reuter and J. Verwoerd

waste. Therefore, the transport phenomena within the whole


combustion system play a very critical role in the waste
incineration process. To catch the information about
distribution of turbulent mixing, temperature and Residence
Time Distribution (RTD) within the whole combustion
space, CFD simulation can provide a good solution.
Hazardous waste in various forms are fed into the rotary
kiln through a number of burners and injection lances/pipes,
or even the loading door so called load chute, as is described
previously. Combustion air is supplied partially together
with the waste as carrier gas and primary air, and partially
supplied through load chute at the kiln front as well as
separate injection lances in SCC. To promote earlier
combustion, a supporting burner (also called main burner at
AVR) is installed to burn high calorific waste or fuels to
ensure a high temperature zone near the feeding end of the
kiln. The cold waste in liquid or solid form is firstly heated
up by the radiation flame and walls, it gets dried and solid
vapourisation/gasification takes place before combustion
reactions. The main combustion will take place in the
gas phase between the fuel vapour and combustion air.
The mixing between the fuels vapour and the air is not
perfect, which will form partially diffusion and partially
pre-mixed flames. Due to the insufficient length of
the kiln and thus the short residence time for the mixture of
the fuel vapour and air, the combustion will continue in the
SCC. To promote complete combustion reactions, additional
air lances are installed. After certain travelling time, the
remaining fuel components including CO gas generated
as intermediate combustion product at relatively high
temperatures (10001100oC) are expected to react with
oxygen before they get into the WHB. It is prescribed by the
regulation that a minimum of 2 seconds is required for the
gas mixture at minimum of 850C for non-halogen waste or
1100C for halogen-containing waste after the last air
injection point in the furnace (Directive 2000/76/EC, 2000).
It is very important that the remaining CO (very critical for
the environmental quality of the off-gas) is fully combusted,
since it cannot be combusted any further downstream in the
APC devices downstream. In practice, minor amount of CO
in the off-gas is often the main cause for the operation
failure.

CFD simulation

The details of the coupled transport phenomena in the


hazardous waste incinerator are very difficult to obtain
through online measurement and high temperature
experiments. CFD simulation offers a lot of benefits for
investigating coupled transport phenomena of fluid flow,
heat transfer and chemical reactions in the incineration
system. However, care needs to be taken of the validity of
physical models and validation of the results, particularly
due to the empiricism introduced by the models of
turbulence, multi-phase flow and fuel combustion.
In the current study, the simulation of gas flow and
temperature distribution as well as waste combustion
in the incinerator has been conducted with CFD code

PHOENICS 3.5 (Cham, 2004). The governing partial


differential equations for conservation of mass, momentum,
and energy in a turbulent flow system are expressed by
time-averaged 2nd order partial differential equations.
The general form is expressed in Equation (1), and it is
solved with a common numerical algorithm in the code. Due
to the turbulent nature of the flow, different turbulence
models such as the standard k- and its modifications were
tested, and the standard k- model was finally chosen in the
study. Due to the assumed adiabatic wall boundary
conditions to be explained later, the simulation with
radiation model did not bring any difference for heat
transfer with PHOENICS and has been thus neglected in the
combustion heat transfer model. This is justified by the fact
that a relatively low heat is lost in the furnace operation.

( ) + div( u ) = div( ,eff grad ) + S


t

(1)

where

a general flow variable (velocity component,


temperature, turbulence kinetic energy or its
dissipation rate, radiosity or radiation temperature,
etc.)
u:
the velocity vector
:
density of the fluid phase
,eff: the effective exchange coefficient, which is the sum
of laminar and turbulent exchange coefficients
general source term for flow variable .
S:
For the flow simulation, the continuity equation and the
equation of momentum conservation (the Navier-Stokes
equations) are solved together with a turbulence model.
For heat transfer simulation, the energy conservation
equation ( becomes enthalpy) is solved, where the
source term S includes the gas phase combustion energy
estimated from a combustion model (global 3-gas or 7-gas
combustion model in this study). The combustion model
will be discussed in more detail.
Apart from waste combustion modelling in the gas
phase, separate simulation on the flow of shredded solid
waste was conducted by using the Inter-Phase Slip
Algorithm IPSA, the Eulerian Eulerian two phase
model of PHOENICS (Cham, 2004). In the current study,
only the isothermal particle flows were simulated. In order
to include the effect of gas temperature variations
(gas cooling, and non-isothermal gas mixing) on the gas
flow field, the convective heat transfer within the gas phase
(non-isothermal gas flow) was included in the simulations.
Combustion models were not included in the gas phase flow
and heat transfer simulation, while the combustion effect
was represented with hot gas jets where the incoming
temperatures were estimated from the 7-gas combustion
model. The effect of gravity on the particle momentum
conservation is included. The gas phase turbulence was
simulated with the standard k- model, however, the effect
of the particles presence on the gas phase turbulence was
neglected. In the current simulation, the inter-phase mass
and heat transfer were not taken into account.

Analysis of transport phenomena in a rotary-kiln hazardous waste incinerator

29

Table 1 summarises the assumptions, the physical


models, boundary conditions, and other model parameters,
for the simulations presented in this paper.

grid of 359,100 cells was used to study effect of extra


cooling air from the ventilation system to the rotary
kiln cooling ring.

Table 1

4.2 Waste combustion modelling

Summary of the general simulation conditions

Item

Model information

Waste processing
rate

7,200 kg/h and 30.51 MW thermal capacity

Waste combustion Gas phase combustion, 3-gas and 7-gas


models
combustion models
Shredded solid
waste

In the combustion model assumed as


vapourised gas fuel. In the flow simulation,
modelled with IPSA 2-fluid model

Turbulence

Standard k- model, in both combustion and


IPSA solid flow simulations

Radiation

Neglected in the combustion simulation, due


to the adiabatic wall boundary conditions

Wall boundary
conditions

Adiabatic for all the walls of the kiln and


Secondary Combustion Chamber

4.1 Computational grid


As can be seen in Figure 2, the incinerator consists of a
rotary kiln with a cylindrical shape and a SCC with a
rectangular shape. A Cartesian grid was used to construct
the incinerator model, with a total 230,394 of cells. In order
to obtain the approximate cylindrical shape of the rotary
kiln, solid blocks were used to form the inactive regions of
the geometry. The computational geometry and grid are
shown in Figure 3, where the cell distribution of a side view
of the rotary kiln and the SCC, as well as the front view of
the kiln is illustrated.
Figure 3

Furnace geometry and the computational grid


generated in the CFD model (57 64 86
= 230,394 cells)

In order to include the thermal behaviour and major species


distribution, waste incineration process has been modelled
with two available combustion models from the CFD code,
the 3-gas and the 7-gas combustion models, as described
below. The waste materials were assumed to combust
in the gas phase, and the gasification and vapourisation
process of the wastes was assumed to be complete upon
entering the incinerator. The solid particle or droplet
flows were not physically modelled. The heat loss through
the furnace walls and energy absorbed into the molten
slag was estimated from the overall energy balance and
thermodynamics, and the heat loss was subtracted from the
energy input of individual waste streams. The errors brought
into the results of the simulation models will be discussed in
the model validation. Due to relatively small heat loss
of the system (510%), the incinerator system was assumed
to be adiabatic, and thus no radiation model was used.
Although both combustion models could not physically and
chemically represent realistic incineration process, they do
offer a practical approximation and give a good indication
of the temperature and species distribution.

4.3 3-gas combustion model


For combustion modelling, a global combustion model
of Simple Chemically-Reacting System (SCRS, or 3-gas
model) of Spalding (Spalding, 1979) is used for chemical
waste combustion reactions in the earlier phase of the
project. In the model, wastes of different types are modelled
as vapourised fuels (gases) upon entry. SCRS model
focuses on the overall effects of combustion, and it cannot
give any detail of the combustion mechanism and chemical
species. It involves a reaction between two reactants
(fuel and oxidant) in which they combine, in fixed
proportions by mass, to produce a unique product:
1kg fuel + s kg oxidant = (1 + s ) kg product + heat

The location and definition of various burners and air inlets


are also indicated. Since each waste stream has different
thermal value and chemical composition, a fuel-rich and a
fuel lean streams were used in different portions to obtain
the individual burner input, a central fuel-rich stream
surrounded by four fuel-lean streams. A proper definition of
a fuel-rich and a fuel-lean streams is required by the 7-gas
combustion model of PHOENICS which handles only one
fuel in a combustion model. In addition, a similar but finer

(2)

where s is the stoichiometric oxidant requirement


(kg oxidant/kg fuel). This reaction is taken as irreversible.
Reaction rates in turbulent flow situations are often more
greatly affected by local turbulence than by chemical
factors. For these situations the Eddy Break-Up (EBU)
model is provided, which rests on the hypothesis that only
turbulence and fuel concentration affect the reaction rate
having the following source term in the mass conservation
equation:

S mfu = CEBU min{m fu mox / s} .


k

(3)

Where CEBU is the Eddy Break-Up reaction constant, mfu is


the mass fraction of unburned fuel, mox is the oxidant
fraction. k and are the turbulence kinetic energy and its

30

Y. Yang, M.J.A. Pijnenborg, M.A. Reuter and J. Verwoerd

dissipation rate. In the current study it is assumed that the


reaction rate for the combustion of the different waste types
is turbulent mixing rate limited and the EBU is chosen to
determine the reaction rate. On the basis of extensive
parametric testing a value at 2550 for CEBU was found to
be reasonable. The stoichiometric ratio s, was taken with
value of 6.52, based on the information in the literature
(Wardenier and van den Bulck, 1997) and estimation from
the average heating value of the waste at AVR.

4.4 7-gas combustion model


More recently, an improved combustion model was studied
in order to understand and predict some chemical aspects of
the waste combustion process. Not only the overall thermal
behaviour could be estimated, but the species distribution
could be predicted as well, such as fuel (hydrocarbons),
oxidant (O2), intermediate and final combustion products
(CO, H2, CO2, H2O), and the inert component of nitrogen
(N2). An extended SCRS model from Phoenics is explored
for this purpose. In order to model the species distribution
together with combustion, the first step is to convert the
different waste streams into a sort of virtual fuel or fuel
mixture. Then the virtual fuel is used for combustion
reaction modelling with both equivalent thermal and
chemical contribution as the complex wastes. A similar
concept has been proposed and used for Municipal Solid
Waste (MSW) incineration (Themelis and Kim, 2003), and
this proved to be a very successful approach in handling
the complicated waste materials. The details of the 7-gas
combustion model are described later in this paper.

4.5 Solid flow model for the shredded solid waste


In the present study, the standard drag coefficient for
spherical solid particles moving in a gas phase suggested by
Clift et al. (1978) was chosen. The drag coefficient Cd is
determined by the particle Reynolds number Re:
Cd =

24.0
0.42
(1.0 + 0.15 Re0.687 ) +
.
Re
1.0 + 4.25 104 Re 1.16

(4)

Equation (4) is valid for the Reynolds number of


Re  3.38 105. The particle Reynolds number is based on
the slip velocity Vslip (the velocity difference between the
particle and the carrier gas), the particle diameter Dp and the
kinematic viscosity of the gas phase g:
Re =

Vslip D p

(5)

In the particle gas incineration system, the maximum


particle Reynolds number is estimated to be around
105, within the boundary value in the criterion
(Re  3.38 105). The highest particle Reynolds number

lies in the areas near the burner zones where gas velocity is
very high.
Both steady and transient situations of solid waste flow
were simulated. For transient studies, the time interval of
0.11 second was used with 28 time steps to catch the
smooth change of the particle flow, especially in the first
second of the transient flow. In total 15 seconds was
simulated in most of the transient models, which could
approach roughly (if not completely) the steady flow of the
solid. In these studies, the steady gas phase flow was firstly
solved for the single gas-phase flow. Afterwards, the
particles were added into the system from the solid waste
burner. The variation of the particle volume fraction was
followed to see the development of the solid flowing
process and the statistical particle trajectories, especially the
dispersion extent of the solid waste in the system.

The Virtual Fuel and waste combustion


models

5.1 Virtual fuel determination


Based on the off-gas composition and mass and energy
balance of each waste streams at AVR-Chemie, a virtual
fuel was determined. The organic compound was estimated
as C102H131O110, however, a hydrocarbon has to be found
due to the limitation of the CFD code that it cannot handle
oxygen at present in the organic compounds. To get the
generalised fuel, two constraints were used:

heat of combustion = 42,000 kJ/kg for the dry mass

H/C molar ratio is in a range of 1.31.7.

For all closely related real existing hydrocarbons


calculations were carried out to determine the generalised
virtual fuel composition. Figure 4 illustrate the concept,
where one can see how a poorly defined waste is converted
to the better defined fuel. To get the balance of oxygen in
the waste, oxygen is introduced to the combustion streams
through other oxygen containing compounds such as CO,
O2, or H2O. Combinations with C2H3 or C3H4 with other
components and for all incoming streams are put in the
solver of Microsoft Excel, both C2H3 and C3H4, as
the main virtual fuel component, are found to fit to the
constraints. In this paper further calculations are performed
with C3H4. C2H3 and C3H4 have very similar heat of
combustion to methane (CH4), and are in gaseous form at
room temperature. Here it may be necessary to emphasise
that the hydrocarbon C2H3 or C3H4 is only the main
component of the fuel streams, and the used fuel streams in
the model are defined as a mixture of the main component.
Other allowable fuel component (by the CFD code) of CO
and non-combustibles such as CO2, H2O, O2 and N2, as will
be explained in the following section (waste stream
optimisation).

Analysis of transport phenomena in a rotary-kiln hazardous waste incinerator


Figure 4

Concept of artificial fuel definition for the hazardous


waste incineration plant

31

Step 2: Secondary reactions


2CO + O 2 = 2CO 2

(7)

2H 2 + O 2 = 2H 2 O.

(8)

The reactions are assumed to be kinetically controlled, and


the EBU is used here to calculate the reaction rate as in the
3-gas model for the turbulent combustion system.

5.3 Waste stream optimisation

However, it should be noted that the use of a real


hydrocarbon to represent the main combustible component
is to get the species distribution of the combustion products
such as CO, CO2 and H2O. The real composition of the
wastes is much more complex and the incineration reactions
are much more complicated than the assumed virtual fuel.
A major benefit here is that it is possible to model the
complex waste combustion process, and with adjustable
model parameters, the thermal effect and distribution of
major species in the incineration system could be obtained
and evaluated.

The current CFD code PHOENICS has limited the


combustion model with only two streams: fuel-rich stream
and fuel lean stream. Therefore, air and water present in
the waste streams have to be distributed into both streams.
With these two distinct streams, all the incoming waste
and combustion air could be defined by either a pure
stream or a combination of both. For this reason a stream
optimisation has been made, and Table 2 illustrates the
optimised two streams, with which all inlet compositions
can be determined by mixing different portions of each
stream. This composition is based on the assumption that
40% of the sludge burner waste is fed as fuel rich stream,
and 60% as fuel lean stream, which indicate an extent of
premixing. Thus, the calculated percentages per inlet are
shown in Table 3. All pure air supplies through load chute
and lances consist out of 100% fuel-lean stream, and
waste feeding inlets are mixtures of the fuel-rich and
fuel-lean streams though separate but adjacent multi-inlets
combination. For instance, the main burner is composed of a
central fuel-rich stream and a few surrounded fuel-lean
stream with predefined fractions.
Table 2

5.2 Assumed combustion reactions


With the defined chemical compositions of all waste
streams, a two-step combustion reaction scheme is assumed
with a primary reaction and two secondary reactions shown
as follows:
Step 1: Primary reaction
C3 H 4 + 1.5O 2 = 3CO + 2H 2
Table 3

(6)

Chemical compositions of the optimised fuel rich


and fuel lean streams
Compositions (wt. %)

Streams

C3H4

Fuel rich

21.64 15.64

Fuel lean

0.28 20.97

O2

Heat of combustion 46.3


(MJ/kg)

CO

CO2 H2O

N2

Total

3.09 0.75 4.41 54.47 100.0


0.06 0.00 5.64 73.05 100.0
10.1

Stream distribution, the mass and thermal input for all inlet streams
Compositions (wt. %)

Input data

Streams Load chute Main burner Sludge burner Solid burner1 SCC burner Air lances Ash sump Total

Stream distribution (wt. %) Fuel rich

0.00

34.45

40.00

37.77

13.10

0.0

0.0

Fuel lean

100.0

65.55

60.00

62.23

86.90

100.0

100.0

Mass flow rate (kg/h)

Total energy input (MW)


1

27,000

11,250

3,900

8,193

6,620

9,030

4,000

69,993

Practice

0.00

11.67

4.33

11.67

2.84

0.00

0.00

30.51

Model

1.00

11.39

4.56

10.67

2.70

0.33

0.15

30.80

In definition of solid waste burner, 15.04% was assumed as inert materials reported to slag. The total mass flow rate and the
stream distribution are based on the mass flow rate to the gases phase only.
2
The mass flow rate reported here includes the wastes, oxidation air and water content. The waste processing rate is estimated at
7,200 kg/h, based on the operational data in practice.

32

Y. Yang, M.J.A. Pijnenborg, M.A. Reuter and J. Verwoerd

Table 3 shows the mass and energy input for all the
incoming waste streams as well as additional air supply.
For heat of combustion and energy input from each
incoming stream, the difference of the model data from the
real operation conditions in practice is illustrated, and the
deviation is relatively small.

Simulation results

6.1 Main flow characteristics


It is known that the flow and mixing of the combustion air
and fuel vapour determine essentially the waste combustion
Figure 5

process. The residence time of the gas particles is crucial for


completing the combustion. From the CFD combustion
models, the flow pattern and gas mixing are predicted, and
Figure 5 illustrates the predicted velocity distribution
across a few regions of the incinerator, and a clear
view of the complexity of the flow pattern and mixing
behaviour could be seen. The 3-dimensional nature and
complex mixing pattern are clearly demonstrated.
Although the flow pattern has not been validated
by flow measurements due to extremely hostile
environment within the incinerator, the prediction could
well indicate the general characteristics of the gas flow and
mixing pattern.

Gas flow pattern inside the rotary kiln and SCC during combustion process

The RTD of different gas streams and the in-flight time


distribution of each incoming streams have been
studied (Yang et al., 2002). Together with temperature
distribution maps, this could well indicate combustion
status. With PHOENICS, it is possible to calculate
the time elapsed for a fluid since its entry into the
flow domain. This time variable must only be
convected by the fluid. The accumulation of time is
accounted for by special sources, which are related to
the fluid residence time in a cell. The calculation of the
in-flight time could be easily performed in a steady
state model. Figure 6 illustrates the predicted time
distribution of the gas streams since its entry point
into the incineration system.

The RTD curve has been obtained by adding inert


tracers from the individual inlets, which follow the same
flow path as the flowing gases. Through transient simulation
by measuring the outgoing tracer amount along time from
the outlet, the RTD curves were constructed with which the
flow characterisation was performed for the incinerator
(mean residence time, dead volume etc.). A typical RTD
curve obtained for the incinerator is illustrated in Figure 7
for the main burner combustion stream (E1()) and the load
chute air stream (E2()). Based on this analysis, a dead
volume of 16% and 20% were identified for the above two
streams. The incinerator is essentially a back-mixed flow
reactor with a back-mixing portion of over 75% for the
studied streams.

Analysis of transport phenomena in a rotary-kiln hazardous waste incinerator


Figure 6

Gas in-flight time distribution since its entry into the


incinerator system (based on the 3-gas combustion
model)

Figure 7

RTD curve, obtained from the transient simulation


model (numerical tracer experiment)

the solid bed for part of the solid waste could be clearly
seen. However, in the 3-gas model, all the solid waste was
fed through the load chute, and the combustion was
assumed to take place completely on a solid bed at the
bottom of the rotary kiln.
It is important to notice the distribution of the main
combustion air, coming from the load-chute. Due to the
physical arrangement, the air enters the kiln front from the
load-chute at an angle of 50 downward. This causes a poor
mixing of the air with waste vapours. From the temperature
distribution maps, one can see the colder zone at the lower
part of the kiln from both Figures 8 and 9. In addition, rather
high temperature gradients of the combustion system can be
observed. The hotter zones in the kiln lie in the upper-half,
and very long flames were formed with a maximum
temperature above 2000C. Combustion from the solid-bed
waste takes place somewhat downstream away from the
bed, due to the cooling effect of the load-chute air.
However, the maximum temperature from the flames is
highly dependent on the rate constant (CEBU) of the EBU
model. In the model shown here, the CEBU constant from
both the primary and secondary reactions was set at 2.0.
The constant at various levels have been tested, the
fine-tuning of the constant needs to be conducted carefully
in the next stage. The cooling effect by the cold air from
injection lances seems to be significant, and this effect could
extend up to the outlet region.
Figure 8

6.2 Temperature distribution


With the 7-gas model, both the thermal and chemical
behaviour of the waste combustion can be simulated.
Various scenarios and parameters have been examined, such
as the premixing portion of fuel and combustion air,
distribution of solid-waste stream between the burner and a
solid-bed which may form at the bottom of the kiln. For the
later case, a variation between the pure solid-waste burner
and pure (100%) solid-bed has been investigated. As an
illustrative example, the shredded solid waste assumed to be
equally distributed between the solid-waste burner and a
solid-bed is used as the show-case in this paper. Figure 8
illustrates the temperature distribution across the various
cross-sections of the incinerator. Figure 9 shows the cross
sections from a few essential planes of the incinerator.
The combustion flames from all burners or lances as well as

33

Predicted temperature distribution (C) across the main


burner and the SCC burner (7-gas model)

34
Figure 9

Y. Yang, M.J.A. Pijnenborg, M.A. Reuter and J. Verwoerd


Temperature distribution (C) across various burners, the load-chute and near the outlet (7-gas model)

6.3 Distribution of chemical species


The 7-gas combustion model provides the mass-fraction
distribution of all defined seven chemical species: main fuel
component C3H4, two intermediate products (also as the
fuels for the secondary combustion reactions) CO and H2,
final combustion products CO2 and H2O, and oxidant O2
and inert species N2 brought by air. C3H4 is present in all
fuel-rich streams, but also in minor amounts in the fuel lean
streams. CO originates both from the fuel-rich and fuel-lean
streams, and from the primary combustion reaction of C3H4.
From the air lances, minor amount of CO could be
observed. It should be noted that H2O and CO2 originate
from both fuel-rich streams and the secondary combustion
reactions.
According to the detailed analysis of the simulation
results, all the primary and secondary fuel components are
reacted quickly upon entry to the system. Since all the
burner streams contain certain amount of oxygen (O2),
partial combustion takes place at early stage upon entry as
partially premixed flames. However, the requirement of
more air from the surrounding air inlets and majority O2

supply from load-chute and air lances makes the combustion


also partially diffusion-controlled. Therefore, air supply and
mixing with the fuel streams become very important.
Figure 10 illustrates the species distribution along the
main-burner axis plane and the line profiles along the burner
axis. Quick consumption of the fuel components can be
clearly seen. The CO and H2 concentrations depend very
much on the relative generation and destruction rate.
At present, all the rate constants were set equal, and
thus C3H4, CO and H2 follow similar trend. However,
if the reaction rate of primary combustion is higher
than the secondary combustion reactions, CO and H2 will
exist more appreciably for a long time (more downstream).
This will be tested further in the following stage of research.
Also obvious is the peak values of both CO2 and H2O as
final combustion products and CO as an intermediate
product lie in the location about 3 m from the burner exit
down stream. The minimum value of CO2 at about 6 m
downstream from burner axis is due to the air mixing
(bending up from the kiln floor originated from the
load-chute).

Analysis of transport phenomena in a rotary-kiln hazardous waste incinerator


Figure 10 Distribution of CO mass fraction across the main
burner plane

35

In addition, other chemical species are calculated from the


7-gas model, and typical distributions of the main fuel and
products of the combustion are illustrated in Figure 12. H2 is
not present in the main fuel, and it is only generated as the
secondary fuel component and is quickly combusted. It has
almost the same distribution pattern as CO.
Figure 12 Distribution of mass fraction of other chemical species
as predicted by the 7-gas combustion model

From the distribution of CO mass fraction across the main


burner plane in Figure 10, only near the burner zones,
appreciable CO could be observed. In the majority areas,
CO concentration is very close to zero. For a better
resolution, a closer look at the lower concentration range is
shown in Figure 11. Attention should be paid to the CO
distribution in the SCC, because of its special status in
emissions control. It can be seen that high CO concentration
is caused either by the burner fuel-rich streams, or by the air
lances. However the latter is an artificial effect due
to the minor presence of CO in the air stream, and is not
present in the practice. Near the outlet, the CO fraction falls
below 3 ppm (wt.). At least in this case, the CO is well
combusted in the SCC, according to the simulation
results. However, more quantitative study needs to be
carried out with fine-tuning of the model against future
measurement data.

6.4 Parametric studies


In addition to the case shown above, effects of various
parametric changes have been studied. The focus was on the
following aspects:

Distribution of the solid waste between the solid waste


burner and the solid bed. The variation covers the solid
waste entry from 100% via the solid waste burner to
100% via the solid bed. This variation causes changes
in flame geometry and energy distribution within the
kiln, since the solid waste burns more slowly in the
solid bed at lower temperatures. But more realistic
distribution needs to be further studied, by using
a two-phase flow model in the future.

Rate constant (CEBU) of the EBU model of the


turbulent combustion system: The tests for the CEBU
value of 0.1, 0.5, 1.0, 2.0 and 4.0 have been conducted.
It was found that the rate constant in this range does not
bring significant changes in the maximum flame
temperature, except the case of CEBU = 0.1. However,
model validation against temperature measurements
shows that CEBU = 1.0 gives better fit to the measured
temperature profile.

Figure 11 Distribution of CO mass fraction in the Secondary


Combustion Chamber

Air mixing was found to be important for the waste


combustion. Therefore, the incoming angle of air from
the load-chute at 50 downward from the horizontal and
the pure horizontal cases were compared. It was found
that horizontal air supply improves the mixing.

36

Y. Yang, M.J.A. Pijnenborg, M.A. Reuter and J. Verwoerd

Air preheating: Air preheating was expected to


have a big influence on thermal homogenisation. Air
preheating at different levels were tested from 50C, to
100C, 200C, 300C. It shows that the air preheating
up to 200C results in a big improvement in reducing
the temperature stratification. This will help to form
more uniform combustion reactions, and reduce CO
concentration possibly originated from low temperature
zones.

Introducing extra cooling air to the kiln: In order to


safely utilise the ventilation air from the shredding
machine, which contains small amount of combustibles,
it is introduced into the rotary kiln through a cooling
ring. The influence of this air stream on the combustion
and the temperature distribution was studied at different
flow rate of 20006500 Nm3/hr. The results show that
introducing certain amount of extra air through the
cooling ring reduces the temperature gradient near the
kiln front, and the cooling effect can also protect the
refractory lining from overheating. However, too much
extra air (e.g., 6500 Nm3/hr) will increase the
temperature gradient again and cool the zone near the
kiln front.

For non-isothermal and non-uniform gas flow system,


simple averaging by cross sectional area could not give
correct mean temperature data. Therefore, a mass flow
weighted approach is developed to obtain the average
temperature and other flow variables (e.g., chemical
species). A typical average temperature profile along the
main flow direction of the rotary kiln and the SCC can be
illustrated in Figure 13. This can be very useful for design
and control purposes of industrial furnaces. The data points
in the figure also include the maximum and minimum
temperatures at each cross-section, and a temperature profile
simply averaged with cross-sectional area. The difference
between the two different averaging methods could be
clearly seen.
Figure 13 Comparison of temperature profiles along the main
flow direction of the incinerator

Air incoming angle from the load chute: The primary


combustion air from the load chute is introduced to the
rotary kiln at an angle of 50 downward from the
horizontal (50). The simulation results indicate a
large temperature gradient and poor mixing of the air
with the waste streams near the kiln front. Thus
variation of the air incoming angle was studied
from 50, 10, 0, +10 and +50 from the horizontal.
The study indicates that horizontal inflow of the air
results in the least radial temperature gradients, but
also in less mixing with the waste streams. The angle
change has to comply with the overall feasibility in
practice and a more systematic study should be
conducted in accordance with the incoming direction of
all the waste streams.

6.5 Post processing: statistical analysis


As it is well known, CFD models provide a lot of
information in a standard form of various graphical formats
and data profiles. However, the large amount of output data
are difficult to comprehend for engineering purposes such as
in design and in actual process control situation. Therefore,
proper post-processing of the results is required to produce
useful and handy data sets, which take simple form
and are easy to understand and use. By using proper
statistical methods, a large amount of CFD output data can
be condensed and average information and performance of a
reactor could be quickly demonstrated. The condensed
information can also be used to construct a database for
process control. The following example of temperature
averaging illustrates the engineering value of CFD
predictions and the possible use of CFD simulation to assist
process control.

Furthermore, a vector approach is developed to point


out the average location, an indication of the bias about
the geometrical axis of the reactor (Yang et al., 2003). The
cross section and its symmetrical axis are attached to a
coordinate system. By means of integration over the whole
cross section for the enthalpy and mass flow rate, the centre
of enthalpy can be calculated and coupled to a temperature
vector. This will make the average temperature a function of
the x and y coordinates at a given z-location along the
reactor axis (suppose z-coordinate is main flow direction).
The centre of the average temperature can be calculated
based on the temperature moment over the coordinates.

Analysis of transport phenomena in a rotary-kiln hazardous waste incinerator

6.6 Flow of the shredded solid waste


From the Eulerian-Eulerian 2-phase flow model, the volume
fractions of both gas and solid phases can be computed. For
the hazardous waste incineration kiln, distribution of the
volume fraction of the shredded solid waste gives a good
indication about where the solid waste could be dispersed
and likely be combusted if they are mixed with the
oxidation air.
Figure 14 shows the solid volume fraction of 20 mm
particles across the solid burner plane from transient
2-phase flow models after 15 seconds of solid injection, for
both low and high injection velocities: (a) with a low
injection velocity of 5.5 m/s, (b) with a high injection
velocity of 55 m/s with the same total solid mass flow rate.
It can be seen that the lower injection velocity leads to
much earlier drop of the solid waste particles to the bottom
of the kiln, and a clear solid bed could be formed. When the
injection velocity is 10 times higher, a much faster solid
stream could be seen and the solid waste will fly to and
settle on the back wall of the SCC and further down to the
ash sump. However, a thinner solid bed is still formed.
Similar distribution patterns are also observed in the steady
flow models. To confirm the inject velocity effect, two
injection velocity for the lower feeding rate of 2.4 t/h with a
solid density of 1400 kg/m3 was simulated. Results indicate
that the five times higher injection velocity confirmed the
expected solid flow pattern. Higher solid injection velocity

37

by introducing more carrier gas is an effective method for


solid dispersion.
To see the particle size effect on the solid particle
dispersion, three different particle sizes at low injection
velocity of 5.5 m/s were studied: 5 mm, 10 mm, and 20 mm.
However, no obvious difference is found in this range.
In principle, larger particles tend to fall down to the bottom
walls due to the greater volume and the resultant gravity
force. However, the effect may be dumped when the particle
momentum is already very low at the injection point. In the
future, high gas injection velocity will be simulated to see
the influence of the particle size.
Compared to particle size, particle density has more
significant effect on the particle flow and accumulation.
In the current study, variation of particle density from 1200,
1400, and 4000 kg/m3 has been investigated under steady
state conditions. Results indicate that increase in density
from 1200 kg/m3 to 1400 kg/m3, a visible difference could
already be seen. Compared with a higher density of
4000 kg/m3 a more significant difference can be found.
Solid particles with higher density have more momentum
(inertia) if they are injected at the same horizontal velocity
(which is the case here), and the particles can travel longer
distance before they are dragged down to the bottom by the
gravitational force. In practice, heavier particles may not
gain the same injection velocity as for the lighter particles
due to the inter-phase slip effect.

Figure 14 Illustration of the accumulation of the shredded solid waste (dp = 20 mm): solid volume fraction across the solid burner plane:
transient model after 15 seconds solid injection (a) low injection velocity (5.5 m/s) and (b) high injection velocity (55 m/s)

Model validation

Model validation plays a critical role in all CFD studies.


In the current model system, it is important to obtain
measurement data both for gas flow and temperature as well
as chemical species distribution. Because of the hostile
environment inside the incinerator and technical restrictions,
only the exit plane of the kiln and the SCC furnace interior
could be accessed through viewports. At present, only
temperature was measured with radiation pyrometer and
thermal couples. In the current project, two temperature
measurement campaigns were carried out at AVR-Chemie

waste incinerator through the viewports. Both 2-meter and


5-metre long thermocouples were used, and the 5-meter
long thermocouple was designed to use flowing water as
cooling medium. Along with the measurement campaigns,
temperature recordings from permanent thermocouples at
SCC were analysed to see the process dynamics, and
general trend of process stability was observed.
In order to validate the combustion model, validation
cases from both 3-gas and 7-gas combustion models were
built. The predicted temperature profiles were compared
with the measurement data, as is illustrated in Figure 15.

38

Y. Yang, M.J.A. Pijnenborg, M.A. Reuter and J. Verwoerd

During the measurement campaign, the burner for shredded


solid waste has not been installed. The total energy input
was about 37.06 MW, and the waste processing rate was
6.67 ton per hour.
Since the temperature measurement lasted about
2 hours, the process was not really running steadily as was
monitored from the permanent thermocouples, in spite of
the great effort from the operators. If taking into account the
process fluctuations, it can be seen that the model
predictions are in a reasonable agreement with the measured
data. The general trend is well predicted, and the larger
discrepancy in all 3-cases near the SCC burner side is
caused by the remaining air from the load chute. The 7-gas
combustion model predicted somewhat higher temperature
in general than the measured data, if heat loss is not
subtracted in the model. Negligence of the heat loss from
the furnace walls and by the slag formation as well as dust
carry-over is estimated approximately at 10%, which could
drag the temperature down roughly by 120C. This means
also that the 3-gas combustion model may have predicted
lower temperatures than the measured data. The 7-gas
model prediction with 10% heat loss shows a better
agreement with the measurements.
Figure 15 Temperature profile crossing view-ports of the rotary
kiln incinerator at AVR-Chemie

Besides the temperature validation, verification of species


distribution is another important aspect for the simulation.
However, the measurement of the off-gas compositions
could not be arranged at AVR-Chemie since its closure.
Then the 7-gas combustion model could be calibrated more
reliably with arrangement somewhere else with similar
operations. In the end, both the thermal and chemical
behaviour of the incineration system could be better
predicted and controlled. For the same reason, the solid
flow model could not be validated through any onsite
particle flow measurement, which in fact is very difficult to
perform. Laboratory down-scaled cold models would be a
future option to validate the shredded solid waste flow.

Summary

Through CFD simulation of the rotary kiln waste


incinerator, more understanding was gained about the

waste combustion process. The incinerator is a highly


heterogeneous reactor, with high gradients in combustion
species and temperature, especially along the rotary kiln, the
primary combustion space. Simulation results indicate that
the gas flow is clearly 3-dimensional and complicated
because of the multi-gas streams and strong heterogeneous
reactive nature of the system, and there is a large room to
improve the mixing of the air and the waste streams.
The 3-gas combustion model, especially the 7-gas
combustion model could provide a lot of information about
the incineration process, provided that the waste streams
with very complex nature are properly defined both in
chemical and thermal aspects. The present concept of
the virtual fuel definition is a novel approach and proved to
be a useful exercise and a testing method for the fuel
optimisation, and any of the similar steps is a precondition
for more detailed combustion modelling of the complex
waste materials. Temperature measurements provided very
useful information for calibrating the CFD model, and
reasonable agreement has been reached between the
measured and predicted temperature data. It is obvious
that the detailed combustion modelling of the multi-phase
reacting system still requires more in-depth research, and
further work with multi-phase flow and combustion of the
shredded solid waste is highly needed.
Through the present study it is found that one of the
most effective ways to increase the solid particle flow
velocity and hence increase the particle dispersion, is to
increase the particle injection velocity. This is already
demonstrated by an extreme case at very high injection
velocity and other models from both steady and transient
simulations. However, it should be noted that too high
injection velocity will result in a shorter residence time of
the particle and less effective suspension combustion.
An optimal injection velocity has to be found to gain a
higher dispersion and a longer mean residence time. Further
research is required for thorough parametric studies and
various design alterations.

Acknowledgements
The authors wish to thank the financial support from
AVR-Chemie in the Netherlands and its permission to
publish the work. The authors are also grateful to the
engineering staff of AVR-Chemie for their interests and
support to this work.

References
Cham (2004) Phoenics: Parabolic Hyperbolic or Elliptic
Numerical Integration Code Series, http://www.cham.co.uk,
accessed in September 2004.
Chen, K.S. and Lee, M. (1995) Visualization studies of
three dimensional flow and solid motion in rotary kiln,
Hazardous Waste and Hazardous Materials, Vol. 12, No. 4,
pp.395409.
Clift, R., Grace, J.R. and Webber, M.E. (1978) Bubbles, Drops and
Particles, Academic Press, New York.

Analysis of transport phenomena in a rotary-kiln hazardous waste incinerator


Directive 2000/76/EC (2000) European Parliament and the
Council, of 4 December 2000 on the incineration of waste,
Official Journal of the European Communities, Vol. 288,
No. 12, pp.L332/91111.
Ficarella, A. and Laforgia, D. (2000) Numerical simulation of
flow-field and dioxins chemistry for incineration plants
and experimental investigation, Waste Management, Vol. 20,
No. 1, pp.2749.
Jakway, A.L., Sterling, A.M., Cundy, V.A. and Cook, C.A. (1996)
Three-dimensional numerical modeling of a field-scale
rotary kiln incinerator, Environmental Science and
Technology, Vol. 30, No. 5, pp.16991712.
Jenkins, B.G. and Molar, F.D. (1980) Modelling of heat transfer
from a large enclosed flame in a rotary kiln, Chemical
Engineering Research and Design, Vol. 59, No. 1, pp.1725.
Khan, J.A., Pal, D. and Morse, J.S. (1993) Numerical modeling of
a rotary kiln incinerator, Hazardous Waste and Hazardous
Materials, Vol. 10, No. 1, pp.8195.
Leger, C.B., Cundy, V.A. and Sterling, A.M. (1993) A three
dimensional detailed numerical model of a field-scale rotary
kiln incinerator, Environmental Science and Technology,
Vol. 27, No. 4, pp.677690.
Niessen, W.R. (1995) Combustion and Incineration Processes,
Marcel Dekker Inc., New York.
Spalding, D.B. (1979) Combustion and Mass Transfer, Pergamon
Press, Oxford.
Themelis, N.J. and Kim, Y.H. (2003) Modeling of transport and
chemical rate phenomena in flash combustion of municipal
solid wastes, in Sohn, H.Y., Itagaki, K., Yamauchi, C. and
Kongoli, F. (Eds.): Proceedings of Yazawa International
Symposium: Metallurgical and Materials Processing
Principles and Technologies Vol. 1: Materials Processing
Fundamentals and New Technologies, Warrendale, PA, TMS,
pp.10311051.

39

Veranth, J.M., Dachun, G. and Silcox, G.D. (1996) Field


investigation of the temperature distribution in a commercial
hazardous waste slagging rotary kiln, Environmental Science
and Engineering, Vol. 30, No. 10, pp.30533060.
Veranth, J.M., Silcox, G.D. and Pershing, D.W. (1997) Numerical
modeling of the temperature distribution in a commercial
hazardous waste slagging rotary kiln, Environmental Science
and Technology, Vol. 31, No. 9, pp.25342539.
Wardenier, K. and van den Bulck, E. (1997) Steady-state waste
combustion and air flow optimization in a field scale rotary
kiln, Environmental Engineering Science, Vol. 14, No. 1,
pp.4354.
Williams, J.D., Becker, A.R. and Gerovich, M.J. (1988) 3-D flow
modeling of a hazardous waste incinerator, JAPCA, Vol. 38,
No. 8, pp.10501054.
Wolbach, C.D. and Garman, A.R. (1984) Modeling of destruction
efficiency in a pilot-scale combustor, American Flame
Research Committee International Symposium on Alternative
Fuels and Hazardous Waste, Tulsa, Oklahoma.
Yang, Y., Reuter, M.A. and Hartman, D.T.M. (2001) Combustion
modeling of a rotary kiln waste incinerator with phoenics,
Phoenics Journal of Computational Fluid Dynamics and its
Applications, Vol. 4, No. 1, pp.5381.
Yang, Y., Reuter, M.A. and Hartman, D.T.M. (2003) CFD
modelling for control of a chemical waste rotary kiln
incinerator, Control Engineering Practice, Vol. 11, No. 1,
pp.93101.
Yang, Y., Reuter, M.A., Voncken, J.H.L. and Verwoerd, J. (2002)
Understanding of hazardous waste incineration through
computational fluid-dynamics simulation, Journal of
Environmental Science and Health, Part A: Toxic/Hazardous
Substance and Environmental Engineering, Vol. A37, No. 4,
pp.693705.

Vous aimerez peut-être aussi