You are on page 1of 19

Waste Management xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Waste Management
journal homepage: www.elsevier.com/locate/wasman

A review of technologies and performances of thermal treatment


systems for energy recovery from waste
Lidia Lombardi a,⇑, Ennio Carnevale b, Andrea Corti c
a
Niccolò Cusano University, via Don Carlo Gnocchi, 3, 00166 Rome, Italy
b
Industrial Engineering Department, University of Florence, via Santa Marta, 3, 50129 Florence, Italy
c
Department of Information Engineering and Mathematics, University of Siena, via Roma, 56, 53100, Italy

a r t i c l e i n f o a b s t r a c t

Article history: The aim of this work is to identify the current level of energy recovery through waste thermal treatment.
Available online xxxx The state of the art in energy recovery from waste was investigated, highlighting the differences for dif-
ferent types of thermal treatment, considering combustion/incineration, gasification and pyrolysis. Also
Keywords: different types of wastes – Municipal Solid Waste (MSW), Refuse Derived Fuel (RDF) or Solid Refuse Fuels
Waste-to-Energy (SRF) and some typologies of Industrial Waste (IW) (sludge, plastic scraps, etc.) – were included in the
Energy efficiency analysis. The investigation was carried out mainly reviewing papers, published in scientific journals
Combustion
and conferences, but also considering technical reports, to gather more information.
Incineration
Gasification
In particular the goal of this review work was to synthesize studies in order to compare the values of
Pyrolysis energy conversion efficiencies measured or calculated for different types of thermal processes and differ-
ent types of waste.
It emerged that the dominant type of thermal treatment is incineration associated to energy recovery
in a steam cycle. When waste gasification is applied, the produced syngas is generally combusted in a
boiler to generate steam for energy recovery in a steam cycle. For both the possibilities – incineration
or gasification – cogeneration is the mean to improve energy recovery, especially for small scale plants.
In the case of only electricity production, the achievable values are strongly dependent on the plant size:
for large plant size, where advanced technical solutions can be applied and sustained from an economic
point of view, net electric efficiency may reach values up to 30–31%. In small-medium plants, net electric
efficiency is constrained by scale effect and remains at values around 20–24%. Other types of technical
solutions – gasification with syngas use in internally fired devices, pyrolysis and plasma gasification –
are less common or studied at pilot or demonstrative scale and, in any case, offer at present similar or
lower levels of energy efficiency.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction

Thermal treatment of waste is an inalienable part of every inte-


Abbreviations: ASR, automotive shredded residues; C&IW, commercial & indus-
trial waste; CHP, Combined Heat and Power; DMS, direct melting system; EfW,
grated waste management system (Porteous, 2005). European
Energy from Waste; FBC, fluidized bed combustor; FGR, Flue Gas Recirculation; FGT, strategy for waste management imposes that ‘‘the following waste
flue gas treatment; GT, gas turbine; GTCC, gas turbine combined cycle; HHV, Higher hierarchy shall apply as a priority order in waste prevention and
Heating Value; HT, high temperature; HW, hazardous waste; ICE, internal management legislation and policy: prevention; preparing for re-
combustion engine; IHW, Industrial Hazardous Waste; IW, Industrial Waste; LCA,
use; recycling; other recovery, e.g. energy recovery; and disposal’’
Life Cycle Assessment; LHV, Lower Heating Value; LT, low temperature; MBT,
Mechanical Biological Treatment; MSW, Municipal Solid Waste; P, pressure; Pcond, (Directive 2008/98/EC). Thus, it is very clear that the use of landfills
condenser pressure; Pel, electrical power output; Pth, thermal power input; RDF, must be residual and devoted to pre-treated wastes (not
Refuse Derived Fuel; SC, separate collection; SCR, selective catalytic reduction; SEP, biologically active or not containing motile hazardous substances).
specific electricity production; SHP, specific heat production; SNCR, selective non- Re-use and recycling are aimed at pursuing effective material
catalytic reduction; SRF, Solid Refuse Fuel; T, temperature; Tout, gas temperature at
boiler outlet; WtE, Waste to Energy; gel, electric efficiency; gth, thermal efficiency.
recovery. For those streams of waste, for which the material recov-
⇑ Corresponding author. Tel.: +39 366 6381000. ery is not effectively applicable, the energy recovery is the path to
E-mail address: lidia.lombardi@unicusano.it (L. Lombardi). be followed, considering also that when applying the waste

http://dx.doi.org/10.1016/j.wasman.2014.11.010
0956-053X/Ó 2014 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
2 L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx

hierarchy, ‘‘measures to encourage the options that deliver the best water content is reduced to 10%) (Lombardi et al., 2012) from the
overall environmental outcome, should be taken. This may require recycled paper process; and in general Commercial and Industrial
specific waste streams departing from the hierarchy where this is Waste (C&IW), for which Lupa et al. (2011) measured an average
justified by life-cycle thinking on the overall impacts of the gener- LHV of 9.47 GJ/Mg. Mixed plastic wastes, obtained as by-products
ation and management of such waste’’ (Directive 2008/98/EC). of the sorting process of end-of-use plastic packaging from separate
Thus an integrated waste management system should be collection, are also rather high energy content waste streams, being
designed on the integration of different types of treatment their LHV about 31.7–40.2 GJ/Mg (Arena et al., 2011).
processes: recycling processes for material recovery and, in case, Pavlas et al. (2009) state that the thermal treatment of waste
biological treatments for appropriate streams, as well as thermal with energy recovery belongs to the preferred sources of renew-
treatments for energy recovery, and should be provided with able energy and that the waste stops to be a problem becoming
service landfills for disposal of residues generated by the other an available fuel. Producing energy from waste contributes to pri-
treatments. mary energy savings in conventional utility systems (Pavlas et al.,
In this text by thermal treatment is meant any thermochemical 2010). According to this approach, two main advantages are high-
conversion process that takes place at relatively high temperatures lighted: the waste is processed and, at the same time, energy is
causing modifications in the chemical structure of the processed produced. For this reason, today, the thermal treatment plants
material. Thus the three main processes available for thermochem- associated to energy production are commonly addressed to as
ical conversion will be included within the analysis: combustion, Waste to Energy (WtE) or even Energy from Waste (EfW) plants.
gasification and pyrolysis of waste. Concerning the degree of renewability of carbon contained in
Nowadays, combustion processes, generally called incineration, the MSW (carbon is about 25% in mass for waste with LHV of
are the most commonly widespread thermal treatments applied 10 GJ/Mg), one should consider that this carbon is bound in a vari-
for different types of waste, including Municipal Solid Waste ety of materials such as food waste, garden waste, waste wood,
(MSW), intended as unsorted residual waste (i.e. the waste left paper, cardboard, textile waste and plastics. Gohlke (2009) affirms
downstream of separate collection), Solid Refuse Fuels (SRF), that more than half of the carbon is biogenic in origin, while the
Industrial Waste (IW), and Industrial Hazardous Waste (IHW). remaining part is fossil in origin, as also confirmed by C14 tech-
Incineration of waste is generally associated to energy recovery, nique analysis of WtE stack gas, that for several plants analyzed
in the form of electricity and/or heat production. Only IHW is often in the USA in 2007–2008 showed that two-thirds of the carbon
disposed by incineration (without energy recovery), since energy in US MSW is biogenic (US Department of Energy, 2007). Palstra
recovery for this waste can result quite complex due to the pres- and Meijer (2010) measured biogenic flue gas CO2 fractions within
ence of several pollutants in the generated flue gas. Obviously, 48–50% at waste incineration plant in The Netherlands and simi-
energy recovery is beneficial also for IHW, reducing operating costs larly Fellner et al. (2007) show that the ratio of biogenic energy
and external energy consumption (Stehlík, 2012), and it is applied source in MSW supplied to a plant in Austria range from 36% to
when possible. 53%.
In recent decades, the main interests toward thermal treat- Gohlke (2009) calculated that for a new WtE plant with moder-
ments were due to their ability to significantly reduce the solid ate steam parameters (48.5 MWLHV combustion power, 40 bar/
waste in mass (about 70–80%) and in volume (about 80–90%), 380 °C, no heat recovery, net electric efficiency 20.6%LHV), assuming
allowing preserving landfill space, as well as to eliminate the ten- that 56% of MSW is biogenic in origin, the specific CO2 emission is
dency of waste to putrefy giving place to sanitary problems (this about 0.4 Mg per MWh of produced electricity, and compared this
last aspect being especially important in the past) (Gohlke and values with the specific emissions of fossil fuel power plants (for
Martin, 2007). example a coal power plant can emit about 0.84 Mg of CO2 per
Nowadays, an important additional attractiveness toward MWh of produced electricity).
waste thermal treatments is given by the possibility of making sig- In this regard, some authors invite to consider that when the
nificant energy recovery, thanks to the technological developments source separation of organic waste, paper and cardboard is carried
achieved in this field (Stehlík, 2012) and, in the case of MSW, to the out successfully, the share of the renewable energy content of
increased energy content with respect to the past, because of the MSW may be lower than previously cited values (Horttanainen
change in the consumers’ habits and the increase in the upstream et al., 2013).
separate collection (Calabrò, 2010). The Lower Heating Value (LHV)
for the major part of MSW incinerated in EU passed from 10.0 to 1.1. Source of data
10.3 GJ/Mg from 2001 to 2010 (Reimann, 2012). Several waste
streams have a relatively high LHV. In the case of SRF in EU (cfr. The data used in this work come mainly from scientific litera-
CEN/TS 15359, 2006), the LHV must be more than 3 GJ/Mg and ture (international journals and conferences) and public reports.
can be higher than 25 GJ/Mg: Arena and Di Gregorio (2014) mea- Table 1 summarizes the data sources reported in the reference list
sured a LHV in the range 18.6–21.3 GJ/Mg for SRF obtained from on the basis of the considered thermal process (incineration, gasi-
MSW; while the previous term ‘‘Refuse Derived Fuel’’ (RDF) was fication, plasma, pyrolysis) and the type of document (‘‘others’’
not given by any legal definition and it was interpreted differently includes different types of reports). Also, in reference to the incin-
across countries. For Automotive Shredder Residue (ASR), classified eration process, statistics are given about how many sources con-
as IHW, the LHV lays in the range 19–29 GJ/Mg (Vermeulen et al., sider MSW rather than RDF/SRF, as well as each of the three
2011). Biostabilised sewage sludge LHV, on dry basis, may range incineration technologies later discussed (see Section 2.1). The
from 7 GJ/Mg (Werle and Wilk, 2010) to 23 GJ/Mg (Tyagi and Lo, total sum of the numbers in the table differs from the number of
2013). Other industrial waste flows, interesting for their energy the sources reported in the reference list. This is because the same
content, are scrap wood (LHV of 16 GJ/Mg); sugarcane bagasse source may refer to both MSW and RDF/SRF, as well as to more
(LHV of 18.6 GJ/Mg); plastics scraps (LHV of 32.8 GJ/Mg) (Tsai, than one technology (e.g. sources that compare incineration to gas-
2010); deinking sludge (HHV1 of 6.4–7 GJ/Mg on dry basis) (Ouadi ification). Moreover, some of the sources in the reference list deal
et al., 2013) and pulper residues (LHV of about 21 GJ/Mg, when the with waste characterization and other topics here discussed.
Among the various source documents, several data were
extracted from the Waste-to-Energy State-of-the-Art Report
1
Higher Heating Value (HHV). (2012) published by the International Solid Waste Association

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx 3

Table 1
Type of data source classified on the basis of the type of addressed thermal process.

Journals Conference Others


Combustion MSW 41 1 4
RDF 5 1 1
Grate 21 1 1
Fluidized bed 4 1
Rotary kiln 3 1
Gasification 15 1 2
Plasma 8
Pyrolysis 6 2 2

(ISWA). It contains the statistics of plants from 19 countries, 472 The incineration process can be fed directly by MSW or several
plants in Europe and 86 plants in USA, which treat MSW (or types of IW. Talking about MSW, its direct incineration is also
SRF), with a capacity of more than 15 Mg/d or 10,000 Mg/y. Plants addressed to as ‘‘mass burn’’, generally in juxtaposition to inciner-
for the treatment of other type of waste (IW, HW, sludge, etc.) that ation of mechanically pre-treated waste, where the mechanical
do not treat MSW are not considered. The authors state that, in treatment (often Mechanical Biological Treatment – MBT) (Di
their knowledge, they have considered all the European countries Lonardo et al., 2012; Ionescu et al., 2013) is aimed at improving
with WtE plants, with the only exception of Luxembourg, which the combustible quality of the waste (i.e. increasing the heating
has only one plant that treats about 130,000 Mg/y of waste. The value by reducing moisture and ash contents) and the technical
data contained in the report (ISWA, 2012) have been collected by and environmental parameters (i.e. reduction of chlorine and mer-
means of a questionnaire regarding design and operational data cury), in order to produce a generic combustible fraction or, in case
(year 2011). The response rate has been very different, not only of complying with technical standards and national laws, to pro-
from country to country, but also on the choice of which data to duce SRF/RDF.
provide. Fig. 1 shows the distribution according to the date of first It is worth to remind that waste can be burnt, without the use of
construction of the lines, for those plants that supplied this type of auxiliary fuels, when its LHV exceeds 5–7 GJ/Mg (Chen and
data in response to the ISWA (2012) questionnaire. Christensen, 2010; Komilis et al., 2013), where low values of LHV
A synthetic and clear comparison of the aims, operating condi- are mainly determined by excess of moisture and ash. If the waste
tions, process outputs and gas cleaning for the different thermal composition is described according to proximate analysis – com-
treatments – incineration, gasification and pyrolysis – are given bustibles, moisture and ash – it can be represented in a triangular
by Arena (2012). In this paper the attention is focused on how diagram (Tanner, 1965), like the one depicted in Fig. 2. In such a
energy recovery may be applied when waste undergoes to thermal representation, recently discussed by Komilis et al. (2013), a zone
treatments. Technical possibilities for improving the current level of self-combustion is identified for moisture content lower than
of energy recovery are discussed. Some examples of existing plants 50% in mass, ash content lower than 60% in mass and combustibles
are reported, with special attention to the most recent plants, higher than 25% in mass.
where particular technical arrangements are applied to seek for Also, it should be recalled that Directive 2000/76/EC requires
increased energy recovery. It is important to stress that, in this that gaseous combustion products (i.e. flue gas) are kept at a tem-
work, other criteria for selecting waste treatment technologies – perature of at least 850 °C (for non-hazardous waste and 1100 °C
as pollution issues, costs, material recovery, etc. – are not consid- for hazardous waste) for at least two seconds.
ered, as the focus is energy recovery. Waste – especially MSW – is generally a complex mixture of
Afterwards, an overview of energy recovery from waste by ther- several materials different in physical and chemical characteristics
mal treatment is reported for EU, USA and some other countries. as well as extremely variable in size. Particularly, the size of fuel
particles affects the time required for their combustion. Hence,
2. Incineration of waste with energy recovery the completion of combustion takes longer time for MSW than

Combustion/incineration is a process aimed at attaining the


complete oxidation of all the suitable elemental species encom-
passed in the feedstock material.

Fig. 1. Distribution of the age (according to the date of first construction) of the Fig. 2. Tanner diagram, used to assess the self-combustibility (shaded area) of
lines of the plants. Elaborated from data reported in ISWA (2012). municipal solid wastes during incineration (adapted from Tanner (1965)).

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
4 L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx

for other solid fuels, because of the requirements of the largest par- When only thermal energy production is pursued, saturated steam
ticles contained (Ruth, 1998). is commonly generated, whereas for electricity production (i.e.
The type and the design of the equipment used to burn the ‘‘power’’) or Combined Heat and Power (CHP) production, super-
waste (i.e. the combustor) are strongly influenced by the fact that heated steam is normally generated. In both these latter cases,
the waste is generally solid and it contains rather big particles the superheated steam is supplied to a steam turbine to make a
(Ruth, 1998). Most of the waste combustors currently used belong Hirn steam cycle (which is the name commonly used for a Rankine
to three main families: mobile/fixed grate, rotary kiln and fluidized cycle with steam superheating). For the production of only power,
bed. a condensing turbine is used (Fig. 3), whereas for CHP, either a back
Castaldi and Themelis (2010) affirm that the technology of pressure turbine (Fig. 4) or an extraction-condensing turbine
mobile grate combustor has reached a high level of development. (Fig. 5) is adopted.
ISWA (2012) data, summarized in Table 2, show how the mobile
grate is the most largely used technology for waste incineration.
The BREF document of the European Commission (2006) states
that mobile grate combustors are the waste incineration equip-
ment offering the largest treatment capacity in terms of thermal
power input per line (up to 120 MW on LHV basis).
Rotary kilns have the major advantage of being able to process
any type of waste – even liquid ones – and they can generally with-
stand higher incineration temperatures than mobile grates can
(1400 vs. 1250 °C – European Commission, 2006), being com-
pletely internally lined by refractory materials. Therefore, they
are particularly suitable for processing those wastes that require
high temperatures and raise severe corrosion concerns, such as
HIW. The BREF document (European Commission, 2006) states that
the maximum treatment capacity of these combustors, in terms of
thermal power input on LHV basis, is equal to 30 MW per line.
In fluidized bed combustors, an upward flowing airstream
keeps in suspension a bed of inert and fuel particles. Efficient heat
transfer and uniform mixing, promoted by the high turbulence,
Fig. 3. Simplified layout of the steam cycle for only electricity production.
enhance combustion. The temperature normally varies between
800 and 900 °C. Van Caneghem et al. (2012), in a complete review
about fluidized bed combustors, report that bubbling, rotating and
circulating fluidized beds have found specific and growing applica-
tions within incineration techniques. However, one of the main
drawbacks of these technologies is related to the requirement of
fuel preparation, especially when dealing with MSW. For this rea-
son, particular types of waste, for instance sludge, available with a
homogenous particle size distribution, are the main candidates to
be burnt in these devices. In the case of fluidized bed, the maxi-
mum treatment capacity per line indicated by the BREF document
(European Commission, 2006) is equal to 90 MW of combustion
power on LHV basis.
Disregarding the type of device, combustion of solid and heter-
ogeneous waste is carried out supplying an amount of air rather in
excess with respect to the stoichiometric requirement, in order to
Fig. 4. Simplified layout of the steam cycle for cogeneration of heat and electricity,
promote the complete oxidation of all the suitable elemental spe- according to back pressure turbine option.
cies. Values of the excess air are typically in the range 40–150%,
leading to a quite high specific production of flue gas (in the range
of 4–10 m3N/kg, depending on the incineration technology, with
higher values for rotary kilns and lower values for grate and fluid-
ized bed combustors – European Commission, 2006), that, in turn,
provides some negative consequences on the energy recovery per-
formances. Indeed, the large flue gas flow rate at the stack entails a
significant energy loss, as well as considerable energy consump-
tions for its treatment.
Energy production is accomplished by partially recovering the
heat content of the combustion products, typically by means of a
steam generator, in most cases integrated with the combustor.

Table 2
Distribution of the combustor type. Elaborated from data reported in ISWA (2012), in
which these types of data were not present for USA plants.

Mobile grate Fluidized bed Rotary kiln


87% 10% 3% Fig. 5. Simplified layout of the steam cycle for cogeneration of heat and electricity,
according to extraction-condensing turbine option.

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx 5

CHP is indicated as the best technique for energy recovery 2.1.1. Small size of WtE plants
(European Commission, 2006). CHP is also found to achieve better The size of WtE plants, in particular when devoted to MSW pro-
performances in Life Cycle Assessment (LCA) evaluations, than only cessing, is mainly determined by the extent of the collection area
electricity production (Damgaard et al., 2010) and, moreover, it is that is served by each plant. The size is generally rather small, also
the technical solution that can warrant high values of the energy considering that small plants may find less social opposition.
recovery index R1 (see Section 5.1 European Union). However, Reimann (2012) classifies as small plants those having a through-
the production of only electricity is the sole possibility in absence put lower than 100,000 Mg/y (about 37.5% of the surveyed ones);
of thermal users (industrial plants or district heating) in the prox- medium plants those in the in the range 100,000–250,000 Mg/y
imity of the WtE plant. Just to give a general idea, the European (about 39.5%); while large plants, those having a throughput larger
average values of energy efficiency were calculated by Reimann than 250,000 Mg/y (about 22.9%). However, for the thermal pro-
(2009) according to the following: cess, it makes more sense to classify the plants on the basis of their
thermal input, which depends both on the waste throughput and
EEp LHV. The LHV of waste is rather low if compared with fossil fuels’
gel ¼ ð1Þ
Ein LHV, being for example for MSW of about 10 GJ/Mg on average
(Reimann, 2012). Thus, WtE thermal input sizes are generally small
ETp – lower than 35 MWLHV- or medium size – lower than 90 MWLHV –
gth ¼ ð2Þ
Ein and in any case definitely much smaller than conventional fossil
fuel plants (1000–2000 MWLHV or even larger). The typical small
EEp;CHP thermal input size has several effects on the WtE performances
gel;CHP ¼ ð3Þ due to several reasons: steam turbine efficiency decreases with
Ein
the size (Consonni and Viganò, 2012); unfavorable volume to sur-
face ratio in the combustion chamber imposes large excess air;
ETp;CHP
gth;CHP ¼ ð4Þ auxiliary devices performances decrease with their size. Addition-
Ein
ally, the plant performances are indirectly affected by the size for
where EEp is weighted average specific produced electricity; Ein is economic reasons. In fact, due to the small size, specific costs of
weighted average specific energy input (including import); ETp is devices are quite high, meaning that, in order to contain the overall
weighted average specific heat produced and used; gel is the electric costs, rather simple configurations should be selected (as better
energy production efficiency of only electricity production WtE explained in the following paragraphs). In these cases, the poten-
plants and is 20.7%, gth is thermal energy production efficiency of tial for technical sophistications, which would be able to increase
facilities with major heat production and is up to 81.3; gel,CHP and the performances, is limited or not feasible by economic reasons
gth,CHP are respectively electric and thermal energy production effi- (Pavlas et al., 2011). On the contrary, in large WtE plants, the
ciency when CHP is applied and are about 14.2% and 45.9% respec- reduction of specific costs allow for using the economic resources
tively, considering that EEp,CHP is weighted average specific for technical sophistication of the plant and performances
produced electricity in cogeneration plants and ETp,CHP is weighted optimization.
average specific heat produced and used in cogeneration plants. Elaborating the data from ISWA (2012), it is possible to calcu-
late the design thermal power input per line, for those plants that
supplied the data about the design values of waste input flow rate
2.1. Power generation and LHV. The percentage distribution of the size – in term of ther-
mal power input – is reported in Table 3.
Focusing on the generation of only electricity (i.e. ‘‘power’’), effi-
ciencies of WtE plants appear quite low when compared with
2.1.2. Conservative steam parameters
those commonly achieved in conventional fossil fuel-fired power
For the Hirn cycle, efficiency increases as the pressure and tem-
plants. The maximum net electric efficiency of WtE plants reported
perature of the superheated steam increase. However, the heat
in the literature is 30% on LHV basis (Murer et al., 2011; Gohlke and
transfer surfaces of WtE boilers must face severe, high tempera-
Martin, 2007). Indeed, in 2007, Graus et al. (2007) determined that
ture, acidic corrosion, caused by both the metal chlorides in the
the international weighted average efficiencies on LHV basis were
fly ash and the high concentration of hydrogen chloride (HCl) in
35% for coal, 45% for natural gas and 38% for oil-fired power gener-
the flue gas (Lee et al., 2007). Also Persson et al. (2007) and De
ation. The highest efficiencies observed for coal was about 42%LHV.
Greef et al. (2013) confirm that chlorine has a key role in the cor-
In the EU, for coal-fired power generation the average efficiency
rosion process. The presence of chlorine in the waste, as well as
increased from 34%LHV in 1990 to 38%LHV in 2005 and it was
those of metals, is unavoidable (e.g. apart from chlorine containing
expected to increase up to 40%LHV by 2015 (Graus and Worrell,
plastics, like PVC, chlorine is a basic constituent of kitchen salt).
2009).
The effectiveness of high temperature acidic corrosion depends
The reasons for relatively low performances, in comparison
mainly on the temperature of the metallic surface. The corrosion
with conventional power plants, are mainly due to the combined
rate increases with temperature, hence, to limit corrosion, the tem-
effects of economic and technical constraints. In particular, the fol-
perature of the surfaces must be limited. This consideration applies
lowing points can be identified:
both to evaporating and superheating surfaces, setting direct limits
both to evaporating pressure (i.e. temperature) and superheating
(i) small size of WtE plants;
temperature.
(ii) conservative steam parameters (i.e. evaporating pressure
and superheating temperature);
(iii) relatively high condensing pressure; Table 3
(iv) simple cycle configuration (no steam reheat, only few water Distribution of the size of lines. Elaborated from data reported in ISWA (2012).
pre-heaters); Pth < 35 MW 35 MW < Pth < 90 MW Pth > 90 MW
(v) high stack loss with no or indirect air pre-heating;
61.4% 36.3% 2.3%
(vi) large rate of in-plant energy consumption.

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
6 L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx

Furthermore, the combination of the limit on the maximum report that grate furnace WtE plants in Europe typically produce
superheating temperature and that on the maximum moisture 546 kWh of electricity per Mg of waste, which corresponds to a
content at the discharge of typical condensing turbines usually sets gross efficiency of 18%, based on waste LHV of 10.44 MJ/Mg (only
an even more stringent limit on the maximum evaporating pres- electricity production). Considering in-plant consumption of an
sure. In fact, without reheating, once the superheating temperature average 150 kWh per Mg of waste, the exported electricity is about
is fixed, to warrant a certain steam quality at the discharge of the 396 kWh/Mg (net efficiency 13%). While most modern plants,
steam turbine, evaporating pressure must not exceed a certain using 40 bar/400 °C, typically produce 640 kWh/Mg of electricity,
level, which depends on the expansion efficiency of the turbine which corresponds to a gross electric efficiency of 22% (LHV of
(i.e. the slope of the expansion line in the enthalpy–entropy dia- 10.44 MJ/Mg). With in-plant consumption of 120 kWh/Mg, this
gram). This situation is exemplified in the thermodynamic diagram ends up in about 520 kWh of electricity exported per Mg of waste
reported in Fig. 6. (net efficiency of 18%).
For the above reasons, steam cycles associated to WtE processes The traditional design of waste-fired steam generators generally
commonly adopt conservative steam parameters, which, with a sets an evaporator as the first heat exchanger in the sequence of
flue gas temperature, at the superheater, not exceeding 650 °C, banks cooling down the flue gas. This solution allows keeping
can prevent high corrosion rates (De Greef et al., 2013). low the temperature of tube surfaces where high heat fluxes are
Fig. 7 shows, on the basis of ISWA data (2012), the dispersion of present. Moreover, this is the most effective way of cooling down
steam temperature and pressure values, for the lines of those flue gas, while recovering energy, to a temperature low
plants that supplied the data. How it can be seen, most boilers enough to limit corrosion on the following banks, especially
adopt temperature values in the neighborhoods of 400 °C. Such a superheaters.
superheating temperature is usually coupled with an evaporating Some evaporator tubes may be installed into the walls of the
pressure of 40 bar. However, in recent years, such a design appears secondary combustion chamber (waterwall) above the grate fur-
more like a baseline, which compares with more advanced nace (Jegla et al., 2010) – or within the riser of fluidized bed com-
solutions. bustor (Van Caneghem et al., 2012) – in order to recovery heat
Gohlke and Martin (2007) report that WtE plants built during exploiting both radiation and convection heat transfer mechanism
the 1980s and early 1990s in the USA used steam parameters of of flue gas. This arrangement also provides adequate cooling of the
60 bar/443 °C and generally were more efficient than plants built flue gases to an acceptable level to enter the downstream super-
in Europe based on 40 bar boilers or plant built in Japan based heater section – without the fear of tube overheating (Jegla et al.,
on 20 bar boilers, during the same time. The same authors also 2010). The excess air required in this type of combustor with inte-
grated heat recovery in walls, is generally lower than old fashion
furnaces, without such an arrangement, thus reducing also the
exhausts sensible heat losses. The available surface for this inte-
grated heat recovery is placed in open chamber that surrounds
the firing equipment (it is generally grate or bed) and extends from
the top of stoker or bed to the exit plane at which the next surface
(usually screen or superheater) starts (Jegla et al., 2010). Exhausts
come into contact with the superheater tubes, when their temper-
ature is lowered down to at least 650 °C.
Specific features and characteristics of heat exchangers used in
WtE plants are described and discussed by Stehlík (2007) and
Stehlík (2011).
Lee et al. (2007) describe several methods of protection of the
boiler tubes from high temperature corrosion, aimed at extending
their lifespan. De Greef et al. (2013) list some primary measures to
reduce high temperature corrosion, among which the Prism, which
Fig. 6. Temperature – entropy diagram for an ideal Hirn cycle, showing that, for a
is a technological boiler feature based on placing a prism-shaped
fixed superheating temperature (Tmax), the increase of evaporating pressure leads to body in the first empty boiler pass, allowing for optimization of
a decrease in steam quality at the discharge of the steam turbine. secondary air injection and combustion control. Secondary mea-
sures, to cope with high temperature corrosion, consist of using
many kinds of corrosion-resistant material systems, among which
Inconel 625 (21Cr–9Mo–3.5Nb–Ni base), that can be overlaid on
waterwall areas and superheater tubes, to protect them from the
attack of HCl/Cl2 (De Greef et al., 2013). This method has been suc-
cessfully used in the waterwall tubes and part of the superheater
tube bundles in many WtE facilities.
Improved boiler design, the use of corrosion resistant material
cladding of the waterwall tubes and adaptation of the water-steam
cycle, open the possibility to increase the superheated steam tem-
perature and, hence, conveniently the pressure. Recently built WtE
plants operate with increased values of pressure and temperature
up to, respectively, 60 bar and 500 °C. In some cases, even higher
pressure levels may be achieved, as shown later on.
High steam parameters are the results of a compromise
between investments, expected decreases of plant availability
and profit related to efficiency increase and electricity sale
Fig. 7. Dispersion of steam temperature and pressure values. Elaborated from ISWA (Pavlas et al., 2011). The increased investment and operating costs
(2012). can be withstood only for large throughput WtE plants.

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx 7

2.1.3. Relatively high condensing pressure 2.1.5. High stack loss with no or indirect air pre-heating
Hirn cycle efficiency is improved also for lowering pressure at Stehlík (2012) reports that exhausts are cooled down in the heat
the condenser. Condenser pressure is in turn influenced by cooling recovery process to 150–380 °C and temperatures exceeding
medium, its temperature and plant capacity. Pressure for air cooled 200 °C are quite common. The temperature at the exit of steam
condensers is significantly related to temperature of surrounding generator is a trade-off between maximum boiler efficiency (corre-
air. Temperature of 10 °C enables pressure of 0.1–0.08 bar, temper- sponding to 150 °C) and applied type of flue gas treatment (FGT),
ature of 20 °C has a negative impact increasing the condensing since different types of processes require different temperature
pressure to 0.17–0.12 bar (European Commission, 2006). Con- levels for proper operation (De Greef et al., 2013).
denser pressure may be lowered by using water cooled condensers Boiler performance is, hence, influenced by exiting exhausts
in a closed loop with atmospheric cooling tower (Pavlas et al., temperature and amount, i.e. thermal losses by sensible heat. Also
2011). feedstock moisture influences the stack losses. However, when the
However in traditional WtE plants air cooled condensers are LHV of waste is considered for losses and boiler efficiency calcula-
quite often used, because less expensive than water cooled con- tion, in place of HHV, the effect of feedstock moisture is already
densers. For large installations, larger surfaces of air cooled con- accounted for. These energy losses are the most important contri-
densers or the use of water cooled condensers may lower the bution to overall process losses ranging from 7% up to 25% of waste
condensing pressure and hence improving the efficiency. Typical energy input according to the exhaust temperature and air excess;
values are not far from 0.15 bar (Gohlke, 2009). while other losses in the overall process (thermal losses by radia-
Barigozzi et al. (2011) study a particular configuration of WtE tion and convection, chemical losses by incomplete combustion,
cogeneration plant equipped with an air cooled condenser in par- thermal losses in unburned fuel) are less important (about 3–4%)
allel with a water cooled condenser. With the aim of increasing (Pavlas et al., 2011). Boiler efficiency – calculated as the ratio of
the power output, hence taking into consideration the energy energy acquired by the steam to the thermal energy entering with
requirements for the two types of condensers, they demonstrate the waste – ranges between 75% and 85%, also depending by foul-
that the air cooled is the best way to reject heat if the ambient tem- ing of heat-exchanging surfaces (Tabasová et al., 2012). According
perature is lower than 15 °C. At higher ambient temperature, the to Pavlas et al. (2011), boiler efficiency reaches 81%, if boiler exiting
condensation should exploit the cooling capacity of the water temperature of 250 °C and 6% oxygen content after the last air sup-
cooled condenser, as much as possible while discarding the ply are assumed: significant increase of efficiency is depending on
remaining heat in the air cooled condenser. decreasing the flue gas temperature, which on the other hand puts
more requirements on FGT system.
2.1.4. Simple cycle configuration One possibility to reduce the stack losses is given by the reduc-
In fossil fuel steam power plants, particular arrangements are tion of flue gas flow rate. This can be achieved by reducing the
also applied aimed at increasing the efficiency: as internal regener- excess air from 1.75 to 1.9 (Gohlke, 2009; Main and Maghon,
ation and reheating (Cengel and Boles, 1998). In particular reheat- 2010) to lower values, as 1.39 (Main and Maghon, 2010), through
ing has the feature of allowing working with quite high pressure an improved combustor design with heat recovery integration (as
level, even if superheating temperature is constrained (Fig. 8). previously discussed).
Indeed, steam at an intermediate expansion pressure is heated Additionally thermal losses by sensible heat at the boiler exit
again with the consequence of drifting away the expansion curve can be reduced by Flue Gas Recirculation (FGR). The flow rate of
from the saturation line, and avoiding excessive steam condensing the re-circulated flue gas is used to control combustor tempera-
in the final turbine stage, which may cause erosion of blades by ture, while the fresh air flow rate is determined in order to settle
liquid droplets. the oxygen content of the exhausts, leaving the combustion cham-
Another possibility is the superheating of live steam, from con- ber, as close as possible to 6%, that is the reference value to ensure
ventional temperature to higher ones, in external superheaters completion of the combustion reactions, and determines the corre-
using oil, gas or biomass (Main and Maghon, 2010). sponding minimum flue gas flow rate. Liuzzo et al. (2007) estimate
These type of measures can be applied in WtE plants (Stehlík, net electric efficiency increase from 1% to 3% points, when FGR is
2012; Gohlke and Martin, 2007) to significantly increase the net applied. Additionally, when FGR is implemented, lower capital
electrical efficiency, in case of high capacity installations, while costs are associated with the smaller installed capacity of the
for facilities with low processing capacity (100,000–150,000 Mg/ FGT. Castaldi and Themelis (2010) highlight that FGR enhances
y), the potential for afore mentioned measures is limited and/or the combustion performance increasing the temperature in the
financially not feasible (Pavlas et al., 2011). reaction zone and reducing the emissions level, allowing more
completely combustion in a short time without adding any high-
grade fuel. However, they also add that FGR supplies complexity
and robustness issues to the system.
FGR is proposed also in association to oxygen enrichment of
combustion air to improve combustion performances (Gohlke
and Busch, 2001; Gohlke and Martin, 2007), even if additional con-
sumption for oxygen separation from air should be considered.
Tsiliyannis (2013) shows that, besides NOx emission control,
suitable manipulation of FGR enhances WtE performances under
waste uncertainty, enabling higher throughput, at the desired tem-
perature and within the allowed FGR residence time range.
In reference to the issues related to steam generator exit tem-
perature and FGT consumption, Poggio and Grieco (2010) investi-
gate the variation in energy efficiency of four different WtE gas
neutralization technologies. Actually the influence is rather lim-
ited, since the efficiency of the most advantageous technology,
Fig. 8. Temperature – entropy diagram showing reheating effect for an ideal which resulted the dry treatment with sodium bicarbonate
Hirn cycle. (NaHCO3), was 24%, while the efficiency of the least advantageous

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
8 L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx

technology, which resulted semi-dry treatment with calcium tricity efficiency is 27%, net efficiency is 24%. Line n. 3, started-up in
hydroxide (Ca(OH)2), was 23.7%. 2004, is fed by waste biomass and sewage sludge (total amount of
Additional measures to improve energy recovery and conver- feeding in 2010 was 257,599 Mg2) and operates with steam param-
sion efficiency are preheating of air or water, using low tempera- eters 73 bar/480 °C; gross electricity efficiency is 28%, net efficiency
ture streams available within the plant, in order to improve the is 25%. The three lines are equipped with a reverse-acting grate.
heat recovery system effectiveness (Stehlík, 2012). Combustion Main and Maghon (2010) report different measures – actually
air preheating, for instance, may be beneficial for the enthalpy bal- realized in some existing plants and described below – for optimiz-
ance in the combustion chamber. One low temperature stream, ing boiler and combustion conditions with respect to conventional
which may be used for air preheating, is the water used from grate WtE, for which they calculate, as reference values, a boiler effi-
cooling circuit (grate cooling is generally used by some manufac- ciency of 86.5% and 26.35% gross power efficiency (at the following
turers when the LHV of waste is higher than 12.5–14.5 GJ/Mg), conditions: 40 bar; 400 °C; 60% excess air; flue gas temperature at
characterized by a temperature of 70–80 °C. Air or water pre-heat- the boiler exit: 190 °C; LHV 11 GJ/Mg; condenser pressure:
ing may be realized also by further heat recovery from exhausts at 0.1 bar).
the exit of the steam generator. A new stoker and boiler unit was installed in the year 2006 at
the WtE plant in Hameln (Germany) (40 MWLHV thermal input,
2.1.6. Large rate of in-plant energy consumption waste LHV 12 GJ/Mg, steam parameters 41 bar and 400 °C). This
The in-plant consumption represents a quite important fraction modification allowed reducing the excess air from 1.9 to 1.39,
of the gross power, with a not negligible contribution coming from increasing the boiler efficiency to 87.65% and gross power produc-
the FGT. Gohlke and Martin (2007) report values of in-plant con- tion to 26.63% (Main and Maghon, 2010).
sumption from 120 to 150 kWh per Mg of entering waste. At the WtE plant in Arhus (Denmark), in operation since the
Damgaard et al. (2010) assume electricity consumption for FGT beginning of 2005 (thermal input 51 MWLHV; waste LHV 10.5 GJ/
from 10 up to 80 kWh per Mg of waste, depending on the type of Mg; steam parameters 42 bar, 400 °C), a modification was imple-
applied technologies. For example, a FGT based on semi-dry flue mented in order to lower the boiler outlet temperature to 100 °C.
gas cleaning, activated carbon injection for removal of dioxin, fol- Additional flue gas cooling was achieved by adding an external
lowed by a bag house filter and selective non-catalytic reduction economizer, to reach 140 °C, and a flue gas heat exchanger to fur-
(SNCR) consumes about 45 kWh/Mg. If flue gases are condensed ther cool to 100 °C. The heat recovered from the last heat exchan-
and selective catalytic reduction (SCR) is used in place of SNCR, ger is used, together with grate cooling heat, for air pre-heating.
the consumption becomes 75 kWh/Mg. In case of wet flue gas The boiler efficiency increased to 92.63% and gross power produc-
cleaning, followed by a bag house filter with activated carbon tion to 28.14% (Main and Maghon, 2010).
injection and SNCR, the consumption is 70 kWh/Mg. If flue gas con- Superheating of live steam from 400 °C to 520 °C in an external
densation is added and SCR is substituted to SNCR, the consump- superheater, consisting of natural gas fired boiler, was realized in a
tion becomes 80 kWh/Mg. new WtE plant in Heringen (Germany) (two units with 58 MWLHV
Depending on the plant size the overall fraction of the electrical thermal input; waste LHV 12 GJ/Mg; steam parameters 81 bar,
power for in-plant consumption is 10–15%, which may increase up 520 °C). The external superheater variant, at similar conditions as
to 21% when waste pre-treatment consumption is required to feed the reference case, has a boiler efficiency of 87.65% and a gross
fluidized bed (Nixon et al., 2013b). power efficiency of 29.68% (Main and Maghon, 2010).
In the new WtE plant in Naples (Italy), in operation since begin-
2.1.7. Advanced solutions to increase electric energy performances ning 2009 (three lines with 113.3 MWLHV thermal input each;
An example of modern WtE plant with very high steam param- waste LHV 15.1 MJ/Mg), the boilers are designed for high steam
eters, is the 530,000 Mg/y, 10 GJ/Mg LHV, Waste Fired Power Plant parameters of 500 °C at 90 bar. The superheaters are located in a
(WFPP) in Amsterdam (NL), started-up in 2007, equipped with a flue gas temperature zone above 800 °C, requiring to be protected
horizontal grate, using high steam parameters of 130 bar and by monolithic SiC concrete, in order to increase their lifetime. For
440 °C and steam-steam reheating, to achieve a peak 30% net elec- these higher steam parameters there is no change in boiler effi-
trical efficiency (34.5% gross) and an availability of 92.4% (Murer ciency of 86.5%, but gross power efficiency reaches 30.2% (Main
et al., 2011). As explained by Murer et al. (2011), the advantage and Maghon, 2010).
of this concept is the high energy efficiency due the high pressure In the Ruedersdorf-Berlin (Germany) WtE plant (110 MWLHV
and the reheating, combined with quite conventional superheater thermal input; LHV 14.5 GJ/Mg; maximum steam pressure and
temperature of 440 °C. In the intermediate reheater, saturated temperature 90 bar and 420 °C), in operation since middle 2008,
steam from the drum reheats the steam exiting from the first stage the intermediate reheat concept was applied. In this case steam
of the turbine (14 bar) to 320 °C. Condenser pressure is kept to at intermediate expansion pressure – 24.4 bars – is reheated at
0.03 bar, by water cooling (in this case sea water). The waterwalls 420 °C. Also in this case there is no change in boiler efficiency of
in the furnace are protected by Inconel (note that steam saturation 86.5%, but gross power efficiency increases to 29.9% (Main and
temperature is increased with respect to conventional values due Maghon, 2010).
to the increased pressure), while the superheater is made of carbon
steel without any particular coating. The same authors report no
pipe changes after 30,000 operational hours per boiler. 2.2. Cogeneration
Another example of modern WtE plant is reported by Gohlke
and Martin (2007), in reference to the Brescia (IT) one. Lines no. A further means to improve the overall energy recovery of WtE
1 and 2, started up in 1998, are fed by MSW, sewage sludge and – and thus the greenhouse gas balance (Gohlke, 2009) – is CHP.
waste biomass (total amount of feeding in 2010 was 551,728 Heat can be supplied to a nearby industrial plant requiring heat
Mg2), and operate with steam parameters 61 bar/450 °C; gross elec- for its own process: in this case a back pressure turbine (i.e. pres-
sure at the turbine exit is usually high to provide the temperature
2
necessary for heat utilization) may be used to supply a constant –
Data from operation report of the Brescia plant, in Italian (Comune di Brescia
Settore Ambiente ed Ecologia. Rapporto dell’Osservatorio sul funzionamento del
over the year – heat rate to the industrial user.
Termoutilizzatore di Brescia relativo agli anni 2008/2009/2010) available at http:// Indeed, the technical design of the plant imposes a fixed ratio
www.a2a.eu/. between electricity and heat production: electricity production is

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx 9

only possible when a heat demand exists, since the return fluid 2.3. MSW pre-treatment and energy recovery
from the heat user is the only means for cooling the steam.
When heat is cogenerated for feeding a district heating net- As anticipated above, MSW can be directly fed to WtE, or may
work, generally the preferred technical solution is a condensing be pre-processed in MBT plants, according to material stream sep-
turbine with intermediate steam extraction, which allows for quite aration concept, consisting mainly in metal recovery and separa-
flexible regulation according to the variable heat demand through tion of generally three streams: inert, humid fraction and
the seasons. Indeed, these plants can operate in condensation combustible fraction. Separation is carried out by mechanical
mode, thereby only producing electricity and cooling off the excess equipment which is able to select different materials but, obvi-
heat, or in extraction mode for cogeneration of heat and electricity. ously, not in pure quality. Indeed, the humid fraction is contami-
In this last case, steam is extracted before the last turbine stage for nated with undesired materials, thus its fate is generally
supplying heat to the thermal user, thereby reducing the electricity landfilling after biostabilization. An alternative route for MSW
production. If the demand for heat is large, most plants are capable pre-treatment is biological drying, in which aerobic biostabiliza-
of entirely shutting off the valves to the last turbine stage, increas- tion takes place at a reduced extent, with the aim of exploiting
ing the heat production even further (operating in back pressure the biological reaction heat of the biodegradable matter to partially
mode). During demand cut-off (summer operation) the most of dry the whole MSW.
steam is utilized for electricity production without having to The combustible stream, still containing some inert and humid
reduce the throughput of the incineration plant; thermal output parts, is characterized by an increased LHV. When the combustible
is reduced and electricity output is increased. Excess low grade stream complies with existing technical standards (for example, in
heat is rejected by the traditional condenser (Pavlas et al., 2011). EU, CEN/TS 15359, 2006) and/or some national laws can be classi-
Pavlas et al. (2011) calculate how the specific electricity produc- fied as RDF/SRF and may find some specific destinations, additional
tion changes in cogeneration mode vs. the ratio of steam through to devoted WtE, as cement kiln plants or thermal power plants,
condensation stage to total flow rate. Steam parameters are ini- substituting solid fossil fuels (i.e. coal).
tially assumed 40 bar/400 °C and two different possibilities for However, one should consider that MBT have their own con-
heat cogeneration are considered: direct steam export at 11 bar sumption – reported in a wide range from 8 kWh/Mg (Bovea
and hot water production by steam at 3 bar. When turbine is oper- et al., 2010) to 80 kWh/Mg (Papageorgiou et al., 2009) – and that
ated in back-pressure mode (i.e. no steam flows through the con- some losses of combustible materials take place in the mechanical
densing stage), specific electricity production ranges within 200– separation. With this in mind, several studies (Consonni et al.,
400 kWh/Mg depending on value of back-pressure (corresponding 2005; Papageorgiou et al., 2009; Cimpan and Wenzel, 2013) were
to net electric efficiency of about 4–10%). If superheated steam carried out aimed at assessing the effective energy balance of alter-
pressure is increased up to 60 bar, specific electricity production native routes for feeding WtE, starting from MSW.
increases by 100 kWh/Mg (corresponding net electric efficiency Consonni et al. (2005) assess four strategies for energy recovery
about 14%). Arrangement in full condensation mode leads to gen- from MSW, downstream of the separate collection, by dedicated
eration of 600–800 kWh/Mg (corresponding calculated net electric WtE plants generating only electricity. They compare direct feed-
efficiency about 20–21%). ing to a grate combustor; light mechanical pre-treatment and feed-
Additional thermal users, besides traditional industrial pro- ing to grate combustor; RDF production and incineration in a
cesses that require heat and district heating networks, are plants fluidized bed combustor. The best energy balance corresponds to
for desalination of sea water, which may play a favorable role in the direct WtE case.
specific locations with arid climate and shortage of water. Dajnak Cimpan and Wenzel (2013) calculate primary energy saving
and Lockwood (2000) make an initial assessment of this concept potential comparing several configurations including mechanical
and concluded that useful quantities of fresh water can be pro- and MBT pre-treatments, which produce SRF/RDF, biogas and/or
duced. Lo Mastro and Mistretta (2006) perform an economic anal- recover additional materials for recycling, and a system based on
ysis of a hypothetical MSW WtE plant coupled with seawater conventional mass burn WtE. They conclude that direct WtE
multi-stage flash desalination plant. achieves the highest savings in scenarios with CHP production;
Fruergaard et al. (2010) highlighted that even if district heating MBT-based systems have similarly high performance if SRF
is mainly relevant in temperate climates, the utilization of surplus streams are co-combusted with coal.
heat from the WtE may be done also in warmer climates e.g. by When RDF/SRF is only used in devoted WtE plants, MBT-based
means of district cooling (in which a temperature difference is systems score lower savings due to inherent system losses and
used to drive a cooling device). additional energy costs.
An example of this approach is given by the WtE plant in Goth- Hence, from an energy recovery point of view, the pre-treat-
enburg (SE), which is one of the many heat suppliers to the town ment of MSW seems not to be advantageous, unless the efforts
large district heating network, where during the summer the heat to produce an improved quality fuel are conveyed to partially sub-
demand is further increased by absorption chillers, spread around stitute fossil fuels, as coal in thermal power plants or cement kilns.
the city (Gohlke, 2009). The feed and return temperature of the dis- Additionally, it should be pinpointed that with the increasing of
trict heating depends mainly on the ambient temperature, indeed, the up-stream separate collection, the LHV of residual MSW should
the feed temperature ranges from 75 to 110 °C, whereas the return be naturally increased (Calabrò, 2010).
temperature lies between 40 and 55 °C. The yearly averages are 80 However, when selecting a waste management strategy, several
and 45 °C (Gohlke, 2009). aspects should be taken into account. In those cases in which the
Hence, expanding cogeneration towards trigeneration can priority is not the energy recovery, MBT application allows for
augment the energy supply for summer months in Europe or reducing the amount of landfilled waste and recovering some types
for year-round cooling in tropical locations, as affirmed by of materials.
Udomsri et al. (2011), who show the benefits that can be
obtained for a MSW WtE plant integrated with absorption tech- 2.4. Integration of WtE and gas turbine combined cycles
nology in Southeast Asia: for a plant with 1350 Mg/day of MSW
input, the system can produce 77 MW of cooling and 21.5 MW When heat recovery cannot be applied and electricity perfor-
of electricity. mances cannot be improved according to the technical possibilities

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
10 L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx

described above, unconventional solutions may be explored 2008). For MSW, a pre-treatment is generally required, especially
(Pavlas et al., 2011). One of the many discussed concepts is the when fluidized bed are used, even if some gasification technologies
integration of municipal waste incinerator with combined steam- are able to accept no-pretreated MSW (Tanigaki et al., 2012). If
gas cycle (Otoma et al., 1997; Korobitsyn et al., 1999; Consonni MSW pre-treatment is necessary, one should consider that MBT
and Silva 2007; Qiu and Hayden, 2009; Poma et al., 2010; processes are applied with the additional energy consumptions,
Udomsri et al., 2010; Balcazar et al., 2013). The main idea is based as highlighted before, that should be included, also in this case,
on the possibility of superheating the steam produced by the WtE in the overall energy balance.
flue gas boiler using cleaner exhausts, as for example gas turbine The exiting syngas contains large amounts of not completely
(GT) exhausts, in order to allow raising the superheated steam oxidized products, mainly carbon monoxide, hydrogen and lower
temperature, without the above mentioned corrosion risks. amounts of methane, hence carrying an energy content that can
Qiu and Hayden (2009) propose two different configurations for be used in some type of energy conversion equipment to generate
the integration of the WtE and natural gas turbine combined cycle an energy carrier. Generally the main problems for using the syn-
(GTCC), showing electrical efficiency values in the range of 30–42%, gas, and in particular the syngas from waste, are related to the
depending on the ratio between the natural gas input and the total presence of undesired compounds such as particulate, tar, alkali
energy input (from 0.1 to 0.6), and on the technical assumptions metals, chloride and sulfide. In particular tars cause problems of
for the two cycles. Poma et al. (2010) calculate a net electric effi- fouling in transport pipeline routes and moving machinery parts,
ciency of 46.3% for an integrated WtE/GT combined cycle, where making necessary the cooling down of syngas and the use of clean-
GT exhausts are mainly used to superheat the steam in the super- ing equipment in order to remove its tarry products and particulate
heater and heat the water in the economizers, while the produc- solid matter (Bebar et al., 2005).
tion of saturated steam is carried out almost completely in the Focusing the attention on the possibility of recovery energy
evaporator of the WtE part. from the gasification of waste, the principal interest is about the
A different way of coupling waste and GT, proposed by de downstream use of the produced syngas. When syngas is com-
Souza-Santos and Ceribeli (2013), is based on MSW incineration busted into a boiler, the problems caused by cooling down the
in a bubbling fluidized bed boiler (at 40 bar), producing flue gases products and condensation of their tarry components are elimi-
that, after cleaning, are injected into gas turbines, and steam to nated and no particular limitations are imposed on the content
feed a steam turbine. For this configuration the authors calculate of tar, dust, alkalis, heavy metals and hydrogen sulfide (Arena,
a net electric efficiency of 33.4%. 2012). In this case the plant is very similar to an incineration plant,
with the difference that oxidation is divided into two steps (two-
step oxidation). If potential improvement in energy efficiency is
3. Gasification of waste with energy recovery sought, it would be more interesting to use the syngas in an inter-
nally-fired cycle, generally characterized by higher conversion effi-
Gasification is a process aimed at converting a solid fuel – solid ciencies than steam cycles, a synthesis process or a system to
waste in the specific case – to a gaseous fuel, called ‘‘producer gas’’ generate hydrogen. In these cases, the challenge is the appropriate
or ‘‘syngas’’, through a partial oxidation of the solid fuel in presence cleaning of the syngas.
of an oxidant amount lower than that required for the stoichiom-
etric combustion. A solid product is generated, too, containing char
(carbonaceous compounds) and ashes. 3.1. Gasification: syngas combustion in a boiler
The gasification of solid waste is a complex process based on
several physical and chemical interactions and generally takes Several authors claim that processes based on waste gasifica-
place at temperatures higher than 600 °C, this last depending tion and syngas combustion in a boiler, however, have some ben-
mainly on the type of reactor and the waste characteristics. The eficial aspects as the combustion of syngas, being a homogenous
different process steps, the operating and performance process gas-phase one, can be performed in more favorable conditions than
parameters are synthesized by Arena (2012). in the case of MSW (Consonni and Viganò, 2012), this means that
The heat required for the gas forming process can be supplied lower amount of air may be supplied and consequent smaller
by partial combustion of the entering fuel (autothermal gasifica- equipment for heat recovery and FGT may be used, because of a
tion) or by an external source (allo-thermal gasification). The substantially lower volume of final flue gas (Bebar et al., 2005).
LHV of the syngas depends strongly on the used gasification med- However, it should be reminded that plants tend to be more com-
ium. When air, oxygen-enriched air or pure oxygen are used, par- plex and costly, more difficult to operate and maintain, less reliable
tial oxidation takes place, the LHV of the syngas ranges from 4– (Consonni and Viganò, 2012).
7 MJ/m3N, when air is used, to 10–15 MJ/m3N, if pure oxygen is used. For instance, operational problems led to the shutdown (since
When steam is used as the only gasification medium, a high hydro- July 2001, after three periods of operation during five years) of
gen concentrated syngas is produced with a LHV of 15–20 MJ/m3N; the gasification plant in Greve in Chianti (Italy), one of the first
in this case no exothermic reaction takes place, requiring an exter- examples of these plants in Italy, commissioned in 1992 for pro-
nal heat source as, for example, thermal plasma (Arena, 2012). cessing 200 Mg/day of RDF in pellet form and based on two atmo-
Several types of gasification reactors are used: fixed bed (up- spheric pressure circulating fluidized bed, working with air at
draft and down-draft), fluidized bed (bubbling fluidized bed, circu- 850 °C, each of 15 MW thermal power input (Morris and
lating fluidized bed, and internally circulated fluidized bed), Waldheim, 1998; Belgiorno et al., 2003).
entrained bed, vertical shaft, moving grate furnace, rotary kiln An example of gasification technology that collected a large
and plasma reactor. Their description can be found in Arena enthusiasm in the past is the Thermoselect one. The Thermoselect
(2012), with a list of selected companies of gasification-based process consists of a degassing channel (where in an air-free envi-
plants, which offer a commercially proven process. ronment, water is evaporated and organic compounds are partially
A variety of wastes may be fed to gasification processes: ASR degasified, by means of external heating), followed by oxygen, high
(Viganò et al., 2010), RDF/SRF (Lombardi et al., 2012; Arena and temperature gasification and melting stage (Calaminus and
Di Gregorio, 2014); mixed plastic waste (Arena et al., 2011); pack- Stahlberg, 1998). The syngas cleaning system generally consists
aging derived fuel (Di Gregorio and Zaccariello, 2012); paper of a quencher, a gas scrubber, a desulfurization stage, a gas drying
industry waste (Ouadi et al., 2013); sewage sludge (Groß et al., system and an activated carbon filter (Richers et al., 1999).

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx 11

Development of this process started in 1989. A demonstration culated net efficiency is about 13%, with about 20% of gross power
plant with a licensed capacity of 4.2 Mg/h (33,000 Mg/y) was taken consumed by Air Separation Unit for pure oxygen production.
into operation in 1992 in Fondotoce (Italy) until 1998, for which Obviously, when gasification with syngas combustion in a boi-
Malkow (2004) reports an electricity yield of about 200– ler is applied, the discussion about how to improve the electric effi-
500 kWh/Mg, using 12 GJ/Mg LHV waste. Then a plant was built ciency of the steam cycle is analogous to the one made previously
in Karlsruhe (DE), for a nominal capacity of 225,000 Mg/y of in reference to waste incineration and, similarly, the steam cycle
MSW and industrial waste (three lines, 10 Mg/h each), 12 GJ/Mg can be configured in order to operate in cogeneration mode, with
LHV, to produce syngas further burnt in two boilers, which produce the same beneficial effects highlighted before.
steam for a back-pressure turbine, with designed power output of As an example, the improved steam parameters, 121 bar/540 °C,
12.5 MW and 50 MW of heat for district heating (Calaminus and at the Metso waste gasification plant (a circulating fluidized bed
Stahlberg, 1998). From the design values, the electric efficiency operating at 850 °C in) in Lahti, Finland, can be used without high
results about 12.5% and the cogeneration efficiency about 62.5%. temperature corrosion risk, thanks to a rather demanding treat-
The plant started official operation in January 2002 (after two years ment of the syngas, based on its cooling to 400 °C and hot gas fil-
of commissioning) and was closed in November 2004. tration by ceramic filters, allowing reaching a net electric
There are some plants in Japan based on Thermoselect technol- efficiency of about 31% (Ligthart, 2013). The plant processes
ogy (originally licensed to the former Kawasaki Steel, now JFE 250,000 Mg of SRF per year and produces 50 MW of electricity
Group), the most famous of which is in Chiba (Japan) (Yamada and 90 MW of district heat.
et al., 2004): its construction started in 1998, for a nominal size Detailed information about waste gasification plants operation
of 300 Mg/day (two lines). In 1999 the plant completed a demon- and performances at commercial scale, are not commonly
strating operation of MSW treatment for a continuous period of reported, mainly because many gasification technologies are at
93 days and a total of more than 130 days, for about 15,000 Mg development or pilot stage. Moreover, a large number of studies
of MSW with 8.5 GJ/Mg LHV. In 2000, the plant began an industrial are focused mainly on the gasification process itself of different
waste treatment/fuel production business which treats industrial inputs, while few of them investigate the downstream use of syn-
wastes, producing a purified syngas which is transferred to the gas in details and the integration between the gasification section
nearby steel works, supplying part of the fuel for the combined- and the power production one.
cycle power plant, together with other byproduct gases generated Tanigaki et al. (2012, 2013) describe the operation of two plants
in the steel works (blast furnace gas, coke oven gas, etc.). To dem- based on the direct melting system (DMS), which is a technology
onstrate the possibility of using the syngas also in other devices, a quite widespread in Japan and South Korea, based on shaft-furnace
1.5 MW gas engine generator was installed at the site of the Chiba type gasification and melting process, classified as an atmospheric
plant, and a demonstration test of gas engine power generation moving bed gasifier. In these plants co-combustion of MSW (LHV
was performed, at full and partial load, using part of the fuel gas 6.8–9.1 GJ/Mg) and coke is applied, also with the aim of promoting
supplied to the steel work. the slag fluidity; very high temperatures are reached in the bottom
Consonni and Viganò (2012) describe and simulate two power of the gasification reactor where combustion and melting occur
plants based on gasification and syngas combustion in a boiler to (see Section 5.3 Japan). The power section is based on a steam
produce steam for feeding a Hirn cycle, reproducing as much as cycle, with 39.2 bar/400 °C steam parameters. Even if not directly
possible two existing gasification commercial technologies: a high comparable with plants fueled only by waste, it is worth to report
temperature grate gasifier resembling the technology proposed by the gross electric efficiency shown by Tanigaki et al. (2012), which
Ener-G; and a low temperature fluidized-bed gasifier, based on the is 18.9% and 23.0%, for the two existing plants.
technology proposed by the Japanese company Ebara. Assuming a
waste LHV equal to 10.34 GJ/Mg and steam parameters at 40 bar/
400 °C, for a plant size of 50 MW of thermal power input, the gross 3.2. Gasification: syngas use in internally-fired devices
efficiency is about 23.7–23.8%, while the net efficiency is about
19.6–19.7% for the two cases. If steam parameters are increased Alternatively with respect to what illustrated in the previous
up to 70 bar/450 °C, in the case of larger plant with 300 MW input paragraph, the syngas may be fed, after an adequate treatment,
thermal power, the gross efficiency may reach 31%, with a corre- to a highly efficient internally-fired cycles, as GT and GTCC or inter-
sponding net efficiency of about 27%. The authors conclude that nal combustion/reciprocating engines (ICE). The use of a gaseous
the energy performance is basically determined by the configura- fuel in highly efficient devices opens potentially the possibility of
tion and the operating parameters of the steam cycle – as it hap- increasing the overall energy efficiency of the process. However,
pens in conventional incineration plants – and by auxiliary as illustrated by some examples reported in this paragraph several
power consumption. constraints – as the necessity of demanding syngas treatment, the
Viganò et al. (2010) calculate a gross electric efficiency of 25.7% atmospheric pressure gasification and the rather low efficiency of
which, after subtracting internal consumptions, results in a net GT based systems at the small scale typical of waste treatment
electric efficiency of 22.5%, in the case of sequential gasification plants (Consonni and Viganò, 2012) – reduce the effective
and syngas combustion, using ASR with 19.2 GJ/Mg LHV, for a ther- improvements achievable by this configuration, reporting the effi-
mal power input of 100 MW (about 140,000 Mg/y) and considering ciency to values similar to those previously seen for incineration or
a steam cycle parameters equal to 30 bar/325 °C. gasification with syngas combustion in a boiler.
Arena et al. (2011) calculate a net electric efficiency of 23.7 for a In particular, when syngas is fed to GT the following limitations
small gasification plant fed by two different types of mixed plastic on pollutants content apply: tar: 10 mg/m3N, dust: 5 mg/m3N, alkalis:
waste (LHV 31.7 GJ/Mg and 40.2 GJ/Mg; entering waste flow rate 0.1 ppm in weight, heavy metals: 0.1 ppm in weight, hydrogen sul-
and 1500 kg/h and 1900 kg/h). fide: 20 ppm in volume; for ICE the limits are the following: tar:
Lombardi et al. (2012) simulate a power plant based on high 100 mg/m3N, dust: 50 mg/m3N, alkalis: 0.1 ppm in weight, heavy
temperature gasification, in pure oxygen, of RDF (about 64 MW metals: 0.1 ppm in weight, hydrogen sulfide: 20 ppm in volume
of thermal power input), with ash melting, coupled with a steam (Arena, 2012). The use of syngas in highly efficient internally-fired
cycle, characterized by steam parameters 56 bar/405 °C, resem- cycles is investigated mainly by simulations aimed at evaluating
bling an existing Italian plant (Zagaroli, 2007). In this case, the cal- the potentially achievable energy production, quite often

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
12 L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx

disregarding the actual syngas pre-treatment required to obtain an The initial approach to apply thermal plasma to waste treat-
adequate quality of syngas for feeding such devices. ment was based on throwing a plasma jet on the solid waste, in
Yassin et al. (2009), consider the gasification of RDF (LHV order to produce their rapid heating and releasing and cracking
16.7 GJ/Mg), performing a mass and energy balance in the case of of volatile matter. This type of approach is described in Zhang
fluidized bed gasification coupled with ICE or GTCC. The authors et al. (2012, 2013). Minutillo et al. (2009) developed a thermo-
show that when a GTCC is implemented, the net electric efficiency chemical model to estimate the syngas composition and the energy
is in the range 24–27%, while in case of ICE application it is in the required for the gasification reactions of RDF (12.9 GJ/Mg, about
23–25% range (lower values for 50,000 Mg/y throughput, i.e. 70 MW thermal power input with waste). The syngas is further
29.4 MW thermal power input; higher values for 100,000 Mg/y supplied to a GTCC producing electric energy with a net efficiency
throughput, i.e. 58.8 MW thermal power input). of 31%. Galeno et al. (2011) simulated the use of the syngas from
Also Lombardi et al. (2012) evaluate by simulation the energy RDF (12.9 GJ/Mg, 12.9 MW thermal power input with waste)
efficiency of using syngas from RDF high temperature gasification plasma gasification to a solid oxyde fuel cell (SOFC). In this case
(64 MW of thermal power input), in pure oxygen with ash melting, the net electric efficiency – after subtracting torch and air separa-
in a GTCC. In this case final net electric efficiency results 14.7%, tion consumptions – reaches 32.7%.
mainly because the internal consumptions are about 69% of the With the aim of limiting the electric power demand required by
gross power output, namely due to the power required for air com- the plasma gasification of the whole solid waste, a second
pression (about 52%). The same authors show that one possibility approach was proposed, based on the use of plasma in a two-stage
for improving the overall efficiency is to increase the integration thermal process: waste is first traditionally gasified, then plasma
between the gasification section and the power section, recovering jet is used to refine the syngas in order to decompose and convert
the heat from syngas cooling before the cleaning process and using tar, to improve its quality for engine feeding. Eventually, the
it to produce additional steam for feeding the bottoming steam plasma jet may be used also to vitrify the gasification solid
cycle (net electric efficiency 17.1%) or for injection in the GT com- residues.
bustion chamber (net electric efficiency 18.7%). An example of this approach is described by Materazzi et al.
Additionally, it should be pinpointed that gasification at pres- (2013) and Taylor et al. (2013). Ray et al. (2012) claim that the cal-
sure higher than atmospheric one may enhance the opportunities culated net electric efficiency for a commercial scale plant based on
to increase energy conversion efficiency in GT based cycles and this approach, using a feeding waste stream with LHV 21–22 GJ/
reduce costs, but a process of this kind would be very challenging Mg, is in excess of 25%.
and it is not yet provided by any company (Consonni and Viganò, Lombardi et al. (2012) calculate a net electric efficiency of 25.3%
2012). for a system based on traditional gasification of RDF (LHV 18 GJ/
As a specific case, Giugliano et al. (2008) propose the gasifica- Mg, about 40 MW of thermal power input with waste) associated
tion of RDF and syngas co-combustion, after significative cleaning with plasma refining of syngas and its utilization in an ICE. If pul-
process (including the removal of tar, particulates, hydrogen sul- per residues are fed to a similar process (LHV 21.4 GJ/Mg) the cal-
fide, ammonia, and chlorine and fluorine compounds), in an exist- culated net electric efficiency is 27.4%.
ing natural gas combined cycle power plant, substituting part of
the natural gas.
Di Gregorio and Zaccariello (2012) study the bubbling fluidized 4. Pyrolysis of waste with energy recovery
bed air blown gasification of packaging derived fuel (LHV 23.2 GJ/
Mg) on small industrial scale application and propose the use of Pyrolysis is a thermal process which takes place in complete
an ICE after appropriate syngas cleaning. For a small system with absence of oxygen, heating the entering feed by an external heat
a net electric power output of about 500 kWe, corresponding to source, at temperatures higher than 400 °C and generally lower
waste entering flow of about 500 kg/h, i.e., about 4000 Mg/y, it is than 800 °C. It produces three output streams-gas, liquid (oil)
possible to calculate a net electric efficiency of about 15%. and solid (char) – the all of them with combustible characteristics.
Thanks to the absence of oxygen, no oxidation occurs, while the
feeding organic material undergoes to a thermal degradation, in
3.3. Plasma some cases with the help of catalysts. The yields of pyrolysis gas,
oil and char depend mainly on the process temperature and heat-
Recently, several studies were carried out regarding the possi- ing rate. Generally the higher the temperature the larger the gas-
bility of realizing the gasification process by applying plasma tech- eous fraction. While greater yields of oil can be obtained for
nology, probably because more efficient and reliable torches for lower temperature and fast heating rate. A high yield of oil is, in
thermal plasma generation have become available with respect general, of particular interest for the easy to carry and store it.
to the past (Tang et al., 2013). During the last thirty years, plasma While a large development of biomass pyrolysis was carried out
technology applied to organic waste disposal was extensively in the last years, pyrolysis of waste is still a step behind and mainly
investigated and small-scale plants were developed, with reference at research/pilot level.
to many types of wastes, including ASR, sludge, asbestos fibers, In particular, several studies about some selected waste pyroly-
medical waste, and MSW (Tang et al., 2013). However, the most sis are available (i.e. different types of plastic wastes, scrap tyres,
important application of thermal plasma process is devoted to wood waste, waste electric and electronic equipment, RDF) which
destruction of hazardous wastes rather than energy recovery. mainly focus on the pyrolysis process itself, rather than on the fur-
Thermal plasma gasification is based on the partial thermal oxi- ther possible uses of the obtained fuels and on an overall energy
dation of waste, producing syngas and solid residue. Steam, air or balance.
oxygen may be supplied to the process as an oxidizing agent. The Actually the only one MSW pyrolysis plant in operation in Eur-
high temperature of the plasma arc greatly reduces the potential ope is in Burgau in Germany and it has been working since 1986
for undesirable compounds in the syngas and the solid residues (Pyrolysis facility, 2004). The technology supplier for the Burgau
are produced in the form of a vitrified slag. plant was WasteGen UK Ltd. The plant processes about
Tang et al. (2013) report an overview of the types of torches to 38,000 Mg/year MSW and sewage sludge. MSW must be prelimin-
generate thermal plasma, the different plasma working gases and ary shredded into a 300 mm maximum size. The heating value of
the various proposed types of reactors. feedstock was originally about 8.5 GJ/Mg, registering an increase

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx 13

over the years up to about 10.2 GJ/Mg. Pyrolysis takes place at Baggio et al. (2008) simulate an innovative system based on the
about 500 °C in a rotary kiln, which is indirectly heated by a por- pyrolysis of MSW (thermal input 34.8 MW, 7200 kg/h) which pro-
tion of the flue gases (approximately 20%) from syngas combus- duces solid char, oil and gas fuels used in a GT combined cycle for
tion. The syngas, approximately 636 m3/Mg, containing hydrogen, electric power generation. The thermal energy input for the pyro-
carbon monoxide, carbon dioxide, methane and other hydrocar- lysis is obtained from the combustion of a fraction of char and oil.
bons, has about 10–14 GJ/m3 LHV, and it is burnt, together with The gas, after filtration and compression, feeds two GTs. The
the landfill gas extracted from the nearby landfill, at temperature exhaust from the GTs, after post-combustion with char and oil,
in excess of 1250 °C. About 80% of the flue gases are fed to the boi- drives a steam turbine cycle. The authors estimate a net electric
ler (which was substituted in 2003 to cope with the increased efficiency of the plant of about 28–30%.
heating value of the MSW), producing steam to drive a 2.2 MW
steam turbine generator for power generation. Residual steam/
5. Overview of WtE at international level
condensate is piped to a nearby greenhouse. No data about effec-
tive electric energy production or efficiency are reported in the
Castaldi and Themelis (2010) report that more than 650 ther-
cited reference, however assuming 38,000 Mg/h, 8000 h/y of oper-
mal treatment plants process about 180 million Mg of MSW in
ation, 10.2 GJ/Mg LHV and gross power output 2.2 MW, the gross
EU, US, Japan, Canada and another 30 nations including the new-
electric conversion efficiency results about 16%, which is obviously
comer of China. The dominant WtE technology is the ‘‘mass burn’’
overestimated, since additional thermal input to the combustion
of as received MSW. In the following, an overview on energy recov-
chamber is supplied by the landfill gas.
ery levels, typical of different countries and regions, is reported.
Pyrolysis – also applied to MSW – collected a wide success in
Japan, where at the beginning of 2000s several plants based on
the Mitsui Recycling 21 technology were built (Hiroaki, 2003). This 5.1. European union
process is based on waste shredding and processing in a pyrolysis
drum. Pyrolysis gas and carbon are burnt in a combustion cham- At European level, the EU Waste Framework Directive (Directive
ber. The boiler produces steam for power production in a steam 2008/98/EC) defines the criteria for determining if WtE plants are
cycle. Alternatively, the pyrolysis carbon could be supplied to Recovery (R1) or Disposal (D10) operations.
cement plants. The process is also able to recover aluminum and The Directive introduces the energy recovery indicator (R1):
steel. R1 ¼ ðEp  ðEf þ Ei ÞÞ=ð0:97ðEw þ Ef ÞÞ ð5Þ
Aside from these examples of MSW pyrolysis plants, which in
any case are based on a conventional steam cycle for the energy where Ep is the weighted sum of gross electric and thermal energy
recovery section and do not exploit the potential of using the gas- generated by the plant (GJ/year), with electric energy weighed by a
eous fuel in highly efficient internally-fired cycles, no additional factor 2.6 and thermal energy weighed by a factor 1.1; Ef is energy
MSW pyrolysis facilities are planned at present (Bilitewski, in fuels contributing to steam production of steam (GJ/year); Ew is
2012). However, pyrolysis is still interesting if used in combination energy in the waste (GJ/year); Ei is energy imported other than Ew
with special waste fractions like plastic, rubber tyres, sewage and Ef (GJ/year); 0.97 is a coefficient accounting for energy losses
sludge and waste wood. associated to bottom ash and radiation.
In particular scrap tyres received special attention from 1990s According to Directive 2008/98/EC, to be considered a recovery
as feeding stream to pyrolysis (Sharma et al., 2000; Fernández plant, a WtE plant must reach R1 higher than 0.65 (or, for plants
et al., 2012; Williams, 2013). A waste tyre pyrolysis plant is in permitted before January 1st, 2009, R1 higher than 0.60).3
operation in Japan (Nippon Steel & Sumikin Engineering, 2014). Additionally, some European Countries – as The Netherlands
Here gas, oil, steel wire, and carbon generated by the carbonization and Austria – and Switzerland put into force some specific criteria
and thermal decomposition of waste tires are reused at steel for evaluating the energy efficiency level of WtE plants with the
plants. aim of defining the subsidies, as better detailed by Gohlke (2009).
However, still large research is carried out in this field First the Confederation of European Waste-to-Energy Plants
(Martinez et al., 2013). Pyrolysis of scrap tyres can produce gases (CEWEP) investigated the R1 approach in reference to the 97 most
with LHV higher than 40 MJ/m3N, which can sufficiently covers efficient European WtE plants (Reimann, 2005). Later on Grosso
needs of the pyrolysis plant, excluding start up period. Pyrolytic et al. (2010) performed the calculation of energy efficiency accord-
gas can be used as a sole fuel or it can be mixed with natural gas ing to the R1 formula for a significant number of European WtE
or propane in a pre-determined ratio while the solid char may be plants.
used either as a smokeless fuel, carbon black or activated carbon The CEWEP, then, published the third edition of its Energy
or can be gasified for the production of fuel gases (Antoniou and Report, with reference to 2007–2010 data, with the aim of calculat-
Zabaniotou, 2013). Pyrolysis oils from waste tyres can reach ing the R1 efficiency for 314 European WtE plants operated by
44 MJ/kg of LHV, depending on the tyre composition and the pro- CEWEP members from 17 European countries (15 EU countries
cess conditions; it can be used directly as a liquid fuels, in particu- plus Switzerland (CH) and Norway (NO)), processing 59.4 mil-
lar as alternative fuel in compression ignition engines, or added to lion Mg/y, which represents a share of 85.5% of the total inciner-
petroleum refinery feedstock (Martinez et al., 2013). Frigo et al. ated MSW of 20 European countries in EU 27 plus CH and NO in
(2014) evaluate the possibility of using the oil obtained by scrap 2009, in 448 plants (Reimann, 2012). In general very low results
tyres in blends with Diesel fuel in Diesel engines. For a blend of with R1 < 0.60 are found in small sized plants (throughput lower
20% of scrap tyres oil and 80% of Diesel (volume basis), the engine than 100,000 Mg/y), located in South-Western Europe producing
performances are not significantly affected. electricity only. For plants producing electricity only it is very dif-
Similarly, Kumar et al. (2013) study the performances of blends ficult to meet R1 criteria as only 37.3% of them meet R1 P 0.60.
of oil, obtained by catalytic pyrolysis of waste high-density poly- The highest R1 results are related to large sized plants (throughput
ethylene, with diesel in a ICE with varying loads. higher than 250,000 Mg/y), located in Northern Europe with CHP
Concerning other type of wastes, Singhabhandhu and Tezuka production (Reimann, 2012). Of the 314 CEWEP plants, 83 operate
(2010) investigate the integrated co-processing of waste cooking
oil, waste lubricating oil and waste plastics by pyrolysis, with the 3
Possible adjustment of the R1 criteria according to the country specific climate
aim of producing pyrolysis oil for energy recovery. condition are left open to Member State.

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
14 L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx

producing only electricity, 47 produce only heat and 184 produce 5.3. Japan
Combined Heat and Power (CHP) (Reimann, 2012).
Even though the R1 compliance criterion is a uniform and for- Castaldi and Themelis (2010) report that Japan processes by
mal parameter for European plants, other parameters can be more thermal treatment about 40 million Mg/y of MSW, being the larg-
widely used for comparison with plants from other countries. est user of this process in the world, and the main used technology
Table 4 shows the specific energy production from WtE plants in is grate incineration of non-pretreated MSW.
EU, which can be compared with other data. Gohlke and Martin (2007) explain that landfilling is tradition-
Reimann (2012) also reports the total annual energy production ally not applied in Japan due to the lack of space and this is the rea-
by the 285 of the considered 314 plants (only EU 27 countries), in son why the largest part of waste is incinerated in a high number of
reference to 2007–2010 investigated years: net produced heat is small plants.
about 54.41 TWh/y and net produced electricity is 23.69 TWh/y. Tabata (2013) reports that, in Japan, about 80% of MSW is incin-
Considering that the final energy consumption of electricity in erated, but only 24.5% MSW incineration plants apply energy
years 2007–2010 in EU 27 is 2836–2822 TWh/y (Eurostat data), recovery. Considering the plants with energy recovery, the national
the production from the considered WtE plants is about 0.84% of average power generation efficiency is quite low (11.3%), produc-
the total consumption. ing, on average, 200 kWh/Mg, with higher values for large installa-
Each EU country is free to set the criteria to decide which por- tions (300–390 kWh/Mg). The same author reports, similarly, that
tion of the energy recovered by incineration plants can be classified heat production is quite low (average 0.76 kWh/Mg) and it is used
as renewable. Such criteria should be based on the effective bio- only for heating pools or hot water in public facilities, while district
mass content of the incinerated waste, but many EU countries heating was introduced only since few years.
use just a 50% ratio. The reason for such an intensive use of thermal treatment –
EurObserv’ER (2012, 2011) data show that EU 27 primary even without energy recovery – is explained by Tanigaki et al.
energy production from renewable waste incineration was about (2012) highlighting how the landfills use reduction is a major con-
99 TWh in 2011 (93 TWh in 2010), generated by municipal and cern and one of the main objectives for MSW management in
other waste incineration plants. Additionally, the renewable elec- Japan. Tanigaki et al. (2012) also add that, generally in Japan, treat-
tricity produced by MSW incineration was about 18.3 TWh in ments of MSW incinerator bottom ash, such as melting, is required
2011 (17.2 TWh in 2010) (EurObserv’ER, 2012, 2011). Thus, consid- before landfilling, due to Japanese strict regulations in waste
ering that renewable electricity production in EU 27 was about management.
661.4 TWh in 2010 (EurObserv’ER, 2011), waste incineration con- These conditions promoted the development of processes able
tributes with 2.6% to the total (5.7% excluding hydropower). to produce vitrified residues (ash melting) as an appendix to con-
Some EU country specific studies are available, concerning WtE ventional incineration processes, or as main modifications of the
plants. incineration process (i.e. oxygen enrichment), or as integrated
Nixon et al. (2013b) report data for 25 operational MSW WtE treatments in high temperature gasification process.
plants in the UK. The most common incineration technology is This is the main reason why a lot of gasification plants for MSW,
moving grate (23 plants), while only one plant is based on fluidized including gasification and melting processes, are under commer-
bed and one on rotary kiln. Waste throughput varies from 56,000 cial operation for energy and material recovery, in Japan
to 500,000 Mg of MSW per year. Reported energy efficiency, (Tanigaki et al., 2012).
assuming MSW LHV 10 GJ/Mg, is about 21–24% for only electricity
production, while it is about 31%, in case of CHP.
5.4. China
Beylot and Villeneuve (2013) investigate 110 French incinera-
tors reporting that 37% of plants produce only electricity with an
Cheng and Hu (2010) show that MSW generation increased in
average efficiency of 14%; 21% of plants produce only heat with
China at an annual rate of 8–10%, producing presently over 150 mil-
43% average efficiency; 26% of plants cogenerate heat and power
lion Mg/y. In the 1980s only 5% of generated MSW were processed
with average efficiency of 33%.
in waste management system, while presently about 55% of MSW is
managed by a combination of landfilling, composting, and inciner-
ation, being landfilling the dominant form of waste disposal (80%),
5.2. USA
while the amount of incinerated MSW increased steadily, becoming
the second most important way of MSW management, mainly with
In the USA, WtE plants are in operation in 25 states. They are
the aim of volume reduction, rather than energy recovery. In regard
fueled by 26.3 million Mg/y of MSW, which represents 7.4% of
to this, it should be mentioned that the typical energy content of
overall produced MSW. Most of the WtE facilities are on the East
MSW in China is generally quite low – with respect to Europe or
coast, corresponding to 66% of the total WtE capacity in the USA.
North America – with an average low heating value of 5 GJ/Mg, with
65 WtE plants (20.05106 Mg/y) are mass-burn plants where the
values down to 4 GJ/Mg in typical Chinese towns and higher values
MSW is fed as collected into large furnaces, while 15 plants
in larger cities (6–7 GJ/Mg) (Nie, 2008). Also Yang et al. (2012)
(5.71106 Mg/y) are fed by RDF (Psomopoulos et al., 2009).
report data about quite low heating value of MSW from six cities
Psomopoulos et al. (2009) report that, according to operating
data collected by the US WtE industry, the specific net electricity Table 4
production from MSW in a modern WtE power plant is 600 kWh/ Specific energy (heat and/or electricity) production in MSW WtE in EU (Reimann,
Mg, on the average. The same authors clarify also that the US 2012).

Department of Energy (US DOE) categorizes WtE plants as a type Only Only CHP
of biomass plant, adding that even if a more stringent definition electricity heat
of the term ‘‘renewable’’ is applied, including only material from Specific gross electricity produced (kWh/Mg) 581 – 444
non-fossil sources, about 64% of the US MSW, after material recov- Specific electricity self used (kWh/Mg) 105 – 106
ery for recycling and composting, is derived from renewable origin. Specific net electricity produced (kWh/Mg) 476 – 338
Specific gross heat produced (kWh/Mg) – 2300 1101
According to this approach, WtE represents 28% contribution to
Specific heat self used (kWh/Mg) 122 146 163
renewable electric energy (excluding hydropower) in the USA Specific net heat produced (kWh/Mg) 122 2154 938
(Psomopoulos et al., 2009).

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
Table 5
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from

Summary of operating conditions and performances.

Source Process/ Waste type LHV Size gel gth SEP SHP Steam Tout Pcond Note
technology parameters
Pel Pth Waste throughput Gross Net Gross Net Gross Net P T
GJ/Mg MW MW Mg/y kWh/ kWh/ kWh/ kWh/ bar °C °C bar
Mg Mg Mg Mg
(bMg/d; cMg/h)
Combustion – data from existing plants
Gohlke and Martin (2007) COMB/GRATE MSW 9.08 590b 737 623 60 443
Gohlke and Martin (2007) COMB/GRATE MSW, sewage sludge, 551,728a 27 24 61 450 135
biomass
a
Gohlke and Martin (2007) COMB/GRATE Biomass, sewage sludge 257,599 28 25 73 480 135
Gohlke and Martin (2007) COMB/GRATE MSW 34 30 130 440 135–180 1
Gohlke (2009) COMB/GRATE MSW 10 15 21.6c 20.6 40 380 209 0.15 2
Main and Maghon (2010) COMB/GRATE 12.00 40 26.63 41 400
Main and Maghon (2010) COMB/GRATE 10.5 51 28.14 42 400 100 3
Main and Maghon (2010) COMB/GRATE 12.00 58 29.68 81 520 4
Main and Maghon (2010) COMB/GRATE 15.10 3x 30.2 90 500

L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx


113.3
Main and Maghon (2010) COMB/GRATE 14.5 110 29.9 90 420 5
Murer et al. (2011) COMB/GRATE MSW 10.00 530,000 34.5 30 130 440 130–180 0.03 6
Stehlík (2012) COMB/GRATE MSW 96,000 115 2025 40 400 7
Combustion – general data or simulations
Consonni and Viganò (2012) COMB/GRATE MSW 10.34 50–300 135,652–813,913 24.6– 20.5– 40–70 400– 135
32 28.5 450
Gohlke and Martin (2007) COMB/GRATE MSW 10.44 18 13 546 396
Gohlke and Martin (2007) COMB/GRATE MSW 10.44 22 18 640 520 40 400
Gohlke (2009) COMB/GRATE MSW 24 21 40 380
Gohlke (2009) COMB/GRATE MSW 17 30 40 380
Pavlas et al. (2011) COMB/GRATE MSW 10.9 100,000 15–20 600– 40–60 400 250 0.08
800
Pavlas et al. (2011) COMB/GRATE MSW 10.9 100,000 4–14 70– 200– 40–60 400 250 11
60 500
b
Udomsri et al. (2011) COMB/GRATE MSW 148 1350 24.3 36 380 200 0.056 8
Udomsri et al. (2011) COMB/GRATE MSW 148 1350b 15.5 74.3 36 380 200 1.65 8
Van Caneghem et al. (2012) COMB/FB SRF, sewage sludge, ASR 27
Gasification: syngas combustion in a boiler
Calaminus and Stahlberg (1998) GAS/HT MSW, industrial waste 12.00 4 225,000 12.5 50
Ligthart (2013) GAS/HT/FB SRF 250,000 31 121 540
Tanigaki et al. (2012) GAS/HT/DMS MSW + COKE 6.8–9.1 252b + (45–49 kgCOKE/ 18.9– 408– 39.2 400
MgMSW) 23 673
Arena et al. (2011) GAS/HT Mixed plastic waste 31.7– 1.5–1.9c 23.7 500
40.2
Gohlke (2009) GAS/HT MSW 10 320– 150–
390 190
Consonni and Viganò, 2012 GAS/HT/ 10.34 50 135,652 23.8 19.7 40 400 135
GRATE
Consonni and Viganò (2012) GAS/LT/FB 10.34 50 135,652 23.7 19.6 40 400 135
Consonni and Viganò (2012) GAS/HT/ 10.34 300 813,913 31 27 70 450 135
GRATE
Consonni and Viganò (2012) GAS/LT/FB 10.34 300 813,913 31.5 27.5 70 450 135
Lombardi et al. (2012) GAS/HT RDF 18.5 12.85c 13.4 56 405 132 0.18
Viganò et al. (2010) GAS/LT/FB ASR 19.2 100 140,625 25.7 22.5 30 325 205 0.06
Gasification: syngas use in internally-fired devices
Di Gregorio and Zaccariello GAS/HT Packaging derived fuel 23.2 0.5 4000 15 ICE
(2012)

15
(continued on next page)
16 L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx

in China in the range of 3.73–5.73 GJ/Mg. As a result, incineration of

GTCC

GTCC
Chinese MSW has several problems related to difficulty in ignition,
Pcond Note

ICE
unsteady and unstable combustion flame, incomplete combustion
of the waste, increased formation of air pollutants. To overcome
bar

these problems supplementary fuel is often used, significantly


increasing the operating cost.
Regarding the type of incineration technology, mass burning
Tout

with a grate furnace is the mostly employed in China (Yang


°C

et al., 2012).
When the energy recovery is present, the achieved efficiency is
°C
T
parameters

quite low. Yang et al. (2012) reported an annual average value of


the gross electricity conversion efficiency of 17%, for the MSW
Steam

mass burning with a grate furnace plant in Shanghai, recovering


bar
P

only electricity. While Cheng et al. (2007) report an overall fuel-


kWh/

to-electricity efficiency of 14.6%, for a novel WtE incineration tech-


Net

Mg

nology based on co-firing of MSW with coal in a grate-circulating


fluidized bed (CFB) incinerator, which was implemented in the
Gross
kWh/
SHP

Mg

Changchun MSW power plant.


However, Cheng and Hu (2010), on the basis of several aspects
kWh/

favoring the diffusion of WtE, among which the development and


Net

Mg

LT gasification is intended at about 600 °C; HT gasification is intended at temperature higher than 700 °C, typically in the 700–950 °C range, but even higher.

availability of local technologies, affirm that WtE process is expected


Gross
kWh/

to supply increasingly greater contribution to produce renewable


SEP

Mg

energy in China, that is presently relying mainly on coal, and at the


same time contributing to appropriate MSW management.
gth

5.5. India
23–25
24–27
14.7–
18.7
Gross Net

Nixon et al. (2013a) report that even if in 2000 the MSW Man-
agement and Handling Rules were introduced in India, the collec-
gel

6 – Intermediate steam reheating (14 bar; 320 °C); water condenser; recovery of available heat at different temperature levels.

tion efficiency of MSW is still only about 70% and waste is


mainly dumped or landfilled, with about 90% of landfills being
unsatisfactory. Sharholy et al. (2008) already reported that MSW
Waste throughput

50,000–100,000
50,000–100,000

treatment and disposal methods in India mainly include landfill-


(bMg/d; cMg/h)

ing, composting (aerobic and vermicomposting) and very few


waste to energy initiatives (incineration, RDF and biomethanation).
12.85c

Calorific value ranges between 3.3 and 4.2 GJ/Mg (Unnikrishnan


Mg/y

3 – External economizer to 140 °C and further to 100 °C in a flue gas heat exchanger, for air pre-heating.

and Singh, 2010).


Singh et al. (2011) conclude that in India WtE facility – includ-
MW MW

ing in the term also biogas production – is not only possible but
Pel Pth

necessary in order to meet the demands of growing cities and


Size

improve environmental conditions.


GJ/Mg

18.5

16.7
16.7
LHV

2 – In-plant consumption 100 kWh/Mg; excess air 1.75 (flue gas O2 of 8.4% dry).

6. Discussion

Table 5 summarizes the main data about energy production col-


8 – Air preheating by further exhaust from steam generator to 200 °C.

lected during this review study, including incineration, gasification


with syngas use in a boiler and gasification with syngas use in
Waste type

Data from operation report of the Brescia plant, in Italian.

internally-fired devices. Data about plasma and pyrolysis are left


outside from the table.
1 – Intermediate steam reheating and water condenser.
RDF

RDF
RDF

Regarding pyrolysis, aside from some theoretical studies, appli-


7 – SEP and SHP calculated from data in the paper.

cation to waste for energy recovery is still limited to few specific


technology

waste streams, quite pure and, hence, able to provide good quality
Process/

4 – External superheaters using natural gas.


GAS/HT

GAS/HT
GAS/HT

oil, which can be used in highly efficient energy conversion devices


(Arena, 2011).
Daily waste throughput (Mg/day).
Hourly waste throughput (Mg/h).

Concerning plasma, few data are available and processes are


5 – Intermediate steam reheating.

still at development stage.


Given that cogeneration can improve in the same way the per-
formances of both systems based on incineration and gasification,
Lombardi et al. (2012)

a synthetic comment to Table 5 is reported here in reference to


Yassin et al. (2009)
Yassin et al. (2009)

achievable electric efficiency levels.


Table 5 (continued)

– WtE incineration plants of relatively low size (throughput lower


than 100,000 Mg/y, that assuming a LHV of 10 GJ/Mg, is about
Source

Notes:

35 MW of entering thermal power) may reach a gross electric


a

c
b

efficiency of 18–26%, relying on steam parameters not higher

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx 17

than 40 bar/400 °C, since, as we stated before, the increased possibility of recovery the valuable energy content of the waste,
investment and operating costs can be withstood only for large in part effectively renewable, and in any case produced daily by
throughput plants; efficiency can be further increased of few the local activities. In this sense waste is a domestic energy source
percentage points by intensive heat recovery from low temper- widespread on the territory and continuously produced. Its contri-
ature fluxes in the plant. bution to countries’ energy production and consumption is pres-
– Large plants (throughput larger than 250,000 Mg/y, that assum- ently still very low, but it may be increased, increasing the plant
ing a LHV of 10 GJ/Mg, is about 88 MW of entering thermal performances and, even more, making up for the lack of thermal
power) may use improved steam parameters – 60–70 bar/ treatment plants in several countries worldwide.
450–500 °C – and reach up to 27–29% gross electric efficiency. The dominant technology in thermal treatment is incineration
– Very modern and large plants can use additional measures to with energy recovery in a steam cycle. Among the furnaces avail-
improve efficiency, as reheating, external superheating, able for waste incineration, the mobile grate has reached a high
demanding heat recovery from low temperature fluxes, steam state of development and is the most popular worldwide.
pressure up to 130 bar and steam temperature up to 500 °C, Waste gasification is applied producing syngas which is gener-
thus reaching a gross electric efficiency of 30–34%. ally combusted in a boiler.
– Air gasification plants, with syngas use in a boiler, of relatively It is important to stress that for both the mentioned possibilities
low size (throughput lower than 100,000 Mg/y), similarly to – incineration or gasification – cogeneration is the mean to
incineration plants, may rely on steam parameters not higher improve energy recovery, especially for small scale plants.
than 40 bar/400 °C again for scale matter and may reach effi- When discussing of only power production, indeed, the energy
ciency levels of about 23–24%. performances, of both plants based on incineration or gasification
– Air gasification plants, with syngas use in a boiler, of large size with syngas use in a boiler, are directly linked to the technological
(throughput higher than 250,000 Mg/y), may reach about 30– level of the designed steam cycle, mainly meaning maximum steam
31% gross efficiency, by using high pressure and temperature pressure and temperature and complex cycle arrangements. High
steam (70–120 bar; 450–540 °C). technological level of the steam cycle may be sustained in general
only for large scale plants, being a compromise between invest-
It is quite evident that for both incineration and gasification ments and profit related to efficiency increase and electricity sale.
(with syngas use in a boiler) the possibility of obtaining high values It this view, it is possible to summarize that large scale – incin-
of electric efficiency is substantially linked to the possibility of eration or gasification plants – may reach up to 30–31% net electric
applying high pressures for the steam cycle. For the most recent efficiency, in only power mode. Small-medium size incineration or
and large plants – both based on incineration or gasification and gasification plants generally operate with steam at 40–50 bar and
providing appropriate measures for avoiding high temperature 400 °C, with maximum net electric efficiency around 20–24%.
corrosion – the maximum values of steam parameters are similar However at comparable size and steam cycle technology, the
and hence similar values of efficiency are found. gasification process is more complex and costly, more difficult to
However at comparable size and steam cycle technology, the operate and maintain and less reliable than the incineration one.
complexity of the process remains higher in the case of gasifica- Moreover, gasification reactors require, for a large number of types
tion. Moreover, gasification reactors require, for a large number of feeding streams, to be fed by pre-treated waste. Hence, in the
of types of feeding streams, to be fed by pre-treated waste. Hence, overall energy balance pre-treatment consumptions and losses
in the overall energy balance pre-treatment consumptions and must be included.
losses must be included. For small-medium size plants – either based on incineration or
Further, since data about efficiency are quite often provided by gasification – the only effective and practicable way to improve the
modeling and simulation, special caution is necessary when com- energy recovery is cogeneration, which also allows accomplishing
paring the values among themselves. In this case it would be more some regulation constraints (R1 criterion in EU) and providing
reliable a comparison of different processes simulated on the basis additional beneficial effects in terms of environmental savings.
of the same assumptions (i.e. type of waste) and same tools. For Studies related to gasification with syngas use in internally-
this reason, results provided by some authors for different types fired devices provided efficiency estimation at levels similar to
of energy conversion systems are valuable for relative comparison. those of conventional plants based on incineration.
Looking at results available from studies related to gasification With the aim of producing a clean syngas, the use of thermal
with syngas use in internally-fired devices, estimated efficiency plasma in a two steps process – made of conventional gasification
levels are still not so far from those of conventional WtE plants followed by syngas refining by plasma – seems to be promising,
based on incineration. but more data about pilot operation should be provided to the sci-
In reference to gasification, in general, it should be kept in mind entific community.
that the overall gasification and energy conversion plant would Pyrolysis application to waste for energy recovery is limited to
become definitely more complex than a conventional WtE, thus few specific waste flows. In particular pure and homogeneous
more difficulties in operation, additional maintenance and reduced waste streams are required to produce good quality oil, which
reliability are expected. In reference to this issue, one should can be used in highly efficient energy conversion devices. Investi-
remind that the first purpose of a waste thermal treatment plant gation of the real possibility of using the produced fuels in such
is waste treatment and, especially when dealing with MSW, fre- devices deserves special attention.
quent or prolonged shutdowns may result in waste accumulation Generally speaking, a lack of published data about plant perfor-
first in the plant bunker and then along the streets. mances was recorded. Being this work a review primarily of the
scientific literature about energy recovery from waste, the authors
limited the reported data to those found in scientific journals and
7. Conclusions conferences (with few exceptions), thus excluding data that can
be found in other sources (plant technical reports, personal com-
The thermal treatment of waste needs to be present within the munications, etc.). In this sense the authors wish to encourage a
integrated waste management system and plays an important role larger publication of real plant data by manufacturers, associations
in it. The basic ability of thermal treatment of reducing mass/vol- and researchers, especially in those cases where particular innova-
ume of wastes and their putrescibility is today coupled with the tive features are implemented.

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
18 L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx

Acknowledgment Di Lonardo, M.C., Lombardi, F., Gavasci, R., 2012. Characterization of MBT plants
input and outputs: a review. Rev. Environ. Sci. Biotechnol. 11, 353–363.
Directive 2000/76/EC of the European Parliament and of the Council on the
The authors are fully grateful to Prof. Stefano Consonni and Dr. Incineration of Waste I, 4 December 2000.
Federico Viganò for their precious suggestions. Directive 2008/98/EC of the European Parliament and of the Council on Waste and
Repealing Certain Directives – Annex II, 19 November 2008.
EurObserv’ER, 2011. The State of Renewable Energies in Europe, 11th ed.
EurObserv’ER Report. <http://www.eurobserv-er.org/> (accessed May 2014).
References EurObserv’ER, 2012. The State of Renewable Energies in Europe, 12th ed.
EurObserv’ER Report. <http://www.eurobserv-er.org/> (accessed May 2014).
European Commission, 2006. Reference Document on the Best Available Techniques
Antoniou, N., Zabaniotou, A., 2013. Features of an efficient and environmentally
for Waste Incineration. <http://eippcb.jrc.ec.europa.eu/reference> (accessed
attractive used tyres pyrolysis with energy and material recovery. Renew.
May 2014).
Sustain. Energy Rev. 20, 539–558.
Fellner, J., Cencic, O., Rechberger, H., 2007. A new method to determine the ratio of
Arena, U., 2011. Gasification: an alternative solution for waste treatment with
electricity production from fossil and biogenic sources in waste-to-energy
energy recovery. Waste Manage. 31, 405–406.
plants. Environ. Sci. Technol. 2007 (41), 2579–2586.
Arena, U., 2012. Process and technological aspects of municipal solid waste
Fernández, A.M., Barriocanal, C., Alvarez, R., 2012. Pyrolysis of a waste from the
gasification. A review. Waste Manage. 32, 625–639.
grinding of scrap tyres. J. Hazard. Mater. 203–204, 236–243.
Arena, U., Di Gregorio, F., 2014. Gasification of a solid recovered fuel in a pilot scale
Frigo, S., Seggiani, M., Puccini, M., Vitolo, S., 2014. Liquid fuel production from waste
fluidized bed reactor. Fuel 117, 528–536.
tyre pyrolysis and its utilisation in a Diesel engine. Fuel 116, 399–408.
Arena, U., Di Gregorio, F., Amorese, C., Mastellone, M.L., 2011. A techno-economic
Fruergaard, T., Christensen, T.H., Astrup, T., 2010. Energy recovery from waste
comparison of fluidized bed gasification of two mixed plastic wastes. Waste
incineration: assessing the importance of district heating networks. Waste
Manage. 31, 1494–1504.
Manage. 30, 1264–1272.
Baggio, P., Baratieri, M., Gasparella, A., Longo, G.A., 2008. Energy and environmental
Galeno, G., Minutillo, M., Perna, A., 2011. From waste to electricity through
analysis of an innovative system based on municipal solid waste (MSW)
integrated plasma gasification/fuel cell (IPGFC) system. Int. J. Hydrogen Energy
pyrolysis and combined cycle. Appl. Therm. Eng. 28, 136–144.
36, 1692–1701.
Balcazar, J.G.C., Dias, R.A., Balestieri, J.A.P., 2013. Analysis of hybrid waste-to-energy
Giugliano, M., Grosso, M., Rigamonti, L., 2008. Energy recovery from municipal
for medium-sized cities. Energy 55, 728–741.
waste: a case study for a middle-sized Italian district. Waste Manage. 28, 39–50.
Barigozzi, G., Perdichizzi, A., Ravelli, S., 2011. Wet and dry cooling systems
Gohlke, O., 2009. Efficiency of energy recovery from municipal solid waste and the
optimization applied to a modern waste-to-energy cogeneration heat and
resultant effect on the greenhouse gas balance. Waste Manage. Res. 27, 894–
power plant. Appl. Energy 88, 1366–1376.
906.
Bebar, L., Stehlik, P., Havlen, L., Oral, J., 2005. Analysis of using gasification and
Gohlke, O., Busch, M., 2001. Reduction of combustion by-products in WTE plants: O2
incineration for thermal processing of wastes. Appl. Therm. Eng. 25, 1045–1055.
enrichment of underfire air in the MARTIN SYNCOM process. Chemosphere 42,
Belgiorno, V., De Feo, G., Della Rocca, C., Napoli, R.M.A., 2003. Energy from
545–550.
gasification of solid wastes. Waste Manage. 23, 1–15.
Gohlke, O., Martin, J., 2007. Drivers for innovation in waste-to-energy technology.
Beylot, A., Villeneuve, J., 2013. Environmental impacts of residual Municipal Solid
Waste Manage. Res. 25, 214–219.
Waste incineration: a comparison of 110 French incinerators using a life cycle
Graus, W., Worrell, E., 2009. Trend in efficiency and capacity of fossil power
approach. Waste Manage. 33, 2781–2788.
generation in the EU. Energy Policy 37, 2147–2160.
Bilitewski, B., 2012. Chances and limitations of pyrolysis. In: Proceedings of Crete
Graus, W., Voogt, M., Worrell, M., 2007. International comparison of energy
2012. 3rd International Conference on Industrial and Hazardous Waste
efficiency of fossil power generation. Energy Policy 35, 3936–3951.
Management, September 12–14, 2012 Chania (Crete, Greece).
Groß, B., Eder, C., Grziwa, P., Horst, J., Kimmerle, K., 2008. Energy recovery from
Bovea, M.D., Ibáñez-Forés, V., Gallardo, A., Colomer-Mendoza, F.J., 2010.
sewage sludge by means of fluidised bed gasification. Waste Manage. 28, 1819–
Environmental assessment of alternative municipal solid waste management
1826.
strategies. A Spanish case study. Waste Manage. 30, 2383–2395.
Grosso, M., Motta, A., Rigamonti, L., 2010. Efficiency of energy recovery from waste
Calabrò, P.S., 2010. The effect of separate collection of municipal solid waste on
incineration, in the light of the new Waste Framework Directive. Waste
lower calorific value of the residual waste. Waste Manage. Res. 28, 754–758.
Manage. 30, 1238–1243.
Calaminus, B., Stahlberg, R., 1998. Continuous in-line gasification/vitrification
Hiroaki, H., 2003. Operating experience with pyrolysis/gasification of wastes and
process for thermal waste treatment: process technology and current status
ash melting in energy recovery from waste materials. Joint meeting of IEA
of projects. Waste Manage. 18, 547–556.
Bioenergy Task 32, 33 and 36. Operating Experience and Techno-economic
Castaldi, M.J., Themelis, N.J., 2010. The case for increasing the global capacity for
Benefits and Environmental Benefits of Energy Recovery from Renewable Waste
waste to energy (WTE). Waste Biomass Valor. 1, 91–105.
Materials. Mielparque Hotel, Tokyo, 28 October, 2003. <http://www.ieabcc.nl/
CEN/TS 15359, 2006. Solid Recovered Fuels – Specifications and Classes. European
workshops/Tokyo_Joint_Meeting/02_Mitsui.pdf> (accessed September 2014).
Committee for Standardisation.
Horttanainen, M., Teirasvuo, N., Kapustina, V., Hupponen, M., Luoranen, M., 2013.
Cengel, Y.A., Boles, M., 1998. Thermodynamics: An Engineering Approach, third ed.
The Composition, Heating Value and Renewable Share of the Energy Content of
McGraw-Hill Companies.
Mixed Municipal Solid waste in Finland. Waste Manage., http://dx.doi.org/
Chen, D., Christensen, T.H., 2010. Life-cycle assessment (EASEWASTE) of two
10.1016/j.wasman.2013.08.017 (in press).
municipal solid waste incineration technologies in China. Waste Manage. Res.
Ionescu, G., Rada, E.C., Ragazzi, M., Marculescu, M., Badea, A., Apostol, T., 2013.
28, 508–519.
Integrated municipal solid waste scenario model using advanced pretreatment
Cheng, H., Hu, Y., 2010. Municipal solid waste (MSW) as a renewable source of
and waste to energy processes. Energy Convers. Manage. 76, 1083–1092.
energy: current and future practices in China. Bioresour. Technol. 101, 3816–
ISWA, 2012. Waste-to-Energy. State-of-the-Art-Report. Statistics, 6th ed. <http://
3824.
www.iswa.org/media/publications/knowledge-base/> (accessed May 2014).
Cheng, H., Guozhang, Y., Meng, A., Li, Q., 2007. Municipal solid waste fueled power
Jegla, Z., Bebar, L., Pavlas, M., Kropac, J., Stehlik, P., 2010. Secondary combustion
generation in China: a case study of waste-to-energy in Changchun city.
chamber with inbuilt heat transfer area – thermal model for improved waste-
Environ. Sci. Technol. 41, 7509–7515.
to-energy systems modelling. Chem. Eng. Trans. 21, 859–864.
Cimpan, C., Wenzel, H., 2013. Energy implications of mechanical and mechanical–
Komilis, D., Kissas, K., Symeonidis, A., 2013. Effect of organic matter and moisture on
biological treatment compared to direct waste-to-energy. Waste Manage. 33,
the calorific value of solid wastes: an update of the Tanner diagram. Waste
1648–1658.
Manage. http://dx.doi.org/10.1016/j.wasman.2013.09.023 (in press).
Consonni, S., Silva, P., 2007. Off-design performance of integrated waste-to-energy,
Korobitsyn, M.A., Jellema, P., Hirs, G.G., 1999. Possibilities for gas turbine and waste
combined cycle plants. Appl. Therm. Eng. 27, 712–721.
incinerator integration. Energy 24, 783–793.
Consonni, S., Viganò, F., 2012. Waste gasification vs. conventional Waste-To-Energy:
Kumar, S., Prakash, R., Murugan, S., Singh, R.K., 2013. Performance and emission
a comparative evaluation of two commercial technologies. Waste Manage. 32,
analysis of blends of waste plastic oil obtained by catalytic pyrolysis of waste
653–666.
HDPE with diesel in a CI engine. Energy Convers. Manage. 74, 323–331.
Consonni, S., Giugliano, M., Grosso, M., 2005. Alternative strategies for energy
Lee, S.H., Themelis, N.J., Castaldi, M.J., 2007. High-temperature corrosion in waste-
recovery from municipal solid waste. Part A: Mass and energy balances. Waste
to-energy boilers. J. Therm. Spray Tech. 16, 104–110.
Manage. 25, 123–135.
Ligthart, F., 2013. Metso waste gasification. Case Lahti Energy. IWWG Workshop:
Dajnak, D., Lockwood, F.C., 2000. Use of thermal energy from waste for seawater
Energy Recovery from MSW and SRF – Case Studies. Sardinia 2013. In: 14th
desalination. Desalination 130, 137–146.
International Waste Management and Landfill Symposium. S. Margherita di
Damgaard, A., Riber, C., Fruergaard, T., Hulgaard, T., Christensen, T.H., 2010. Waste
Pula (Ca)-Italy, 30 September – 4 October, 2013.
Manage. 30, 1244–1250.
Liuzzo, G., Verdone, N., Bravi, M., 2007. The benefits of flue gas recirculation in
De Greef, J., Villani, K., Goethals, J., Van Belle, H., Van Caneghem, J., Vandecasteele, C.,
waste incineration. Waste Manage. 27, 106–116.
2013. Optimising energy recovery and use of chemicals, resources and materials
Lo Mastro, F., Mistretta, M., 2006. Thermoeconomic analysis of a coupled municipal
in modern waste-to- energy plants. Waste Manage. 33, 2416–2424.
solid waste thermovalorization–MSF desalination plant: an Italian case study.
de Souza-Santos, M.L., Ceribeli, K., 2013. Technical evaluation of a power generation
Desalination 196, 293–305.
process consuming municipal solid waste. Fuel 108, 578–585.
Lombardi, L., Carnevale, E., Corti, A., 2012. Analysis of energy recovery potential
Di Gregorio, F., Zaccariello, L., 2012. Fluidized bed gasification of a packaging
using innovative technologies of waste gasification. Waste Manage. 32, 640–
derived fuel: energetic, environmental and economic performances comparison
652.
for waste-to-energy plants. Energy 42, 331–341.

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010
L. Lombardi et al. / Waste Management xxx (2014) xxx–xxx 19

Lupa, C.J., Ricketts, L.J., Sweetman, A., Herbert, B.M.J., 2011. The use of commercial Ruth, A.L., 1998. Energy from municipal solid waste: a comparison with coal
and industrial waste in energy recovery systems – a UK preliminary study. combustion technology. Prog. Energy Combust. Sci. 24, 545–564.
Waste Manage. 31, 1759–1764. Sharholy, M., Ahmad, K., Mahmood, G., Trivedi, R.C., 2008. Municipal solid waste
Main, A., Maghon, T., 2010. Concepts and experiences for higher plant efficiency management in Indian cities – a review. Waste Manage. 28, 459–467.
with modern advanced boiler and incineration technology. In: Proceedings of Sharma, V.K., Fortuna, F., Mincarini, M., Berillo, M., Cornacchia, G., 2000. Disposal of
the 18th Annual North American Waste-to-Energy Conference NAWTEC18 May waste tyres for energy recovery and safe environment. Appl. Energy 65, 381–394.
11–13, 2010, Orlando, Florida, USA. Singh, R.P., Tyagi, V.V., Allen, T., Ibrahim, M.H., Kothari, R., 2011. An overview for
Malkow, T., 2004. Novel and innovative pyrolysis and gasification technologies for exploring the possibilities of energy generation from municipal solid waste
energy efficient and environmentally sound MSW disposal. Waste Manage. 24, (MSW) in Indian scenario. Renew. Sustain. Energy Rev. 15, 4797–4808.
53–79. Singhabhandhu, A., Tezuka, T., 2010. The waste-to-energy framework for integrated
Martinez, J.D., Puy, N., Murillo, R., Garcia, T., Navarro, M.V., Mastral, A.M., 2013. multi-waste utilization: waste cooking oil, waste lubricating oil, and waste
Waste tyre pyrolysis – a review. Renew. Sustain. Energy Rev. 23 (2013), 179– plastics. Energy 35, 2544–2551.
213. Stehlík, P., 2007. Heat transfer as an important subject in waste-to-energy systems.
Materazzi, M., Lettieri, P., Mazzei, L., Taylor, R., Chapman, C., 2013. Thermodynamic Appl. Therm. Eng. 27, 1658–1670.
modelling and evaluation of a two-stage thermal process for waste gasification. Stehlík, P., 2011. Conventional versus specific types of heat exchangers in the case
Fuel 108, 356–369. of polluted flue gas as the process fluid – a review. Appl. Therm. Eng. 31, 1–13.
Minutillo, M., Perna, A., Di Bona, D., 2009. Modelling and performance analysis of an Stehlík, P., 2012. Up-to-date technologies in waste to energy field. Rev. Chem. Eng.
integrated plasma gasification combined cycle (IPGCC) power plant. Energy 28, 223–242.
Convers. Manage. 50, 2837–2842. Tabasová, A., Kropác, J., Kermes, V., Nemet, A., Stehlík, P., 2012. Waste-to-energy
Morris, M., Waldheim, L., 1998. Energy recovery from solid waste fuels using technologies: impact on environment. Energy 44, 146–155.
advanced gasification technology. Waste Manage. 18, 557–564. Tabata, T., 2013. Waste-to-energy incineration plants as greenhouse gas reducers: a
Murer, M.J., Spliethoff, H., De Waal, C.M.W., Wilpshaar, S., Berkhout, B., Van Berlo, case study of seven Japanese metropolises. Waste Manage. Res. 31, 1110–1117.
M.A.J., Gohlke, O., Martin, J.J.E., 2011. High efficient waste-to-energy in Tang, L., Huang, H., Hao, H., Zhao, K., 2013. Development of plasma pyrolysis/
Amsterdam: getting ready for the next steps. Waste Manage. Res. 29, 20–29. gasification systems for energy efficient and environmentally sound waste
Nie, Y., 2008. Development and prospects of municipal solid waste (MSW) disposal. J. Electrostat. 71, 839–847.
incineration in China. Front. Environ. Sci. Eng. China 2, 1–7. Tanigaki, N., Manako, K., Osada, M., 2012. Co-gasification of municipal solid waste
Nippon Steel & Sumikin Engineering, 2014. <http://www.eng.nssmc.com/english/ and material recovery in a large-scale gasification and melting system. Waste
business/environment/e_01> (accessed September 2014). Manage. 32, 667–675.
Nixon, J.D., Dey, P.K., Ghosh, S.K., Davies, P.A., 2013a. Evaluation of options for Tanigaki, N., Fujinaga, Y., Kajiyama, H., Ishida, Y., 2013. Operating and
energy recovery from municipal solid waste in India using the hierarchical environmental performances of commercial-scale waste gasification and
analytical network process. Energy 59, 215–223. melting technology. Waste Manage. Res. 31, 1118–1124.
Nixon, J.D., Wright, D.G., Dey, P.K., Ghosh, S.K., Davies, P.A., 2013b. A comparative Tanner, V.R., 1965. Die Entwicklung der Von-Roll-MÃ1=4 llverbrennungsanlagen (The
assessment of waste incinerators in the UK. Waste Manage. 33, 2234–2244. development of the Von-Roll incinerators). Schweiz Bauzeitung 83, 251–260.
Otoma, S., Mori, Y., Terazono, A., Aso, T., Sameshima, R., 1997. Estimation of energy http://dx.doi.org/10.5169/seals-68135 (last accessed May 2014) (in German).
recovery and reduction of CO2 emissions in municipal solid waste power Taylor, R., Ray, R., Chapman, C., 2013. Advanced thermal treatment of auto shredder
generation. Resour. Conserv. Recy. 20, 95–117. residue and refuse derived fuel. Fuel 106, 401–409.
Ouadi, M., Brammer, J.G., Kay, M., Hornung, A., 2013. Fixed bed downdraft Tsai, W.T., 2010. Analysis of the sustainability of reusing industrial wastes as energy
gasification of paper industry wastes. Appl. Energy 103, 692–699. source in the industrial sector of Taiwan. J. Clean. Prod. 18, 1440–1445.
Palstra, S.W.L., Meijer, H.A.J., 2010. Carbon-14 based determination of the biogenic Tsiliyannis, C.A., 2013. Flue gas recirculation and enhanced performance of waste
fraction of industrial CO2 emissions – application and validation. Bioresour. incinerators under waste uncertainty. Environ. Sci. Technol. 47, 8051–8061.
Technol. 101, 3702–3710. Tyagi, V.K., Lo, S.L., 2013. Sludge: a waste or renewable source for energy and
Papageorgiou, A., Barton, J.R., Karagiannidis, A., 2009. Assessment of the greenhouse resources recovery? Renew. Sustain. Energy Rev. 25, 708–728.
effect impact of technologies used for energy recovery from municipal waste: a Udomsri, S., Martin, A.R., Fransson, T.H., 2010. Economic assessment and energy
case for England. J. Environ. Manage. 90, 2999–3012. model scenarios of municipal solid waste incineration and gas turbine hybrid
Pavlas, M., Bebar, L., Kropac, J., Stehlik, P., 2009. Waste to energy – an evaluation of dual-fueled cycles in Thailand. Waste Manage. 30, 1414–1422.
the environmental impact. Chem. Eng. Trans. 18, 671–676. Udomsri, S., Martin, A.R., Martin, V., 2011. Thermally driven cooling coupled with
Pavlas, M., Tous, M., Bebar, L., Stehlik, P., 2010. Waste to energy – an evaluation of municipal solid waste-fired power plant: application of combined heat, cooling
the environmental impact. Appl. Therm. Eng. 30, 2326–2332. and power in tropical urban areas. Appl. Energy 88, 1532–1542.
Pavlas, M., Tous, M., Klimek, P., Bebar, L., 2011. Waste incineration with production Unnikrishnan, S., Singh, A., 2010. Energy recovery in solid waste management
of clean and reliable energy. Clean Technol. Environ. Policy 13, 595–605. through CDM in India and other countries. Resour. Conserv. Recy. 54, 630–640.
Persson, K., Broström, M., Carlsson, J., Nordin, A., Backman, R., 2007. High US Department of Energy, 2007. Methodology for Allocating Municipal Solid Waste
temperature corrosion in a 65 MW waste to energy plant. Fuel Process. to Biogenic and Non-Biogenic Energy. <http://www.eia.gov/totalenergy/data/
Technol. 88, 1178–1182. monthly/pdf/historical/msw.pdf> (accessed May 2014).
Poggio, A., Grieco, E., 2010. Influence of flue gas cleaning system on the energetic Van Caneghem, J., Brems, A., Lievens, P., Block, C., Billen, P., Vermeulen, I., Dewil, R.,
efficiency and on the economic performance of a WTE plant. Waste Manage. 30, Baeyens, J., Vandecasteele, C., 2012. Fluidized bed waste incinerators: design,
1355–1361. operational and environmental issues. Prog. Energy Combust. 38, 551–582.
Poma, C., Verda, V., Consonni, S., 2010. Design and performance evaluation of a Vermeulen, I., Van Caneghem, J., Block, C., Baeyens, J., Vandecasteel, C., 2011.
waste-to-energy plant integrated with a combined cycle. Energy 35, Automotive shredder residue (ASR): reviewing its production from end-of-life
786–793. vehicles (ELVs) and its recycling, energy or chemicals’ valorization. J. Hazard.
Porteous, A., 2005. Why energy from waste incineration is an essential component Mater. 190, 8–27.
of environmentally responsible waste management. Waste Manage. 25, 451– Viganò, F., Consonni, S., Grosso, M., Rigamonti, L., 2010. Material and energy
459. recovery from Automotive Shredded Residues (ASR) via sequential gasification
Psomopoulos, C.S., Bourka, A., Themelis, N.J., 2009. Waste-to-energy: a review of the and combustion. Waste Manage. 30, 145–153.
status and benefits in USA. Waste Manage. 29, 1718–1724. Werle, S., Wilk, R.K., 2010. A review of methods for the thermal utilization of
Pyrolysis Facility, 2004. <http://lacitysan.org/solid_resources/strategic_programs/ sewage sludge: the Polish perspective. Renew. Energy 35, 1914–1919.
alternative_tech/technical_reports.htm> (accessed May 2014). Williams, P.T., 2013. Pyrolysis of waste tyres: a review. Waste Manage. 33, 1714–
Qiu, K., Hayden, A.C.S., 2009. Performance analysis and modeling of energy from 1728.
waste combined cycles. Appl. Therm. Eng. 29, 3049–3055. Yamada, S., Shimizu, M., Miyoshi, F., 2004. Thermoselect Waste Gasification and
Ray, R., Taylor, R., Chapman, C., 2012. The deployment of an advanced gasification Reforming Process. JFE Technical Report No. 3 (July 2004).
technology in the treatment of household and other waste streams. Process Saf. Yang, N., Zhang, H., Chen, M., Shao, L.M., He, P.H., 2012. Greenhouse gas emissions
Environ. 90, 213–220. from MSW incineration in China: impacts of waste characteristics and energy
Reimann, D.O., 2005. CEWEP Energy Report (Status 2001–2004). Updated July 2006. recovery. Waste Manage. 32, 2552–2560.
<http://www.cewep.eu/information/publicationsandstudies/studies/climate- Yassin, L., Lettieri, P., Simons, S.J.R., Germanà, A., 2009. Techno-economic
protection/223.CEWEP_Energy_Efficiency_Report_Status__-_.html> (accessed performance of energy-from-waste fluidized bed combustion and gasification
May 2014). processes in the UK context. Chem. Eng. J. 146, 315–327.
Reimann, D.O., 2009. CEWEP Energy Report II (Status 2004–2007). <http:// Zagaroli, M., 2007. La Centrale di Termogassificazione di Malagrotta. In Recupero di
www.cewep.eu/information/publicationsandstudies/studies/climate- energia e materia da rifiuti solidi i processi, le tecnologie, le esperienze, le
protection/360.CEWEP_Energy_Efficiency_Report_Status__-_.html> (accessed norme. In: Arena, U., Leone, U., Mastellone, M.L. (Eds.), AMRA S.c.a r.l. Napoli.
May 2014). <http://www.amracenter.comdocpubblicazioni_libri.htm> (accessed January
Reimann, D.O., 2012. CEWEP Energy Report III. (Status 2007–2010). <http:// 2014, in Italian).
www.cewep.eu/information/publicationsandstudies/studies/climate- Zhang, Q., Dor, L., Fenigshtein, D., Yang, W., Blasiak, W., 2012. Gasification of
protection/978.CEWEP_Energy_Efficiency_Report_Status_-.html> (last accessed municipal solid waste in the Plasma Gasification Melting process. Appl. Energy
May 2014). 90, 106–112.
Richers, U., Vehlow, J., Seifert, H., 1999. Evaluation Program for Municipal Solid Zhang, Q., Wu, Y., Dor, L., Yang, W., Blasiak, W., 2013. A thermodynamic analysis of
Waste Incineration Plants. FZKA 6298. Institut für Technische Chemie – Bereich solid waste gasification in the Plasma Gasification Melting process. Appl. Energy
thermische Abfallbehandlung. Forschungszentrum Karlsruhe GmbH, Karlsruhe. 112, 405–413.

Please cite this article in press as: Lombardi, L., et al. A review of technologies and performances of thermal treatment systems for energy recovery from
waste. Waste Management (2014), http://dx.doi.org/10.1016/j.wasman.2014.11.010

You might also like