Sunteți pe pagina 1din 6

Available online at www.sciencedirect.

com

ScienceDirect
Availableonline
Available onlineatatwww.sciencedirect.com
www.sciencedirect.com
Energy Procedia 00 (2017) 000–000

ScienceDirect
ScienceDirect
www.elsevier.com/locate/procedia

Energy
EnergyProcedia 142
Procedia 00(2017)
(2017)3480–3485
000–000
www.elsevier.com/locate/procedia

9th International Conference on Applied Energy, ICAE2017, 21-24 August 2017, Cardiff, UK

A comparison of energy recovery from MSW through plasma


The 15th International Symposium on District Heating and Cooling
gasification and entrained flow gasification
Assessing
Luca Mazzoni1the feasibility
, Manar Almazrouei of1,using
Chaouki the heat2, demand-outdoor
Ghenai and Isam Janajreh1*
temperature function for a long-term district heat demand forecast
Khalifa University of Science and Technology, Masdar Institute, Mechanical Engineering Dept. Abu Dhabi 54224, UAE
1

2
Sustainable and Renewable Energy Engineering Dept, University of Sharjah, Sharjah 27272, UAE
I. Andrića,b,c*, A. Pinaa, P. Ferrãoa, J. Fournierb., B. Lacarrièrec, O. Le Correc
Abstract
a
IN+ Center for Innovation, Technology and Policy Research - Instituto Superior Técnico, Av. Rovisco Pais 1, 1049-001 Lisbon, Portugal
b
Veolia Recherche & Innovation, 291 Avenue Dreyfous Daniel, 78520 Limay, France
c
This work presentsDépartement Systèmes between
a comparison Énergétiques
twoetthermochemical
Environnement - conversion
IMT Atlantique, 4 rue Alfred
pathways Kastler,recovery
for energy 44300 Nantes,
fromFrance
waste, entrained
flow, and plasma gasification. The software Aspen Plus was used to build two equilibrium models to simulate the behavior of the
two gasification processes when the feedstock consists of hand sorted municipal solid waste (MSW) with medium-high heating
value. The impact of various level of air oxygen enrichment is investigated by comparing side-to-side the performance of the two
Abstractin terms of syngas composition, syngas lower heating value and cold gas efficiency (CGE). As the results suggest, plasma
processes
gasification, with a maximum CGE of 74.8% can be considered a better option than entrained flow gasification with a CGE of
District heating networks are commonly addressed in the literature as one of the most effective solutions for decreasing the
71.6%.
greenhouse gas emissions from the building sector. These systems require high investments which are returned through the heat
©sales.
2017 Due to the changed
The Authors. Published climate conditions
by Elsevier Ltd. and building renovation policies, heat demand in the future could decrease,
prolonging the investment return period.
Peer-review under responsibility of the scientific committee of the 9th International Conference on Applied Energy.
The main scope of this paper is to assess the feasibility of using the heat demand – outdoor temperature function for heat demand
forecast. Plasma
Keywords: The district of Alvalade,
gasification; located
entrained flow in Lisbon
gasification; cold(Portugal), was
gas efficiency; used assolid
municipal a case
waste;study.
Aspen The district is consisted of 665
Plus modeling.
buildings that vary in both construction period and typology. Three weather scenarios (low, medium, high) and three district
renovation scenarios were developed (shallow, intermediate, deep). To estimate the error, obtained heat demand values were
1.compared with results from a dynamic heat demand model, previously developed and validated by the authors.
Introduction
The results showed that when only weather change is considered, the margin of error could be acceptable for some applications
(the error in annual demand was lower than 20% for all weather scenarios considered). However, after introducing renovation
One of the main challenges of today’s municipalities is to find an effective solution to mitigate the problem of solid
scenarios, the error value increased up to 59.5% (depending on the weather and renovation scenarios combination considered).
waste disposal.
The value of slopeWhen looking
coefficient at theonfield
increased averageof within
waste the management
range of 3.8% thatup deals
to 8% with energythatrecovery
per decade, through
corresponds to the
thermochemical conversion,
decrease in the number several
of heating options
hours are available.
of 22-139h during the However, the most
heating season widespread
(depending on thepathway is incineration,
combination of weatherthat
and
isrenovation
based on scenarios
mass burning of waste
considered). Oninthegrate-fired
other hand,furnaces
function[1]. Someincreased
intercept Europeanforcountries,
7.8-12.7%like per Sweden, Denmark,onand
decade (depending the
the Netherlands
coupled have
scenarios). almost
The valuesreached
suggesteda zero
couldwaste to landfill
be used to modify strategy, with aparameters
the function consistentfor presence of waste
the scenarios to energy,
considered, and
improve the accuracy of heat demand estimations.

© 2017 The Authors. Published by Elsevier Ltd.


Peer-review under responsibility of the Scientific Committee of The 15th International Symposium on District Heating and
* Corresponding author. Tel.: +97128109130; fax: +97128109901.
Cooling.
E-mail address: ijanajreh@masdar.ac.ae
Keywords: Heat demand; Forecast; Climate change
1876-6102 © 2017 The Authors. Published by Elsevier Ltd.
Peer-review under responsibility of the scientific committee of the 9th International Conference on Applied Energy.

1876-6102 © 2017 The Authors. Published by Elsevier Ltd.


Peer-review under responsibility of the Scientific Committee of The 15th International Symposium on District Heating and Cooling.
1876-6102 © 2017 The Authors. Published by Elsevier Ltd.
Peer-review under responsibility of the scientific committee of the 9th International Conference on Applied Energy.
10.1016/j.egypro.2017.12.233
Luca Mazzoni et al. / Energy Procedia 142 (2017) 3480–3485 3481
2 Mazooni, Ghenai, and Janajreh / Energy Procedia 00 (2017) 000–000

with Denmark incinerating around 50% of its total waste [2]. Aside from incineration, pyrolysis and gasification are
also applied in different configuration for waste to energy purposes [3-5]. Gasification is especially interesting, since
the waste is converted into syngas (mainly CO and H2) which can then be burned in gas engines or gas turbines [2],
yielding higher efficiency than incineration, or it can be converted to liquid fuels thus allowing a more flexible end-
use [6].Among the different types of gasification reactors, entrained flow gasifiers, are commercially well established
for coal and biomass gasification [7-10]. However, for waste to energy applications, various authors, identified plasma
gasification as a superior technology [3, 11, 12]. Its distinctive features are the high-energy density and temperature
of thermal plasma which yields fast reaction times and higher flexibility on the waste that it can process [12]. A
commercial installation in Japan demonstrated some practical challenges and limitations of this technology [13].
Various modeling approaches can be adopted to study gasification processes. Among these, equilibrium modeling
is well studied in the literature for its ability to capture the main trend in the gasifiers operation and syngas composition
without being bounded to a specific design thus providing a preliminary indication of the process performance.
Furthermore, its application to model entrained flow and plasma gasification is justified by their high operating
temperature resulting in high reaction rates [14, 15]. Different types of entrained flow coal gasifiers are successfully
modeled with equilibrium models using Aspen Plus with reported cold gas efficiencies ranging from 72% for a wet
feed to 83% for a dry feed [16]. For plasma gasification of solid waste, various equilibrium models have been proposed
[17-19], including one based on Aspen Plus which reported a plasma gasification efficiency of 69% [20]. A preliminary
comparison between plasma and air gasification of different materials was proposed by [21]. However, no studies have
compared entrained flow gasification with plasma gasification when MSW is considered as a feedstock. In this study,
the two processes are compared by varying the amount of oxygen introduced into the reactor as it is one of the main
parameter affecting the gasification efficiency and the syngas quality.

2. Methodology

2.1. Material characterization

Typically, MSW exhibits a very high variability regarding composition, depending on the geographical location,
and the local economy. This latter factor determines the amount of high heating value combustible leftovers such as
plastic, paper, rubber, cloth, leather, and wood. However, a representative MSW composition in terms of proximate
and ultimate analysis is taken from the literature [22] and reported in Table 1. The HHV of MSW is 17.57 MJ/kg [22]
whereas the calculated lower heating value (LHV) is 16.42 MJ/kg. The considered MSW composition here considered
agrees well with other literature data as well [4, 23].

Table 1. MSW proximate analysis as received (ar) and ultimate analysis on a dry and ash free basis (daf).

Proximate analysis (ar) Weight fraction (%) Ultimate analysis (daf) Weight fraction (%)
Moisture 7.56 Carbon 59.64
Volatile matter 53.61 Hydrogen 6.37
Fixed carbon 22.38 Nitrogen 1.50
Ash 16.45 Sulphur 0.37
Oxygen 32.12

2.2. Entrained flow gasification model

An entrained flow gasification model was developed with Aspen Plus and is based on a representative Shell
entrained flow gasifier operating at 40 bar [24]. It is assumed that the gasifier is adiabatic in addition to uniform
temperature and perfect mixing inside the reactor. The adoption of equilibrium modeling is justified by the high
operating temperature, which is set at 1,300C. As shown in Fig. 1 the waste feed is entered in a RYield reactor as
stream 1 where the waste is modeled as a non-conventional solid where it is decomposed according to its ultimate
analysis. The equilibrium is calculated using an RGibbs reactor which determines the product species concentration
3482 Luca Mazzoni et al. / Energy Procedia 142 (2017) 3480–3485
Mazooni, Ghenai, and Janajreh / Energy Procedia 00 (2017) 000–000 3

at the exit of the gasifier by using the Gibbs free energy minimization approach. The species considered are N2, O2,
H2, CO, CH4, C2H2, CO2, H2O, HCL, H2S, COS, HCN, NH3, NO2, NO, S, SO2, SO3, Cl2, and solid carbon. For a fixed
waste feed the amount of air, stream 6, and oxygen, stream 7, introduced in the reactor is calculated by a design
specification to have the RGIBBS reactor operating at 1,300C. The cleaned syngas then exits the gasifier as stream
4 after separation, in block SEP1, of the inorganic fraction (stream 5).

Fig. 1. Aspen Plus flowsheet of the entrained flow gasifier’s model.

2.3. Plasma gasification model

The Aspen Plus model developed for the plasma gasifier is based on [20]. As for the entrained flow gasifier, the
waste is modeled as a non-conventional solid and is represented by stream 1 in Fig. 2. The heating of the waste as it
enters the reactor is modeled with a couple of heat exchanger blocks, H1 and H2 and the heat stream HEAT2. The
non-conventional solid material, stream 2 is then decomposed in the RYIELD reactor similarly as in the model for the
entrained flow gasifier. Then stream 3 enters the separator block SEP1, where 85% of the moisture is diverted as
stream 14. The decomposed solid waste is then sent as stream 4 into the HTR block, which is an RGibbs reactor
simulating the high-temperature zone of the gasifier where the solid feed comes in direct contact with the thermal
plasma, stream 13. The reactor assumes the same species considered for the RGIBBS reactor in the entrained
gasification model. It is assumed that thermal plasma is generated with a DC non-transferred plasma torch modeled
as a heat exchanger PLTORCH. This block heats up the plasma forming gas, stream 12, up to 4,000C. The plasma
forming gas results from the mixing, through block MIX1 of air and oxygen, stream 10 and 11, respectively which
are determined by a design specification which sets the temperature of the stream 5 to 2,500C. The inorganic fraction
of the waste is separated from stream 6 as vitrified slag, through block SEP2. Then, stream 7 enters a second RGibbs
reactor, LTR block, which models the low-temperature zone of the gasifier where the syngas formation is completed.
The syngas exiting the LTR block as stream 8 proceed to the block MIX2 where it is mixed with the moisture, stream
14. The temperature of the syngas exiting the gasifier as stream 9 is between 1,250C and 1,300C.

Fig. 2. Aspen Plus flowsheet of the plasma gasifier’s model.


Luca Mazzoni et al. / Energy Procedia 142 (2017) 3480–3485 3483
4 Mazooni, Ghenai, and Janajreh / Energy Procedia 00 (2017) 000–000

The model has been validated using the experimental data from a recent study on the plasma gasification of refused
derived fuel [25]. The syngas composition of the experimental study is compared with the one predicted by the
developed model in Table 2. The syngas composition predicted by the model agrees well with the experimental one,
apart from an under-prediction of the methane content, which is typical for equilibrium-based models.

Table 2. Syngas composition resulting from refused derived fuel plasma gasification. Comparison between simulation and experimental.

Mole fraction (dry vol%) CO H2 CO2 CH4


Current work 32.4 60.6 7.0 0.0
Agon et al. [25] 33.0 58.4 4.2 4.4

2.4. Oxygen ratio

In this work, the amount of oxygen introduced in the entrained flow gasifier and the plasma forming gas is varied
to determine its effect on the process efficiency. Thus, for convenience, the performance metrics discussed hereafter
are examined for different oxygen ratios which are calculated according to Eq. 1.

ṁO2
ω= (1)
ṁO2 + ṁair

Where ṁO2 and ṁair are the mass flow rate of oxygen and air respectively.

3. Results and discussion

The two gasification processes are compared by looking at the syngas composition, syngas lower heating value
(LHV) and cold gas efficiency (CGE). The syngas composition in terms of CO, H2, N2, CO2, and H2O for the entrained
flow and plasma gasification as the oxygen ratio is varied is depicted in Fig. 3. It is worth noting that in the case of
entrained flow gasification ω need to be at least 21% to reach the specified temperature of 1,300C in the reactor.

Fig. 3. Syngas composition comparison between entrained flow (solid lines) and plasma gasification (dashed lines).

For both gasification processes when the oxygen ratio is increased from zero to one an increasing trend is evident
in the mole fractions of the combustible species H2 and CO as a result of a less diluted syngas, since the N2 mole
fraction concurrently decreases. Overall, it is evident that plasma gasification produces a syngas with sensibly higher
3484 Luca Mazzoni et al. / Energy Procedia 142 (2017) 3480–3485
Mazooni, Ghenai, and Janajreh / Energy Procedia 00 (2017) 000–000 5

H2 and CO mole fractions compared to entrained flow gasification. When the oxygen ratio is one, the syngas
composition in the case of plasma gasification is 51% vol. CO and 29% vol. H2 while that of entrained flow
gasification is 43% vol. CO and 22% vol. H2. On the other hand, the products of combustion reactions like CO2 and
H2O are higher in the case of entrained flow gasification as one would expect from an allothermal process such as
entrained flow gasification. Also, N2 is higher in the case of entrained flow gasification due to a higher air requirement
to reach the specified temperature in the reactor. The syngas composition directly affects its lower heating value
(LHV) which is shown in Fig.4 (a) for both gasification processes as the oxygen ratio is varied from zero to one. It is
evident that plasma gasification yields a syngas with a higher LHV, ranging from 7.6 MJ/kg to 10.7 MJ/kg than that
of entrained flow gasification which ranges from 3.5 MJ/kg to 7.8 MJ/kg.The process efficiency was also assessed
using the CGE defined as the ratio between the power output and the power input to gasifier as per Eq. 2.

ṁsyngas ∙ LHVsyngas
CGE = (2)
ṁfeed ∙ LHVfeed + Ẇtorch

Where ṁsyngas and ṁfeed are the mass flow rate of the syngas and the waste feed respectively, while LHVsyngas
and LHVfeed are the lower heating value of the syngas and the waste feed respectively. In the case of plasma
gasification, the power required to produce the thermal plasma is also taken into consideration with the term Ẇtorch .
As can be seen in Fig. 4 (b) the CGE of plasma gasification is higher than that of entrained flow gasification. It ranges
from 65.9% to 74.8% for the former process and from 60.4% to 71.6% for the latter one.

Fig. 4. Comparison between entrained flow and plasma gasification for syngas LHV (a) and CGE (b).

4. Conclusion

In this paper, we compared entrained flow and plasma gasification when MSW is used as a feedstock, and different
levels of oxygen enrichment of air are considered. It was evident that plasma gasification performed better than
entrained flow gasification, yielding a higher calorific syngas with higher mole fractions of CO and H 2. Overall, the
CGE was also higher in the case of plasma gasification thus meaning that the energy required to create the thermal
plasma did not negatively affect the overall process efficiency. For both processes, it is evident that increasing the
oxygen ratio introduced in the reactor is highly beneficial increasing significantly all performance metrics.

Acknowledgements

The authors acknowledge the partial support of the Takeer Research Center (TRC) and Tadweer in Abu Dhabi.
Luca Mazzoni et al. / Energy Procedia 142 (2017) 3480–3485 3485
6 Mazooni, Ghenai, and Janajreh / Energy Procedia 00 (2017) 000–000

References

[1] Leckner B. Process aspects in combustion and gasification Waste-to-Energy (WtE) units. Waste Management.
2015;37:13-25.
[2] Yassin L, Lettieri P, Simons SJ, Germanà A. Techno-economic performance of energy-from-waste fluidized bed
combustion and gasification processes in the UK context. Chemical Engineering Journal. 2009;146:315-27.
[3] Arena U. Process and technological aspects of municipal solid waste gasification. A review. Waste management.
2012;32:625-39.
[4] Gang X, Jin B-s, ZHONG Z-p, Yong C, NI M-j, CEN K-f, et al. Experimental study on MSW gasification and
melting technology. Journal of Environmental Sciences. 2007;19:1398-403.
[5] Klinghoffer N, Castaldi M. Gasification and pyrolysis of municipal solid waste (MSW). Waste to Energy
Conversion Technology. 2013:146-76.
[6] Trippe F, Fröhling M, Schultmann F, Stahl R, Henrich E. Techno-economic assessment of gasification as a process
step within biomass-to-liquid (BtL) fuel and chemicals production. Fuel Processing Technology. 2011;92:2169-84.
[7] Guo X, Dai Z, Gong X, Chen X, Liu H, Wang F, et al. Performance of an entrained-flow gasification technology
of pulverized coal in pilot-scale plant. Fuel processing technology. 2007;88:451-9.
[8] Svoboda K, Pohořelý M, Hartman M, Martinec J. Pretreatment and feeding of biomass for pressurized entrained
flow gasification. Fuel Processing Technology. 2009;90:629-35.
[9] Van der Drift A, Boerrigter H, Coda B, Cieplik M, Hemmes K, Van Ree R, et al. Entrained flow gasification of
biomass. Ash behaviour, feeding issues, and system analyses. 2004.
[10] Watanabe H, Otaka M. Numerical simulation of coal gasification in entrained flow coal gasifier. Fuel.
2006;85:1935-43.
[11] Fabry F, Rehmet C, Rohani V, Fulcheri L. Waste gasification by thermal plasma: a review. Waste and Biomass
Valorization. 2013;4:421-39.
[12] Gomez E, Rani DA, Cheeseman C, Deegan D, Wise M, Boccaccini A. Thermal plasma technology for the
treatment of wastes: a critical review. Journal of Hazardous Materials. 2009;161:614-26.
[13] Willis KP, Osada S, Willerton KL. Plasma gasification: lessons learned at Eco-Valley WTE facility. 18th Annual
North American Waste-to-Energy Conference: American Society of Mechanical Engineers; 2010. p. 133-40.
[14] Prins MJ, Ptasinski KJ, Janssen FJ. From coal to biomass gasification: comparison of thermodynamic efficiency.
Energy. 2007;32:1248-59.
[15] Wang Z, Yang J, Li Z, Xiang Y. Syngas composition study. Frontiers of Energy and Power Engineering in China.
2009;3:369-72.
[16] Kunze C, Spliethoff H. Modelling, comparison and operation experiences of entrained flow gasifier. Energy
Conversion and management. 2011;52:2135-41.
[17] Mountouris A, Voutsas E, Tassios D. Solid waste plasma gasification: equilibrium model development and
exergy analysis. Energy Conversion and Management. 2006;47:1723-37.
[18] Mountouris A, Voutsas E, Tassios D. Plasma gasification of sewage sludge: Process development and energy
optimization. Energy Conversion and Management. 2008;49:2264-71.
[19] Zhang Q, Wu Y, Dor L, Yang W, Blasiak W. A thermodynamic analysis of solid waste gasification in the Plasma
Gasification Melting process. Applied Energy. 2013;112:405-13.
[20] Minutillo M, Perna A, Di Bona D. Modelling and performance analysis of an integrated plasma gasification
combined cycle (IPGCC) power plant. Energy Conversion and Management. 2009;50:2837-42.
[21] Janajreh I, Raza SS, Valmundsson AS. Plasma gasification process: Modeling, simulation and comparison with
conventional air gasification. Energy conversion and management. 2013;65:801-9.
[22] Adeyemi I, Janajreh I. Gasification of two untapped resources: El-lajjun oil shale and municipal solid waste.
International Journal Of Modern Engineering.19:61.
[23] Kathirvale S, Yunus MNM, Sopian K, Samsuddin AH. Energy potential from municipal solid waste in Malaysia.
Renewable energy. 2004;29:559-67.
[24] Cormos C-C. Integrated assessment of IGCC power generation technology with carbon capture and storage
(CCS). Energy. 2012;42:434-45.
[25] Agon N, Hrabovský M, Chumak O, Hlína M, Kopecký V, Mas̆láni A, et al. Plasma gasification of refuse derived
fuel in a single-stage system using different gasifying agents. Waste Management. 2016;47:246-55.

S-ar putea să vă placă și