Você está na página 1de 9

1 Article

2 Analysis of Coal-Cofiring with Hydrothermally-


3 Treated Empty Fruit Bunch and Palm Kernel Shell in
4 Drop Tube Furnace
5 Arif Darmawan1,2,*, Dwika Budianto2, Muhammad Aziz3 and Koji Tokimatsu1
6 1 Department of Transdisciplinary Science and Engineering, Tokyo Institute of Technology;
7 tokimatsu.k.ac@m.titech.ac.jp
8 2 Agency for the Assessment and Application of Technology (BPPT); dwikabudianto@gmail.com

9 3 Institute of Innovative Research, Tokyo Institute of Technology; aziz.m.aa@m.titech.ac.jp

10 * Correspondence: arif.d.aa@m.titech.ac.jp; Tel.: +81-808-191-8922

11 Academic Editor: name


12 Received: date; Accepted: date; Published: date

13 Abstract: Annual production of crude palm oil in Indonesia was about 28 Tg in 2015. Potential
14 biomass residues from palm oil industry such as palm kernel shell (PKS) and empty fruit bunch
15 (EFB) should be performed utilizing the appropriate technology to optimize its economic benefit as
16 well as minimize the environmental impacts. In this study, cofiring behaviors of each
17 hydrothermally-treated empty fruit bunch (HT-EFB) and PKS with coal in a drop tube furnace (DTF)
18 are analyzed in terms of thermal behavior including temperature distribution and produced gases
19 (CO2, CO, O2, NOx and SOx) composition through computational fluid dynamics (CFD). Several
20 different HT-EFB and PKS mass fractions are evaluated: 0%, 10%, 25%, and 50%. In general, HT-
21 EFB mass fractions of 10 to 25% seems to be the most preferable cofiring condition based on
22 temperature profile and produced gases composition. On the other hand, PKS mass fractions of 10
23 to 15 % shows a good combustion performance compared with other mass fractions. Moreover, PKS
24 and EFB supply capacity from palm mills surrounding the power plants is also an important
25 parameter to be considered for developing large scale biomass cofiring and integrated supply chain.

26 Keywords: coal-cofiring; hydrothermal treatment; empty fruit bunch; computational fluid


27 dynamics (CFD); drop tube furnace; simulation
28

29 1. Introduction
30 The demand for energy sustainability has encouraged researchers to study the use of renewable
31 energy sources in replacement of fossil fuel. In Indonesia, among numerous available energy sources,
32 biomasses including agricultural wastes play very important role in the energy matrix. Recently,
33 palm plantation is massively expanding due to high demand in palm oil products throughout the
34 globe [1]. According to data from the Indonesian Ministry of Agriculture, the total area of palm trees
35 plantations was around 8×104 km2 in 2015 or twice as much as in 2000 (4×104 km2). This number is
36 projected to increase to 1.3×105 km2 by 2020 [2]. Annual production of crude palm oil in Indonesia
37 was 27.78 Tg in 2013. This production is expected to reach 37 Tg in 2019 with annual growth of 4.59%
38 [3]. Palm oil production is mainly located in Sumatera (70%) and Kalimantan [4].
39 The massive increase of palm oil production has led to production of huge amount of
40 agricultural waste. It is assumed that about 90% of the whole palm tree has no significant utilization
41 [5]. It is estimated that the annual production of EFB and PKS are 24.82 and 7.67 Tg, respectively [3].
42 Huge amount of EFB and PKS are generated in Indonesia but poorly utilized, arising many problems
43 associated with the improper disposal practices of the palm oil waste.

Energies 2016, 9, x; doi: FOR PEER REVIEW www.mdpi.com/journal/energies


Energies 2016, 9, x FOR PEER REVIEW 2 of 9

44 Cofiring with coal has been identified as one of the least expensive and most efficient
45 technologies for converting these palm oil wastes to electricity. Unfortunately, cofiring generally
46 requires biofuels with a uniform quality and high energy density, which can be processed in the fuel
47 handling and combustion equipment of existing coal-fired power plants.
48 PKS Different to PKS, raw EFB has drawbacks of high moisture content, up to 70 wt% on wet
49 basis (wb), and low bulks density [5]. New techniques have also been studied to increase the cofiring
50 rates to desired levels for EFB. Hydrothermal treatment (HT) which is performed as a pretreatment
51 process prior to the thermo-chemical conversion of biomass offers significant merits such as high
52 conversion efficiency, the elimination of energy-extensive drying process, and relatively low
53 operation temperature compared to the other thermal processes [6,7].
54 Considering the high potential of both EFB and PKS, especially in Indonesia, the utilization of
55 these biomasses to be cofired into the existing or being planned coal-fired power plants becomes very
56 important. Unfortunately, to the best knowledge of the authors, there is almost no study deals with
57 the effort to evaluate the effect of hydrothermally-treated EFB (HT-EFB) and PKS cofiring to coal-
58 fired combustor. Therefore, in this study, coal-cofiring behaviors with each HT-EFB and PKS in a
59 drop tube furnace (DTF) are modeled and analyzed through computational fluid dynamics (CFD) in
60 terms of thermal behaviors including temperature profiles and composition of exhausted gases (CO 2,
61 CO, O2, NOx and SOx).

62 2. Coal-cofiring of HT-EFB and PKS


63 Figure 1 shows the conceptual diagram of integrated coal-cofiring system with HT-EFB and PKS
64 for power generation. Solid and dotted lines represent material and energy (electricity, heat) streams,
65 respectively. Raw EFB is shredded to smaller size initially before being hydrothermally treated.
66 Generally, HT is performed under subcritical region of water with relatively low temperature, lower
67 than 250 °C [8]. Some researchers have evaluated the application of HT on several biomasses to
68 produce hydrochar [9-11]. To eliminate the drying process after HT, continuous HT with temperature
69 of slightly higher than saturated one is adopted. Therefore, the moisture inside the EFB is also
70 evaporated and exhausted together with the steam required for HT. The produced HT-EFB is
71 exhausted from the reactor in relatively low moisture content.
72 PKS can be mixed directly after grinding without any pretreatment process including drying.
73 The moisture content of PKS is very low, about 3-5 wt% wb [12]. On the other hand, coal is ground
74 and dried initially to lower moisture content before being mixed with HT-EFB and PKS. The mixed
75 HT-EFB, PKS and coal is then cofired in the combustor producing high temperature heat for steam
76 generation using boiler. The generated steam then expands in the steam turbine generating the
77 electricity. In addition, the exhausted flue gas from boiler is utilized mainly in HT and coal drying.
78 This section may be divided by subheadings. It should provide a concise and precise description of
79 the experimental results, their interpretation as well as the experimental conclusions that can
80 be drawn.

81
82 Figure 1. Basic schematic diagram of HT-EFB and coal cofiring system.

83
84
Energies 2016, 9, x FOR PEER REVIEW 3 of 9

85 3. Numerical modeling and calculation

86 3.1. Schematic diagram of coal-cofiring with each HT-EFB and PKS


87 To observe the cofiring performance and its feasibility for each mixture of coal-HT-EFB and coal-
88 PKS, cofiring using DTF is modeled and evaluated in terms of temperature distribution and produced
89 gases composition. Figure 2 shows the schematic diagram of coal-cofiring system with each HT-EFB
90 and PKS in a DTF. The raw wet-coal is initially dried to certain low moisture content and ground to
91 achieve small and uniform size particles. Raw EFB is hydrothermally treated before being ground
92 and fed together with the dried coal to the DTF. Hydrothermal treatment is based on the experimental
93 results which were conducted at temperature of 493 K for 60 min [13]. On the other hand, PKS has
94 no pretreatment process except grinding to small and uniform size of particles.
95 In this study, four mass fractions of each HT-EFB and PKS to total mass of mixed fuel are
96 evaluated: 0% (100% coal), 10%, 25%, and 50%.
97

(a) (b)

98 Figure 2. Basic schematic diagram of coal cofiring system: (a) coal-HT-EFB, (b) coal-PKS

99 3.2. Drop Tube Furnace


100 Currently, in spite of any inherent difference of combustion behavior between DTF and realistic
101 combustion conditions, DTF is still considered as an effective equipment to evaluate certain
102 combustion behavior of fuel [14]. DTF is capable to facilitate an environment simulating the industrial
103 condition, such as fast heating rate, short residence time and dilute particle phases [15]. In general,
104 DTF consists of fuel feeding system, a reactor and a particle collection system at the outlet side. Each
105 HT-EFB and PKS is mixed initially with coal before being fed through the feeding system.
106 In simulation, a laboratory scale of DTF, which is a vertical tubular furnace, has a capacity of 1
107 kWth and dimensions of 1.5 m in height and 0.07 m in diameter is used. Detailed geometry of used
108 DTF is shown in Fig. 3. In addition, to facilitate an isothermal condition along the furnace, electric
109 heaters are installed and divided into three heating zones. The combustion process takes place inside
110 the tubular furnace and in downward direction.
111

112
113 Figure 3. Geometry of DTF used in simulation

114
Energies 2016, 9, x FOR PEER REVIEW 4 of 9

115 Combustion air supply is divided into primary and secondary air with volumetric flow rates of
116 3 and 4 L min-1, respectively. The former is utilized to feed the fuels to combustor, while the latter
117 covers the rest of the demand for combustion. Mixed fuel particles are fed at 45–60 kg h-1 through
118 injection probe mounted at the top of DTF. The produced gas is exhausted from the bottom of DTF.

119 3.3. Materials


120 The coal used in the simulation is originated from Kalimantan, Indonesia. This coal is classified
121 as low rank coal having high moisture content and lower calorific value. Table 1 shows the
122 compositions of used coal, HT-EFB and PKS including proximate and ultimate analyses.

123 Table 1. Material composition

Component Coal [16] HT-EFB [13] PKS [16]


Proximate analysis
Fixed Carbon 40.23 28.62 24.35
Volatile Matter 41.57 62.57 66.77
Moisture 17.30 3.00 3.86
Ash 0.90 5.82 5.02
Ultimate analysis
C 56.98 52.74 43.77
H 3.69 5.33 5.85
O 18.13 31.95 42.32
N 2.83 0.85 0.89
S 0.17 0.00 0.00
Calorific value (MJ kg-1) 22.34 22.22 17.68

124 3.3. Computational Modeling


125 CFD is an effective tool to model and calculate the fluid flow, heat and mass transfers, chemical
126 reaction, and solid and fluid interactions [17]. CFD modelling method for combustion of biomass
127 particle is a challenging work. There have been still a limited number of numerical analysis of
128 biomass cofiring using CFD models especially for HT-EFB cofiring. Compared to physical
129 investigation through experiment, CFD modelling is significantly time and cost effective, safe and
130 easy for scaling up, hence it is usually adopted before performing an experimental study. Related to
131 co-firing, CFD analysis is expected to be able to clarify the combustion performance for all stages of
132 the combustion including combustion temperature, kinetics behavior and concentration of the
133 produced gases.
134 In this simulation, a commercial CFD software ANSYS DesignModeler and Fluent ver. 16.2
135 (ANSYS Inc.) are used to build 3D combustor model and analyze the cofiring behavior, respectively.
136 Cofiring simulation includes some considerations of governing dynamics equations (mass
137 conservation, momentum and enthalpy), turbulences (k-ε turbulence model), radiation heat transfer
138 (P-1 model), particle and reaction in both particle (Eulerian-Lagrangian model) and gas (global 2-
139 steps reactions) phases [18]. Some additional boundary conditions include: (1) fuel and air inlet flow
140 rates are 1.38×10-5 kg s-1 and 1.6 ×10-4 kg s-1 at 300 K, (2) furnace wall temperature, wall roughness
141 and internal emissivity are set to 1300 K (isothermal), 0.5 and 1, respectively, and (3) feeding wall is
142 considered isothermal at 300 K.

143 Reaction mechanism of HT-EFB cofiring

144 Coal char

C + 0.5 O2 → CO (1)
Energies 2016, 9, x FOR PEER REVIEW 5 of 9

145 HT-EFB char

C + 0.5 O2 → CO (2)

146 Coal volatile matter

CH0.39O0.24N0.043S0.0011 + 0.48O2 → CO + 0.195H2O + 0.0011SO2+ 0.022N2 (3)

147 HT-EFB volatile matter

CH1.21O0.45N0.014 + 0.5775O2 → CO + 0.605H2O + 0.007N2 (4)

CO + 0.5 O2→ CO2 (5)

148 Reaction mechanism of PKS cofiring

149 Coal char

C + 0.5 O2 → CO (6)

150 PKS char

C + 0.5 O2 → CO (7)

151 Coal volatile matter

CH0.39O0.24N0.043S0.0011 + 0.48O2 → CO + 0.195H2O + 0.0011SO2+ 0.022N2 (8)

152 PKS volatile matter

(CH0.8O0.725 N0.0174) + 0.3375O2→ CO + 0.4H2O + 0.0087N2 (9)

CO + 0.5 O2→ CO2 (10)

153 5. Results and Discussion


154 5.1. Temperature Distribution
155 An HT-EFB after hydrothermal treatment was found become more uniform and coal-like.
156 Hydrothermal treatment also could improve the drying and dehydration performance, thus the
157 moisture content of the HT-EFB decreased to approximately 3%. This characteristic is very important
158 in the combustion system. For PKS, since it has lower moisture content than EFB, it can be burned
159 with coal without any pre-treatment except crushing. Figure 3 and 4 show the temperature
160 distribution along the combustor under different cofiring mass fractions. In general, higher HT-EFB
161 and PKS mass fraction will increase the temperature inside the combustor. HT-EFB mass fractions of
162 50% results in the highest outlet temperature (maximum of 1536 K).
163
Energies 2016, 9, x FOR PEER REVIEW 6 of 9

(a) (b)
164 Figure 4. Temperature distribution inside the combustor: (a) HT-EFB cofiring (b) PKS cofiring

165 For comparison, the highest outlet temperature in case of no HT-EFB is 1347 K. The change in
166 HT-EFB mass fraction leads to the change in heating value of the mixture, therefore the flame shape
167 and temperature profile within the combustor changes accordingly. HT-EFB contains higher volatile
168 matter and lower fixed carbon than coal. The location of the high-temperature region corresponds to
169 the combustion location of volatile matter and the oxygen availability. The devolatilization of HT-
170 EFB particles occurs faster and earlier than the coal particles because the co-fired HT-EFB has lower
171 moisture content than the coal. On the other hand, since coal has higher significant water content, its
172 particles require longer time for drying before being devolatilized. As the result, in high HT-EFB
173 mass fraction, the combustion temperature remains high and distributed more uniformly although it
174 is located in lower part of the combustor. In addition, as HT-EFB has lower moisture content than
175 coal, high HT-EFB mass fraction leads to the lower total moisture content of the mixed fuel of HT-
176 EFB and coal in the combustor system. Finally, this condition impacts the combustion temperature
177 as water has relatively high heat capacity. Compared to the main combustion area, a lower
178 temperature is observed in the upper the combustor. This is due to heating process of the mixed fuel.
179 The mixture is pyrolyzed and then evolves as volatile matter.
180 In general, higher PKS mass fraction leads to higher flame temperature inside the combustor as
181 well as exhaust gases flowing to the superheater for energy recovery. As can be observed from Table
182 1, PKS has higher volatile matter content than coal. Therefore, higher PKS mass fraction results in
183 higher total volatile matter content during cofiring.
184
185 5.2. Distribution of produced CO, CO2, SO2 and NO
186
187 In contrast to the fully coal combustion, HT-EFB and PKS cofiring has increased carbon
188 monoxide concentrations and nitrogen monoxide in the combustion. Figs. 5-8 present further
189 information about the concentration of produced gases during cofiring, including CO, CO 2, NO, and
190 SO2. Regarding the produced CO concentration, higher mass fraction of HT-EFB and PKS leads to
191 the increase of CO mass fraction during initial reaction of combustion. The volatile matter, especially
192 from HT-EFB and PKS, is oxidized under high combustion temperature forming CO. Afterward, CO
193 reacts further with O2 (air) along the combustor forming CO2. In addition, coal cofiring with HT-EFB
194 and PKS results in lower CO2 concentration following the increase of both HT-EFB and PKS mass
195 fraction (Fig. 6).
196 On the other hand, the formation of NO is considered to be dominated by thermal NO due to
197 high combustion temperature. Higher HT-EFB and PKS mass fraction results in significantly higher
198 NO concentration, especially mass fraction of 50% (Fig 7). In addition, SO 2 compound is formed due
199 to interaction between Sulphur with air. Since HT-EFB and PKS contain no Sulphur (Table 1), higher
200 proposition of HT-EFB and PKS will impact in reduction of SO2 formation (Fig. 8.)
201
Energies 2016, 9, x FOR PEER REVIEW 7 of 9

202

(a) (b)
203 Figure 5. CO mass fraction along the height of the combustor for each different mass fraction:
204 (a) HT-EFB cofiring (b) PKS cofiring

205

(a) (b)
206 Figure 6. CO2 mass fraction along the height of the combustor for each different mass fraction:
207 (a) HT-EFB cofiring (b) PKS cofiring

(a) (b)
208 Figure 7. NO mass fraction along the height of the combustor for each different mass fraction:
209 (a) HT-EFB cofiring (b) PKS cofiring
Energies 2016, 9, x FOR PEER REVIEW 8 of 9

(a) (b)
210 Figure 8. SO2 mass fraction along the height of the combustor for each different mass fraction:
211 (a) HT-EFB cofiring (b) PKS cofiring

212 6. Results and Discussion

213 Cofiring of behavior of coal with HT-EFB and coal with PKS has been modelled and evaluated using CFD
214 analysis under different mixing mass fractions. In general, HT-EFB mass fraction PKS mass fraction of 10 to 25%
215 seems to be the most preferable cofiring condition in terms of temperature and produced gas compositions.

216
217 Acknowledgments: The author greatly appreciate the support of The Indonesia Endowment Fund for Education
218 (LPDP). The proximate and ultimate analyses of PKS and coal are provided by Agency for the Assessment and
219 Application of Technology (BPPT), Indonesia. The authors also thank to Annisa Nurdiawati (Tokyo Institute of
220 Technology) and M. Kunta Biddinika (Tokyo Institute of Technology).

221 Author Contributions: D.B. modelled and performed the CFD analysis for co-firing; M.A. and A.D. performed
222 overall research plan and results analysis; M.A. and A.D. wrote the manuscript; K.T. analyzed the results and
223 checked the manuscript.

224 Conflicts of Interest: The authors declare no conflict of interest.

225 References
226 1. Aziz, M.; Oda, T.; Kashiwagi, T. Innovative steam drying of empty fruit bunch with high energy efficiency.
227 Drying Technol., 2015, 33, 395-405.
228 2. Tarigan, S.D.; Widyaliza, S. Expansion of Oil Palm Plantations and Forest Cover Changes in Bungo and
229 Merangin Districts, Jambi Province, Indonesia. Procedia Environ. Sci., 2015, 24, 199-205.
230 3. BPS. Statistik Kelapa Sawit Indonesia. Badan Pus. Stat. Indonesia., 2014 pp. 96.
231 4. Aziz, M.; Oda, T.; Kashiwagi, T. Design and analysis of energy-efficient integrated crude palm oil and palm
232 kernel oil processes. J Jpn. Inst Energy., 2015, 94, 143- 150.
233 5. Aziz, M.; Prawisudha, P.; Prabowo, B.; Budiman, B.A. Integration of energy-efficient empty fruit bunch
234 drying with gasification/combined cycle systems. App Energy.,2015, 139, 188-195.
235 6. Liu, Y.; Aziz, M.; Kansha, Y.; Tsutsumi, A. A novel exergy recuperative drying module and its application
236 for energy-saving drying with superheated steam. Chem Eng Sci., 2013, 100, 392-401.
237 7. Parshetti, G.K.; Hoekman, S.K.; Balasubramanian, R. Chemical, structural and combustion characteristics
238 of carbonaceous products obtained by hydrothermal carbonization of palm empty fruit bunches.
239 Bioresource Technol., 2013, 135, 683-689.
240 8. Hu B, Yu SH, Wang K, Liu L, Xu XW. Functional carbonaceous materials from hydrothermal carbonization
241 of biomass: an effective chemical process. Dalton Trans., 2008, 40, 5414 – 5423.
242 9. Jamari, SS.; Howse, JR. The effect of the hydrothermal carbonization process on palm oil empty fruit bunch.
243 Biomass Bioenergy. 2012., 47, 82–90
244 10. Erlach, B.; Harder, B.; Tsatsaronis, G. Combined hydrothermal carbonization and gasification of biomass
245 with carbon capture. Energy. 2012., 45, 329-338.
Energies 2016, 9, x FOR PEER REVIEW 9 of 9

246 11. R. Roman S.; Nabais JMV.; Laginhas C.; Ledesma B; Gonzales JF. Hydrothermal carbonization as an
247 effective way of densifying the energy content of biomass. Fuel Proc. Tech., 2012, 103, 78 – 83.
248 12. Bazargan, A.; Rough S.L.; McKay, G. Compaction of palm kernel shell biochars for application as solid fuel.
249 Biomass Bioenergy. 2014, 70, 489-497
250 13. Novianti, S.; Biddinika, M. K.; Prawisudha, P.; Yoshikawa, K. Upgrading of Palm Oil Empty Fruit Bunch
251 Employing Hydrothermal Treatment in Lab-scale and Pilot Scale. Procedia Environ. Sci, 2014, 20, 46–54.
252 14. Du, S.W.; Chen, W.H.; Lucas, J.A. Pulverized coal burnout in blast furnace simulated by a drop tube
253 furnace. Energy., 2010, 35, 576-581.
254 15. Manquais, K.L.; Snape, C.; McRobbie, I.; Barker, J.; Pellegrini, V. Comparison of the combustion reactivity
255 of TGA and drop tube furnace chars from a bituminous coal. Energy Fuels, 2009, 23, 4269-4277.
256 16. Aziz, M.; Budianto, D.; Oda, T. Co-firing of palm kernel shell into coal fired power plant. Energies, 2016, 9,
257 137.
258 17. Ranade, V.V.; Gupta, D.F. Computational Modeling of Pulverized Coal Fired Boilers: CRC Press: Boca
259 Raton, USA, 2015, pp. 19-31.
260 18. Budianto, D.; Aziz, M.; Cahyadi; Oda, T. Numerical Investigation of Co-Firing of Palm Kernel Shell into
261 Pulverized Coal Combustion. J. Jpn. Inst. Energy, 2016, 95, 605-614.

262 © 2016 by the authors. Submitted for possible open access publication under the
263 terms and conditions of the Creative Commons Attribution (CC-BY) license
264 (http://creativecommons.org/licenses/by/4.0/).

Você também pode gostar