Sei sulla pagina 1di 289

Renewable, Local Electricity Generation from

Palm Oil Mill Residues: A Case Study from


Peninsular Malaysia

By
Ida Fahani Md Jaye

Submitted in fulfilment of the requirements for the


degree of Doctor of Philosophy

Centre for Environment and Sustainability


Faculty of Engineering and Physical Sciences
University of Surrey
February 2019
Abstract

The potential for generating renewable electricity from palm oil mill residues (POMR) has
received policy support from the Malaysian Government for almost two decades. However,
uptake of the technology is still relatively low. A significant issue dominating the discussion
for many years is how to translate the renewable electricity generation potential from POMR
into actual implementation. The research seeks to understand the opportunities and barriers for
the use of POMR for a Renewable Electricity System (POMR-RES) in Peninsular Malaysia by
assessing the technical, techno-economic and environmental feasibility of generating
renewable electricity from palm oil mill residues focusing mainly on empty fruit bunches
(EFB) and biogas.
A combination of mathematical analysis and simulation using Aspen PlusTM software was
employed to assess the technical feasibility of the system. Techno-economic analysis was
combined with Life Cycle Assessment (LCA) to integrate the environmental impact
perspective into the POMR-RES evaluation. The results show that EFB has better technical
and techno-economic performance than the biogas. Furthermore, the on-site POMR-RES with
an installed capacity of 5.70 MW or greater is technically feasible, economically viable and
environmentally favourable. The electricity generated from a POMR-RES of this scale is
sufficient to meet a mill’s operational electricity demand, the parasitic load of the POMR-RES
and provide surplus electricity to the national grid. An economically feasible size POMR-RES
are expected to provide:
1. a 20% return on investment (ROI) with five to seven-year payback period (PP).
2. a positive net present value (NPV) with break-even point (BEP) of five to seven-year.
The electricity generated in POMR-RES emits 95% less GHG emissions compared with
current Malaysian electricity grid average when the emission from LUC is excluded from the
electricity generation process. Thirty-five mills in Peninsular Malaysia were identified as
having sufficient EFB supply to operate at or above this economically feasible size with the
total accumulated generation capacity of 200 MW. This accumulated capacity would account
for 25% of the 2020 target for palm oil biomass under National Renewable Energy Policy and
Action Plan.
This research provides a positive case for generating local, renewable electricity from EFB that
can be used as evidence and practical recommendations for various actors such as prospective
investors, analysts, stakeholders, mill owners and policymakers and government agencies such

1
as the Sustainable Energy Development Authority Malaysia (SEDA) for efficient and
sustainable utilization of POMR. This study also makes a positive contribution towards
achieving the national renewable energy target for additional renewable power supplies and as
a contribution towards improved global sustainability.

2
Declaration of Originality

This thesis and the work to which it refers are the results of my own efforts. Any ideas, data,
images or text resulting from the work of others (whether published or unpublished) are fully
identified as such within the work and attributed to their originator in the text, bibliography or
in footnotes. This thesis has not been submitted in whole or in part for any other academic
degree or professional qualification. I agree that the University has the right to submit my work
to the plagiarism detection service TurnitinUK for originality checks. Whether or not drafts
have been so-assessed, the University reserves the right to require an electronic version of the
final document (as submitted) for assessment as above.

Ida Fahani MD JAYE

5th February 2019

3
Table of Contents

Abstract............................................................................................................................................. 1
Declaration of Originality .................................................................................................................. 3
List of Tables .................................................................................................................................... 8
List of Figures ................................................................................................................................. 13
List of Abbreviations ....................................................................................................................... 15
Acknowledgements ......................................................................................................................... 17
CHAPTER ONE ............................................................................................................................. 18
1.1 Introduction ......................................................................................................................... 18
1.2 Electricity Generation Potential from Palm Oil Mill Residues .............................................. 20
1.3 National Aspirations and Targets ......................................................................................... 22
1.4 Renewable Electricity from POMR – The Reality ................................................................ 23
1.5 Research Aims and Objectives ............................................................................................. 24
1.6 Overview of Thesis .............................................................................................................. 25
1.7 Research Contributions ........................................................................................................ 26
CHAPTER TWO ............................................................................................................................ 27
2.1 POMR as a source for RE .......................................................................................................... 27
2.2 Generating RE from POMR in Malaysia .................................................................................... 31
2.2.1 Residues Supply.................................................................................................................. 33
2.2.2 Energy Conversion Technology .......................................................................................... 34
2.2.3 RE Transmission and Distribution ....................................................................................... 36
2.3 Barriers and Opportunities for RE Generation from POMR in Peninsular Malaysia .................... 39
2.4 Summary ................................................................................................................................... 43
CHAPTER THREE ......................................................................................................................... 44
3.1 Scope of Research ..................................................................................................................... 44
3.1.1 Study Area .......................................................................................................................... 45
3.1.2 Types of Residues ............................................................................................................... 47
3.2 Research Framework Summary ................................................................................................. 50
3.3 Data Collection Approach .......................................................................................................... 54
CHAPTER FOUR ........................................................................................................................... 57
4.1 Introduction ............................................................................................................................... 57
4.2 POMR-RES Modelling .............................................................................................................. 59
4.2.1 Mathematical Modelling ..................................................................................................... 59
4.2.2 Computer-Aided Simulation Modelling ............................................................................... 62
4.2.3 Malaysia Palm Oil Mill Standard Energy Calculation Method ............................................. 63
4.3 Data Observation ....................................................................................................................... 64

4
4.4 POMR-RES Power Plant ........................................................................................................... 68
4.4.1 EFB-based POMR-RES Power Plant................................................................................... 68
4.4.2 Biogas-based POMR-RES Power Plant ............................................................................... 70
4.5 MPOB Standard Calculation Method ......................................................................................... 72
4.6 POMR-RES CHP Plant.............................................................................................................. 74
4.6.1 EFB-based POMR-RES CHP Plant ..................................................................................... 74
4.6.2 Biogas-based POMR-RES CHP Plant ................................................................................. 76
4.7 Biomass Dryer ........................................................................................................................... 76
4.7.1 Biomass Dryer Technical Modelling ................................................................................... 77
4.7.2 Dryers heat requirement ...................................................................................................... 79
4.7.3 Dried EFB-based POMR-RES Power Plant ......................................................................... 80
4.8 Conclusion ................................................................................................................................ 83
CHAPTER FIVE ............................................................................................................................. 86
5.1 Introduction ............................................................................................................................... 86
5.2 Feasibility Study for POMR-RES .............................................................................................. 88
5.3 POMR-RES Feasibility Assessment Framework ........................................................................ 92
5.3.1 Technical Feasibility Assessment ........................................................................................ 93
5.3.2 Techno-Economic Feasibility Assessment ........................................................................... 93
5.3.2.1 Cost Estimation ............................................................................................................ 94
5.3.2.2 Revenue Interpretation ................................................................................................. 99
5.3.2.3 Cash Flow and Profitability Evaluations ..................................................................... 100
5.3.2.4 Levelized Cost of Electricity ...................................................................................... 100
5.4 Sensitivity Analysis ................................................................................................................. 101
5.5 Outcomes Illustration ............................................................................................................... 101
5.5.1 OnS case study: On-site POMR-RES installation .............................................................. 102
5.5.1.1 EFB-based POMR-RES ............................................................................................. 102
5.5.1.2 Biogas-based POMR-RES .......................................................................................... 112
5.5.1.3 Sensitivity Analysis .................................................................................................... 116
5.5.2 OffS case study: Off-site POMR-RES installation ............................................................. 120
5.5.2.1 Sensitivity Analysis .................................................................................................... 122
5.5.3 Co-op case study: Co-operative POMR-RES Installation ................................................... 123
5.5.3.1 On-site Co-operative POMR-RES Installation ............................................................ 123
5.5.3.2 Off-site Co-operative POMR-RES Installation ........................................................... 128
5.5.3.3 Sensitivity Analysis .................................................................................................... 128
5.6 Technical and Techno-Economic Assessment of Dried EFB-based Power Plant ....................... 130
5.6.1 Cost for Drying the EFB ................................................................................................... 131
5.7 Summary of Findings............................................................................................................... 132

5
5.8 Economically Feasible Size of POMR-RES ............................................................................. 134
5.9 Conclusions ............................................................................................................................. 137
CHAPTER SIX ............................................................................................................................. 138
6.1 Introduction ............................................................................................................................. 138
6.2 Life Cycle Assessment for Biomass-Based Electricity Generation ............................................ 140
6.3 Life Cycle Assessment for POMR-RES ................................................................................... 142
6.3.1 Goal and Scope ................................................................................................................. 142
6.3.2 LCIA Results and LCA Interpretations and Discussion...................................................... 149
6.3.2.1 Land Use Change ....................................................................................................... 150
6.3.2.2 Oil Palm Plantation .................................................................................................... 152
6.3.2.3 EFB Production Life Cycle Phase (Cumulative, after Allocation) ............................... 154
6.3.2.4 EFB Combustion Life Cycle Phase (Cumulative, after allocation)............................... 155
6.3.2.5 EFB-based Power Plant .............................................................................................. 157
6.3.2.6 Overall Impacts of Electricity Generated in POMR-RES ............................................ 157
6.3.3 LCA Interpretation ............................................................................................................ 159
6.3.3.1 Hotspot Analysis of Electricity Generation via POMR-RES ....................................... 159
6.3.3.2 Limitations of the LCA .............................................................................................. 161
6.3.4 Comparison of Electricity Generation via POMR-RES with Peninsular Malaysia Grid
Average Generation ................................................................................................................... 161
6.3.5 Carbon Emission Saving from EFB-based Power Plant ..................................................... 164
6.4 Summary ................................................................................................................................. 165
CHAPTER SEVEN ....................................................................................................................... 167
7.1 Conclusion .............................................................................................................................. 167
7.2 Suggestions for Further Work .................................................................................................. 170
Reference List ............................................................................................................................... 172
Appendix A ................................................................................................................................... 196
Appendix B ................................................................................................................................... 199
Appendix B-1 – Data Checklist Form ........................................................................................ 199
Appendix B-2 – Statistic information of available EFB and biogas in Peninsular Malaysia ......... 200
Appendix B-3 – Mathematical Modelling Spreadsheet ............................................................... 201
Appendix B-3-1 – Semi-dry EFB-based POMR-RES Power Plant ......................................... 202
Appendix B-3-2 – Biogas-based POMR-RES Power Plant ..................................................... 207
Appendix B-4 – Dryer Enthalpy Balance and Kinematics Modelling Spreadsheet ...................... 214
Appendix B-5 – Mathematical Modelling Spreadsheet for Dried EFB Power Plant .................... 218
Appendix C ................................................................................................................................... 219
Appendix C-1 – Technical Assessment Spreadsheet ................................................................... 219
Appendix C-2 – Techno-Economic Spreadsheet......................................................................... 221
Appendix D ................................................................................................................................... 282

6
Appendix D-1 – Oil Palm Plantation Stage ................................................................................ 282
Appendix D-2 – PO Milling Process Stage................................................................................. 283
Appendix D-3 – Combustion Process Stage ............................................................................... 285
Appendix D-4 – EFB- based Power Plant .................................................................................. 286
Appendix E – Publication .............................................................................................................. 288

7
List of Tables

Table 1.1: RE Potential in Malaysia………………………………………………………......20


Table 1.2: RE Capacity and Carbon Avoidance Target.............................................................22
Table 2.1: Standard Biomass Extraction Rate………………………………………………...28
Table 3.1: Oil Palm Plantation Area (in ha) in Peninsular Malaysia…………………………..45
Table 3.2: Palm Oil Mills According to State in Peninsular Malaysia 2015…………………..46
Table 3.3: Palm Oil Mills According to State in Peninsular Malaysia……………………….47
Table 3.4: POMR Availability in 2015……………………………………………………….48
Table 4.1: Ultimate and Proximate Analysis of EFB………………………………………….63
Table 4.2: Ultimate and Proximate Analysis of Biogas……………………………………….63
Table 4.3: Estimation of Electricity Generation from EFB…………………………………..64
Table 4.4: Estimation of Electricity Generation from Biogas………………………………..64
Table 4.5: Production Profile of Visited Mills……………………………………………….65
Table 4.6: Estimation of EFB Availability………………………………...…………………66
Table 4.7: Estimation of Biogas Availability………………………………………………....66
Table 4.8: Mathematical Modelling of EFB-based Power Plant……………………………...69
Table 4.9: Mathematical Modelling of biogas-based Power Plant………………………...….71
Table 4.10: MPOB Modelling Method of EFB-based Power Plant…………………………..73
Table 4.11: MPOB Modelling Method of biogas-based Power Plant…………………….…..73
Table 4.12: Computer-Aided Modelling of EFB-based CHP Plant……………..…………….75
Table 4.13: Computer-Aided Modelling of biogas-based CHP Plant…………………….…..75
Table 4.14: Dryers Heat Requirement………………………………………….……………..79
Table 4.15: Dried EFB Power Plant Mathematical Modelling with Steam Extraction for Drying
Process……………………………………………………………………………….……….82
Table 4.16: Dried EFB Power Plant Mathematical Modelling without Steam Extraction for
Drying Process……………………………………………………………….……………….82
Table 5.1: Feasibility Assessment of POMR-RES in Malaysia……………………………....89
Table 5.2: POMR-RES Equipment List……………………………………….……………...94
Table 5.3: Lang Factor for Capital Cost Estimation…………………………..………………96
Table 5.4: Feedstock Transportation Linear Equations………………………….……………98
Table 5.5: FiT Rate Entitlement for POMR-RES…………………………………..…………99
Table 5.6: Technical Performance of EFB-based Power Plant……………..………………102
Table 5.7: Cost Estimation of On-Site EFB-based Power Plant………………………..……103

8
Table 5.8: Techno-Economic Performance of EFB-based Power Plant…………….……….105
Table 5.9: Technical and Techno-Economic Performance of On-Site EFB-based CHP Plant
………………………………………………………………………………………………108
Table 5.10: Feasibility Assessment of On-Site Biogas-based Power Plant………..………...111
Table 5.11: Feasibility Assessment of On-Site Biogas-based CHP Plant…………..………..114
Table 5.12: Summary of POMR-RES Performance………………..………………………..115
Table 5.13: Technical and Techno-Economic performance of On-Site Co-Operative
Installation (Co-op 1)………………………………………………….…………………….124
Table 5.14: Technical and Techno-Economic Performance Comparison between Individual
Installations with Co-op 1 Installation……………………………………………………....125
Table 5.15: Costing Comparison of the Individual Installations with Co-op Installation……126
Table 5.16: Technical and Techno-Economic Performance of On-Site Co-Operative
Installation (Co-op 2)………………………………………………………………………..127
Table 5.17: Technical and Techno-Economic Performance Comparison between Individual
Installations with Co-op 2 Installation……...……………………………………………….131
Table 5.18: Estimated Cost for Drying the EFB……………………………………………..136
Table 5.19: POMR-RES Costing Profile……………………………………………………136
Table 6.1: Life Cycle Inventory for Oil Palm Plantation and CPO Milling Process- Data
Comparison…………………………………………………………………………………146
Table 6.2: Economic Allocation for Products and Co-products of CPO Milling Process….149
Table 6.3: Environmental Impacts from the Production of 9.6 kg of FFB at Plantation……152
Table 6.4: Environmental Impacts from EFB Production (Cumulative, Allocated)………....155
Table 6.5: Environmental Impacts from EFB Combustion(Cumulative, Allocated)………..156
Table 6.6: Environmental Impacts from the assigned proportion of the infrastructure 5.7 MW
POMR-RES…………………………………………………………………………………157

Table 6.7: Environmental Impacts for the FU of 1 kWh electricity from EFB after inclusion
of the power plant (Cumulative, Allocated)………………………………………………...158

Table 6.8: Environmental Impacts Comparison of 1 kWh Electricity……...………………..162


Table 6.9: Carbon Emission Saving from Economically Feasible Size POMR-RES………..164
Table A-1: RE Policies and Initiatives in Malaysia………………………………………….190
Table B-1: Availability of EFB and biogas in the mill according to their FFB processing
capacity……………………………………………………………………………………...194
Table B-2: Description of Data Used for Modelling the POMR-RES Power Plant………….195

9
Table B-3: Modelling of semi-dry EFB-based Power Plant in 10 t/hr mill…………………196
Table B-4: Modelling of semi-dry EFB-based Power Plant in 20 t/hr mill…………………196
Table B-5: Modelling of semi-dry EFB-based Power Plant in 30 t/hr mill…………………197
Table B-6: Modelling of semi-dry EFB-based Power Plant in 40 t/hr mill…………………197
Table B-7: Modelling of semi-dry EFB-based Power Plant in 50 t/hr mill…………………198
Table B-8: Modelling of semi-dry EFB-based Power Plant in 60 t/hr mill…………………198
Table B-9: Modelling of semi-dry EFB-based Power Plant in 70 t/hr mill…………………198
Table B-10: Modelling of semi-dry EFB-based Power Plant in 80 t/hr mill………….…….199
Table B-11: Modelling of semi-dry EFB-based Power Plant in 90 t/hr mill…………………199
Table B-12: Modelling of semi-dry EFB-based Power Plant in 100 t/hr mill………………..200
Table B-13: Modelling of semi-dry EFB-based Power Plant in 110 t/hr mill………………..200
Table B-14: Modelling of semi-dry EFB-based Power Plant in 120 t/hr mill………………201
Table B-15: Modelling of biogas-based Power Plant in 10 t/hr mill………………..………201
Table B-16: Modelling of biogas-based Power Plant in 20 t/hr mill……………….……….202
Table B-17: Modelling of biogas-based Power Plant in 30 t/hr mill…………….………….202
Table B-18: Modelling of biogas-based Power Plant in 40 t/hr mill…………………………203
Table B-19: Modelling of biogas-based Power Plant in 50 t/hr mill…………………………203
Table B-20: Modelling of biogas-based Power Plant in 60 t/hr mill…………………………204
Table B-21: Modelling of biogas-based Power Plant in 70 t/hr mill…………………………204
Table B-22: Modelling of biogas-based Power Plant in 80 t/hr mill…………………………205
Table B-23: Modelling of biogas-based Power Plant in 90 t/hr mill…………………………205
Table B-24: Modelling of biogas-based Power Plant in 100 t/hr mill………………………206
Table B-25: Modelling of biogas-based Power Plant in 110 t/hr mill………………………206
Table B-26: Modelling of biogas-based Power Plant in 120 t/hr mill………………………207
Table B-27: Description of Data Used for Modelling the EFB Dryer………………………207
Table B-28: Heat Requirement to Dry EFB in 10 t/hr mill…………………………………..208
Table B-29: Heat Requirement to Dry EFB in 20 t/hr mill…………………………………208
Table B-30: Heat Requirement to Dry EFB in 30 t/hr mill…………………………………209
Table B-31: Heat Requirement to Dry EFB in 40 t/hr mill…………………………………209
Table B-32: Heat Requirement to Dry EFB in 50 t/hr mill…………………………………210
Table B-33: Heat Requirement to Dry EFB in 60 t/hr mill…………………………………210
Table B-34: Heat Requirement to Dry EFB in 70 t/hr mill…………………………………211
Table B-35: Heat Requirement to Dry EFB in 80 t/hr mill…………………………………211
Table B-36: Heat Requirement to Dry EFB in 90 t/hr mill…………………………………211

10
Table B-37: Heat Requirement to Dry EFB in 100 t/hr mill…………………………………211
Table B-38: Heat Requirement to Dry EFB in 110 t/hr mill…………………………….….212
Table B-39: Heat Requirement to Dry EFB in 120 t/hr mill…………………………………212
Table C-1: Technical Feasibility Assessment of 0.9 MW Power Plant……………………...213
Table C-2: Technical Feasibility Assessment of 1.8 MW Power Plant……………………...213
Table C-3: Technical Feasibility Assessment of 2.7 MW Power Plant……………………...213
Table C-4: Technical Feasibility Assessment of 3.6 MW Power Plant……………………...213
Table C-5: Technical Feasibility Assessment of 4.4 MW Power Plant……………………...213
Table C-6: Technical Feasibility Assessment of 5.3 MW Power Plant…………………...…214
Table C-7: Technical Feasibility Assessment of 6.2 MW Power Plant………………...……214
Table C-8: Technical Feasibility Assessment of 7.1 MW Power Plant……………...………214
Table C-9: Technical Feasibility Assessment of 8.0 MW Power Plant……………………...214
Table C-10: Technical Feasibility Assessment of 8.9 MW Power Plant…………………….214
Table C-11: Technical Feasibility Assessment of 9.8 MW Power Plant…………………….215
Table C-12: Technical Feasibility Assessment of 10.7 MW Power Plant…………………...216
Table C-13: Techno-Economic Assessment of the Private-Funded 0.9 MW Power Plant…..218
Table C-14: Techno-Economic Assessment of the Private-Funded 1.8 MW Power Plant…..220
Table C-15: Techno-Economic Assessment of the Private-Funded 2.7 MW Power Plant…..222
Table C-16: Techno-Economic Assessment of the Private-Funded 3.6 MW Power Plant…..224
Table C-17: Techno-Economic Assessment of the Private-Funded 4.4 MW Power Plant…..226
Table C-18: Techno-Economic Assessment of the Private-Funded 5.3 MW Power Plant…..228
Table C-19: Techno-Economic Assessment of the Private-Funded 6.2 MW Power Plant…..230
Table C-20: Techno-Economic Assessment of the Private-Funded 7.1 MW Power Plant…..232
Table C-21: Techno-Economic Assessment of the Private-Funded 8.0 MW Power Plant…..234
Table C-22: Techno-Economic Assessment of the Private-Funded 8.9 MW Power Plant…..236
Table C-23: Techno-Economic Assessment of the Private-Funded 9.8 MW Power Plant…..238
Table C-24: Techno-Economic Assessment of the Private-Funded 10.7 MW Power Plant…240
Table C-25: Techno-Economic Assessment of the Self-Funded 0.9 MW Power Plant……...242
Table C-26: Techno-Economic Assessment of the Self-Funded 1.8 MW Power Plant……...244
Table C-27: Techno-Economic Assessment of the Self-Funded 2.7 MW Power Plant……...246
Table C-28: Techno-Economic Assessment of the Self-Funded 3.6 MW Power Plant……...248
Table C-29: Techno-Economic Assessment of the Self-Funded 4.4 MW Power Plant……...250
Table C-30: Techno-Economic Assessment of the Self-Funded 5.3 MW Power Plant……...252
Table C-31: Techno-Economic Assessment of the Self-Funded 6.2 MW Power Plant……...252

11
Table C-32: Techno-Economic Assessment of the Self-Funded 7.1 MW Power Plant……...254
Table C-33: Techno-Economic Assessment of the Self-Funded 8.0 MW Power Plant……...256
Table C-34: Techno-Economic Assessment of the Self-Funded 8.9 MW Power Plant……...258
Table C-35: Techno-Economic Assessment of the Self-Funded 9.8 MW Power Plant……...260
Table C-36: Techno-Economic Assessment of the Self-Funded 10.7 MW Power Plant…….262
Table C-37: DCF of the Private-Funded 0.90 MW Power Plant…………………………...270
Table C-38: DCF of the Private-Funded 1.80 MW Power Plant…………………………...271
Table C-39: DCF of the Private-Funded 2.70 MW Power Plant…………………………...272
Table C-40: DCF of the Private-Funded 3.60 MW Power Plant…………………………...273
Table C-41: DCF of the Private-Funded 4.50 MW Power Plant…………………………...274
Table C-42: DCF of the Private-Funded 5.40 MW Power Plant…………………………...275
Table C-43: DCF of the Private-Funded 6.30 MW Power Plant…………………………...276
Table C-44: DCF of the Private-Funded 7.20 MW Power Plant…………………………...277
Table C-45: DCF of the Private-Funded 8.00 MW Power Plant…………………………...278
Table C-46: DCF of the Private-Funded 8.90 MW Power Plant…………………………...279
Table C-47: DCF of the Private-Funded 9.80 MW Power Plant…………………………...280
Table C-48: DCF of the Private-Funded 10.70 MW Power Plant…………………..……...281
Table D1: Life Cycle Inventories for Oil Palm Plantation to Produce 1 kg of FFB………….283
Table D2: Life Cycle Inventories for Wet Milling Process to Produce 1 kg CPO………….285
Table D3: Life Cycle Inventories for Combustion 1 kg EFB………………...………………286
Table D4: Life Cycle Inventories for 5.7 MW POMR-RES……………………...………….287

12
List of Figures

Figure 1.1: Electricity Generation Fuel Mix in Peninsular Malaysia………………………….19


Figure 1.2: The POMR-based Electricity Generation………………………………………...23
Figure 2.1: Bioenergy Conversion Routes……………………………………………………31
Figure 2.2: POMR RE Generation Value Chain………………………………………………33
Figure 2.3: Relationship between the Theoretical Availability of Residues and Constraints on
their Availability…………………………………………………………………………...…38
Figure 2.4: Peninsular Malaysia Electricity Supply System……...…………………………...47
Figure 3.1: Wet Milling Process of CPO……………………………………………………...47
Figure 3.2: Research Framework……………………………………………………………..54
Figure 4.1: POMR-RES Power Plant………………………………………………………....58
Figure 4.2: POMR-RES CHP Plant…………………………………………………………..58
Figure 4.3: POMs EFB and Biogas Production Profile……………………………...……….67
Figure 4.4: Variation between Mathematical and MPOB Modelling of POMR-RES Power
Plant…………………………………………………………………………………………..72
Figure 4.5: Actual EFB Availability…………………………………………………………84
Figure 5.1: POMR-RES Feasibility Assessment Framework…………..…………………….92
Figure 5.2: Impact of Interest Rate Variation on ROI of on-site POMR-RES Intallation…....117
Figure 5.3: Impact of Exclusion of New Grid Installation Cost on ROI of on-site POMR-RES
Intallation…………………………...………………………………………………………118
Figure 5.4: Impact of Cost Variation on ROI of on-site POMR-RES installation…………..118
Figure 5.5: ROI Value Comparison between Connecting All Electricity and Surplus Electricity
to Grid……………………………………………………….………………………………119
Figure 5.6: Impact of Interest Rate Variation on NPV of On-Site POMR-RES
installation…………………………………………………………………………………..120
Figure 5.7: Impact of Interest Rate Variation on ROI of Off-Site POMR-RES Installation...122
Figure 5.8: POMR-RES Energy System…………………………………………………….133
Figure 6.1 (a): GHG Emission from Renewable Electricity…………………………...…….139
Figure 6.1 (b): GHG Emission from Fossil-Based Electricity…………………...…………..140
Figure 6.2: EFB-Based Power Plant………………………………………………….……..144
Figure 6.3: Peninsular Malaysia Oil Palm Plantation and CPO Production Trend…………151
Figure 6.4: Environmental Impacts from Production of 9.6 kg FFB at Plantation…….……152
Figure 6.5: Environmental Impacts Contributors for FFB Plantation……………………….154

13
Figure 6.6: Hotspot Analysis of Electricity Generation Process…………………………….159
Figure 6.7: Environmental Impacts Comparison of 1 kWh Electricity from POMR-RES with
the National Grid……………………………………………………………………………162

14
List of Abbreviations

AD Anaerobic Digester
BEP Break-even Point
CEPCI Chemical Engineering Plant Cost Index
CHP Combined Heat and Power
CH4 Methane
CO2 Carbon Dioxide
CO2eq Carbon Dioxide Equivalent
Co-op Cooperative Installation
CPO Crude Palm Oil
dLUC Direct Land Use Change
EC Energy Commission
EFB Empty Fruit Bunches
EIs Environmental Impacts
EPU Economic Planning Unit
ETP Economic Transformation Program
FiT Feed-in Tariff
FEP Freshwater Eutrophication Potential
FFB Fresh Fruit Bunches
FFS Fossil Fuel Scarcity
FU Functional Unit
GHG Greenhouse Gasses
GW Giga Watt
GWP Global Warming Potential
HTPsc Human Carcinogenic Toxicity Potentials
HTPsnc Human Non-Carcinogenic Toxicity Potentials
iLUC Indirect Land Use Change
IRR Internal Rate of Return
IPPs Independent Power Producers
LCA Life Cycle Assessment
LCI Life Cycle Inventories

15
LCIA Life Cycle Impact Assessment
LUC Land Use Change
MF Mesocarp Fibres
MSW Municipal Solid Waste
MW Mega Watt
MWh Mega Watt-hour
MY Malaysia
NKEA National Key Economic Area
NPV Net Present Value
NREAP National Renewable Energy Action Plan
N2O Nitrous Oxide
PKS Palm Kernel Shell
PKO Palm Kernel Oil
POME Palm Oil Mill Effluent
POMR Palm Oil Mill Residues
POMs Palm Oil Mills
POMR-RES Palm Oil Mill Residues for a Renewable Electricity System
PV Photovoltaic
OffS Off-site Installation
OnS On-site Installation
OPF Oil Palm Fronds
OPT Oil Palm Trunks
RE Renewable Energy
REPPA Renewable Energy Power Purchasing Agreements
ROI Return on Investment
RSPO Roundtable on Sustainable Palm Oil
SEDA Sustainable Energy Development Authority
SREP Small Renewable Energy Program
SESB Sabah Electricity Sdn Bhd
SESCO Sarawak Electricity Supply Cooperation
TAP Terrestrial Acidification Potential
TNB Tenaga Nasional Berhad

16
Acknowledgements

All praises to God who gave me the strength and opportunities to complete this thesis after all
the challenges and hard works. Then to my sponsor, Majlis Amanah Rakyat (MARA), for their
trust and financial support to make this PhD become a reality. I could never have reached this
point without the help, support, guidance, and efforts of a lot of people. My heartfelt and sincere
thanks to my two supervisors, Prof. Richard Murphy, and Dr. Jhuma Sadhukhan for their spot
on and challenging comments and continuous support, which have always been enlightening.
Their professionalism and expertise have been and will continue encouraging me to become a
better researcher in the future. Richard, thank you for spending your time to attend a few
meetings and discussions in Kuala Lumpur, Malaysia. Jhuma, your patience while guiding me
to understand all the chemical engineering principal and the time spent to develop the model
means a lot to me.

My PhD will not be as smooth as without generous and genuine help from Moira Foster, the
centre administrator. Thank you Moira, for your countless help, advice and nags. You have
been the most reliable support system that I can wish for since the beginning of my PhD
journey. I owe you a lot. I would like to convey my sincere appreciation to my sparring partner
for the past two years, Punch. From a stranger, we become a sister. Thank you for all your
spiritual and emotional support and the most important was for hitting my ‘head and face’
during our time in Osaka. This appreciation also goes to all my other colleagues, Tyler,
Richard, Raya, Khadeejah, Yi, Nini, Mercio, Marcio and Melissa (I’m sorry if I missed anyone
of you). I would like to take this opportunity to thank my friend Kok Siew for his constant
reply to my emails and messages. Your simple way of explaining stuff makes Chemical
Engineering doable for me. Thank you to everyone in CES for all the help and sharing all sort
of tricks to survive PhD. Not to forget my comrade Sarah and Ireena for their constant supports,
motivations and prayers.

A zillion thanks to my family, especially to my mom and dad. I want you to know that this
effort is dedicated to you. Thank you for your countless prayers, comfort words and
encouragements. To my siblings, thank you for your support and for being there with mom and
dad when I’m away. To my five little and not so little monkeys, hopefully, this journey will be

17
an inspiration for you to keep on pursuing your dreams. Lastly, thank you to everyone who has,
directly and indirectly, supported me since the beginning.

CHAPTER ONE

INTRODUCTION

1.1 Introduction

Providing adequate, affordable and clean electricity has become the global agenda. Various
initiatives have been taken globally to ensure the electricity generated is adequate to drive
economic growth, preserve the environment and maintain price competitiveness. Amongst all
the initiatives for electricity generation, reducing reliance on fossil fuels is central to achieving
this agenda. The International Energy Agency’s energy statistics for 2015 reported that an
average of 66% of the world’s electricity generation was based on fossil fuels. This is followed
by hydro (16%), nuclear (11%) and non-hydro renewable resources (7%) (International Energy
Agency, 2017). It was further reported that the high percentage of fossil fuel utilization in the
electricity generation sector has contributed roughly 25% of global greenhouse gas (GHG)
emissions (World Energy Council, 2016b).

In regard to this, decarbonizing the electricity generation sector using alternatives to fossil fuel
will play a vital role in the dynamics of achieving the global agenda. Numerous initiatives have
been undertaken by various parties to reduce the reliance on fossil fuel to generate electricity,
and the share of non-hydro renewable resources is expected to grow significantly (REN 21,
2017). An expanding body of literature has acknowledged renewable resources; which include
biomass, biogas, solar photovoltaic, wind and geothermal as suitable alternative resources to
decarbonize the electricity generation sector (Shuit et al., 2009; Rodriguez et al., 2011;
Borhanazad et al., 2013; Turconi, Boldrin and Astrup, 2013). Amongst these, biomass has been
reported to have the highest potential in generating low-carbon electricity owing to its ‘carbon-
neutral’ properties (Hansen, Olsen and Ujang, 2012). This potential can further benefit
countries with abundant biomass resources like Malaysia in its generation of low-carbon
electricity (Gallagher, 2001; Sulaiman et al., 2011; Bakhtyar et al., 2013).

In Malaysia, fossil-based electricity generation is the largest single source to the national
carbon emission (Mohd Safaai et al., 2010; Nakata, Silva and Rodionov, 2011).The Peninsular

18
Malaysia Electricity Supply Outlook in 2017 reported a total installed electricity generation
capacity of 20,900 MW with significant dominance by coal and natural gas (Figure 1.1). As a
result, the electricity generation sector emitted about 60 million tonnes of GHG in 2005 which
will further see an increase to 153 million tonnes by 2020 (Shamsuddin, 2012).

Coal
Hydro
53%
5.50%

Renewables
7% Biomass
0.40%

Biogas
Natural Gas 0.10%
40%
Solar
1.00%

Natural Gas Coal Hydro Biomass Biogas Solar

Figure 1.1 Electricity Generation Fuel Mix in Peninsular Malaysia (Energy Commission,
2017)

Driven by population growth and robust economic progress, the Malaysian electricity peak
demand is expected to increase to 23,100 MW by 2030 (Chua and Oh, 2010) which puts extra
pressure on the electricity generation sector to deliver a secure and reliable electricity supply.
This scenario is of serious concern as Malaysia has experienced rapid depletion of natural gas
reserves (Khor and Lalchand, 2014) and limited local coal resources which have led to
significant imports from Indonesia and Australia (Ong, Mahlia and Masjuki, 2011). In addition,
it is further reported that no additional source of hydropower will be available after 2025 as the
current resources become the last hydro resources in Peninsular Malaysia (Energy
Commission, 2017). The above-mentioned scenarios indicate an urgent need for additional
electricity generation resources in meeting the future electricity demand. Additional resources
reduce the likelihood of Peninsular Malaysia experiencing peak power deficits in the
foreseeable future and at the same time could mitigate carbon emission from the sector
(KeTTHA, 2008, 2017). That said, the non-hydro renewable resources are required to step up
and uplift in order to balance the effects of power deficits and further contribute towards
decarbonizing the electricity generation sector. The following sub-sections give the overview

19
of the electricity generation potential from various renewable energy resources in Malaysia
focusing mainly on palm oil mill residues (POMR) as the core interest of this thesis, along with
the national renewable energy aspirations and targets. It also discusses the reality of generating
renewable electricity from POMR in Peninsular Malaysia.

1.2 Electricity Generation Potential from Palm Oil Mill Residues

As a tropical country rich in natural resources, Malaysia is blessed with a variety of alternative
energy resources as a substitute for fossil fuels for electricity generation. Various academic
publications and statistical reports have appraised the annual electricity generation potential
from these resources as indicated in Table 1.1. From the analysis, biomass and solar
photovoltaic (PV) are identified as having the highest electricity generation potential. Biomass
holds 80% of the total potential followed by solar PV with 18% with the gross annual
cumulative economic value cited as US$ 3,951 million and US$ 77 million, respectively
(Sulaiman et al., 2010, 2011; Abdullah and Sulaiman, 2013).

Table 1.1: Annual RE Potential in Malaysia for 2013 (Ahmad and Tahar, 2014)

Resources Potential (GW)


Biomass 29.0
Solar PV 6.5
Mini Hydro 0.5
Biogas 0.3
Total 36.3

The electricity generation potential from biomass includes the potential from forestry and
agricultural residues, and municipal solid waste. Regarding the agricultural residues, those
from the palm oil industries have the highest electricity generation potential (Umar, Jennings
and Urmee, 2014a). The Malaysian Innovation Agency reported that the residues from milling
the palm oil fresh fruit bunches (FFB) accounts for 86% of the total mass of agricultural
residues in the country (Malaysia Innovation Agency, 2013; KeTTHA, 2017). This is evident
in the reported 94 million tonnes of combined POMR, consisting of empty fruit bunches (EFB),
mesocarp fibres (MF), palm kernel shell (PKS) and palm oil mill effluent (POME) (Energy
Commission, 2016; Sadhukhan et al., 2018). In line with growing demand for PO, the Malaysia
Innovation Agency under the National Biomass Strategy 2020 has anticipated that about 100
million tonnes of dry residues and 100 million tonnes of POME will be generated by the year

20
2020. These residues are available mainly for bioenergy and for several other value-added
products (Malaysia Innovation Agency, 2013).

The feasibility of electricity generation from POMR is due to the perennial availability of
POMR, their ‘carbon neutrality’ with regard to their GHG balance and their high energy
content ranging from 8.5 MJ/kg to 20 MJ/kg. Shamsuddin (2012) has quoted the maximum
potential from POMR as 3500 MW, and Shafie et al. (2012) have estimated about 53.7 TWh
of electricity could be generated annually from PKS, MF, and EFB. EFB accounts for the
highest electricity generation potential compared with the other POMRs. It is forecast that by
2020 about 70 TWh of electricity (approximately 875 MW installed capacity) can be generated
using POMR (Shafie et al., 2012). Although less reported, the biogas extracted from anaerobic
digestion of 0.65 tonne POME is capable of generating 25 – 40 kWh of electricity (Hasanudin
et al., 2015). Elsewhere, Ali et al. (2015) suggested about 1.8 kWh of electricity is generated
from every m3 of biogas. The annual cumulative potential electricity generation capacity of
biogas in 2015 was reported at 168 MW (Mohd Ghazi and Muhammad Nasir, 2017; Md Jaye,
Sadhukhan and Murphy, 2018).

The environmental advantages of using POMR over fossil fuel is manifest in the level of GHG
emitted per kWh of electricity generated. Andiappan, Ng and Bandyyopadhyay reported that
every 1 kilowatt-hour of electricity generated from fossil fuel in Malaysia releases about 0.77
kg CO2eq – the highest compared to neighbouring countries such as Indonesia, Thailand and
Singapore. Mahlia (2002) confirmed the capability of POMR-based electricity to lowering the
carbon emission against the conventional fossil-based electricity. Sadhukhan et al. (2018)
hypothesised that POMR-based electricity in Malaysia releases 0.65 kg CO2eq/kWh,
representing a reduction of 0.12 kg CO2eq/kWh.

Based on this hypothesis, incorporating electricity generated from POMR into the Malaysian
electricity generation portfolio would be a positive way to gradually transition from a highly
fossil fuel dependant electricity generation system towards a more sustainable one in the future
(Basri, Ramli and Aliyu, 2015). The following section details the national RE aspirations and
targets to describe the vital role that POMR could offer in diversifying the electricity generation
mix.

21
1.3 National Aspirations and Targets

The national RE aspiration began in 1980 under the Fourth-Fuel Diversification Strategy with
the aim to increase energy security and further reduce dependency on fossil fuel. Towards this
reduction, hydro was named as the fourth fuel after oil, coal and natural gas (Hashim and Ho,
2011). The aspiration continued in the 8th Malaysia Plan in 2001-2005 when the Malaysian
Government streamlined the RE portfolio by incorporating the non-hydro resources – biomass,
biogas, solar PV, mini-hydro, and geothermal – as the fifth fuel under the Fifth Fuel Policy
(Maulud and Saidi, 2012). The aim of this streamlining was to diversify the national energy
mixture as a proactive measure in reducing over-reliance on fossil fuel in the electricity
generation sector. This streamlining further signified the national enthusiasm to meet the global
commitments to reduce GHG emissions (Basri, Ramli and Aliyu, 2015; Oh et al., 2017).

In reference to reducing over-reliance on fossil fuel, it is reported that by the end of 2005, RE
was expected to contribute 500 MW to the grid in Peninsular Malaysia and Sabah. The small
renewable energy program (SREP) was introduced to spearhead the development of the grid-
connected RE-based electricity generation. However, the outcome of the program seems
disappointing given that only 12 MW of electricity was added to the grid (Shamsuddin, 2012).
However, the Malaysian Government continued to drive the national aspiration by introducing
various incentives to foster the development of grid-connected RE-based electricity generation
plants. The incentives include the institution of National Renewable Energy Action Plan
(NREAP) and Renewable Energy Act, initiation of RE policies, introduction of feed-in tariff
(FiT) mechanism, and establishment of the Sustainable Energy Development Authority
(SEDA) to manage the implementation of the FiT (Mukherjee and Sovacool, 2014;
Shahmohammadi et al., 2015). Targets for renewable energy-based grid-connected electricity
and annual CO2eq avoidance for Peninsular Malaysia and Sabah were reviewed and recalibrated
in line with the introduction of these incentives. Both targets are presented in Table 1.2 with
particular attention given to the palm oil biomass (see KeTTHA, 2008 pg. 32).

Table 1.2: RE Capacity and Carbon Avoidance Target (Umar, Jennings and Urmee, 2014a)

Year RE Target Annual CO2eq Avoidance Target Palm Oil Biomass Target
Ending (MW) (million tonnes) (MW)

2015 985 12 330


2020 2,080 46 800
2030 4,000 163 1,340

22
1.4 Renewable Electricity from POMR – The Reality

The potential for generating sustainable electricity from POMR is considered substantial and
has received continuous support from the Malaysian Government for almost two decades.
However, adoption of the technology remains low and has not attracted significant investment.
A major issue that has dominated the discussion for many years is how to translate the
renewable electricity generation potential from POMR into actual implementation. The
difficulty with implementation is mainly because POMR-based electricity generation is
sandwiched between the two most competitive and profitable sectors; crude palm oil (CPO)
production and highly subsidized fossil-based electricity generation. The complex structure of
both sectors with the involvement of multiple stakeholders has limited adoption of the
technology and investment interest for the POMR-based electricity generation (Loh and Abdul
Majid, 2016). Figure 1.2 shows the position of the POMR-based electricity generation system
and the stakeholders involved.

POMR-based
MPOB, MPIC Electricity
Generation
Private Estate/GLC
Independent Fossil-based National and
Small Holders Palm Oil Electricity State Utilities
Industries Generation
FELDA/FELCRA Sector IPPs, DPPs
/RISDA
Co-generators
Millers
Energy
Commission
EPU,
KeTTHA,
SEDA

Figure 1.2: The Actors Involved in POMR-based Electricity Generation

In 2016, the Energy Commission (EC) cancelled three POMR-based electricity generation
licences due to multiple technical and economically unfeasible conditions (Energy

23
Commission, 2017). By the end of 2017, only 12 MW RE capacity from POMR was installed
and connected to the grid (Oh et al., 2017). These amounts are very far from the optimistic
target of having 800 MW contributions from POMR by 2020. As stated earlier, this
disappointing performances in generating RE from POMR has heightened the need for a critical
feasibility assessment of RE generation proposals. However, to date, there has been little
consensus on strategies to be adopted to conduct this feasibility assessment (Sovacool and
Drupady, 2011).

1.5 Research Aims and Objectives

As presented in the previous sections, the actual potential of generating electricity from POMR
has not been translated into practise due to limited scientific evidence on the technical and
techno-economic feasibility of the electricity generation process. This circumstance has
resulted in a limited number of successful business implementation cases, which creates
uncertainties for future investors when considering entering the POMR-based electricity
generation business. Therefore, expectations for electricity generations from POMR and
national renewable energy targets were not met. This research explores a number of key issues
relevant to the translation of the POMR-based electricity generation potential into
implementation. These are:

i. Bounded information regarding the actual availability of POMR that can be used for
renewable electricity generation.
ii. Non-conducive environment for the RE generation business due to unfavourable
subsidies and incentives offered for generating electricity from POMR.
iii. Limited rigorous technical, techno-economic feasibility and environmental analysis
available for POMR-based electricity generation.
iv. Absence of sound framework to conduct such integrated feasibility assessments.
v. Lack of comprehensive regulatory framework to foster the sustainable deployment of
POMR-based electricity generation via a clear roadmap to guide prospective investors,
mill owners and other stakeholders in the design, development and feasibility of new
electricity generation developments.

As such, this research aims to provide practical insights to promote translation of the potential
for generating renewable electricity generation from POMR into implementation. This can be
achieved by scientific, transparent assessment of the feasibility of POMR-based electricity

24
generation under a comprehensive technical and techno-economic feasibility and
environmental impact assessment. Four objectives were established for the research:

• Model the palm oil mill residues to renewable electricity system (POMR-RES) to better
evaluate the electricity generation potential from the available EFB and biogas from
anaerobic digestion of POME in the palm oil mills.
• Assess the technical and techno-economic feasibility of the POMR-RES under different
scenarios (i.e. interest rate variations, FiT rate revisions, grid cost exclusions) using
appropriate case studies.
• Provide a clear financial analysis of economically feasible operating conditions for
POMR-RES based on favourable returns on investment (ROI), net present value
(NPV), and payback period (PP) and break-even point (BEP).
• Develop the environmental impact profile of economically feasible POMR-RES plants
or consortia.

1.6 Overview of Thesis

The thesis is organized in seven chapters. Chapter 1 provides a brief introduction to the
electricity generation sector in Malaysia together with the aims and objectives of the research.

Chapter 2 presents a critical review of generating electricity from POMR in a macro-economic


context. This includes a review of the barriers to, and drivers for, implementation of such
systems.

Chapter 3 provides an overview of the research by detailing the research scope and summarizes
the proposed research framework.

Chapter 4 explains the data collection for the research and the procedures to model the POMR-
RES power plant and combine heat and power (CHP) plant. The primary aim of this chapter is
to establish the electricity and heat generation profile from combusting the available EFB and
biogas from anaerobic digestion of POME in the mills.

Chapter 5 comprises two parts. The first assesses the technical and techno-economic feasibility
of POMR-RES. Assessment of the technical feasibility of the system is based on the amount
of electricity generated. The techno-economic feasibility is based on the projected cash flow to
quantify the ROI and discounted cash flow to quantify the NPV. Additionally, the PP and BEP
of POMR-RES is also estimated. The procedures to obtain the economically feasible size of

25
the POMR-RES and the estimation of the amount of residues required to operate at this size
are identified in the second part of the chapter. The number of mills in Peninsular Malaysia
that can operate at the economically feasible size is also identified in the latter part of the
chapter.

Chapter 6 describes the Life Cycle Assessment (LCA) procedure to develop the environmental
profile of the economically feasible size EFB-based power plant.

Chapter 7 presents the overall conclusions from the research and offers suggestions on future
works.

1.7 Research Contributions

The findings from the research offer clear evidence in regarding the feasibility of electricity
generation from POMR in Peninsular Malaysia. Firstly, this study is considered to be the first
attempt to establish a site-specific electricity and heat generation profile for EFB and biogas.
Second, this research provides an explicit technical and techno-economic feasibility
comparison between EFB and biogas to identify the most suitable residues for electricity
generation. A major contribution from this research is the proposal of a framework to indicate
an economically feasible size of POMR-RES for profitable electricity generation. This
establishment of an economically feasible scale for POMR-RES plant provides the necessary
underpinning to promote investment and resource utilization allowing the palm oil industries
to exploit the opportunities in the electricity generation sector.
In the broader context, this research provides evidence and practical recommendations for
various actors such as analysts, policymakers and government agencies such as SEDA to
restructure the existing mechanism of the RE quota distributions and allocations specifically
for biomass and biogas. Such restructuring could promote improved distributions of the RE
quota to eligible recipients, thus reinforcing the effectiveness and sustainability of the RE Fund.
It is further expected to support policymakers and analyst in their efforts developing reliable
policies for sustainable RE in the future. Finally, it expands the current body of knowledge
associated with POMR utilization for electricity generation through its novel research findings.

26
CHAPTER TWO

RENEWABLE ELECTRICITY GENERATION IN MALAYSIA

2.1 POMR as a source for RE

Biomass is one of the oldest renewable resources known to humans. It comprises of resources
such as energy crops, forestry, agricultural and municipal waste, effluents, and manure and
sewage sludge. These resources provide energy sources and help to create renewable products
such as biofuel, bioenergy, biomaterials and biochemicals (Ragauskas et al., 2006; Black et al.,
2011). The energy provided by biomass resource can be extracted through thermo-chemical
process by heating and refining the resources into combustible liquid fuels and through bio-
chemical processes by using enzymes, microbes and catalysts to make fuels as shown in Figure
2.1 (Faaij, 2006; Black et al., 2011; World Energy Council, 2016a).

Conversion Technology Products Market

Thermochemical Biochemical

Combustion Heat Electricity

Gasification Biochar

Pyrolysis Syngas Heat

Liquefaction Synoil Biofuels

Fermentation Bioethanol

Biomaterials
Digestion Biogas
and chemicals

Esterification Biodiesel

Figure 2.1: Potential Bioenergy Conversion Routes (after Black et al., 2011; Ozturk et al.,

2017)

27
Research and practice has identified that the generation of electricity from biomass has
numerous benefits, for instance, reducing dependency on fossil fuels and substantially
mitigating greenhouse gas (GHG) emissions (Panwar et al., 2011; Martinez-Hernandez et al.,
2013; Petinrin & Shaaban, 2015; Codina Gironès et al., 2017). Other authors (e.g. Demirbas,
Balat and Balat, 2009; Loorbach, 2010; Panwar et al., 2011) have highlighted the potential to
improve the security of electricity supply, contribute to poverty reduction, promote practical
waste management and elevate the value of such resources as reasons for using biomass for
electricity generation. The utilization of biomass for electricity generation has become part of
the energy policies agenda in many nations, has received government subsidies and incentives,
and has created knowledge sharing and technology transfer between countries (Reiche &
Bechberger, 2004; Haas et al., 2011; El Fadel et al., 2013; Oh et al., 2017). However, Plevin
et al. (2010) emphasized that inappropriate exploitation of biomass for electricity can/has led
to deforestation issues with substantial GHG emissions resulting from the land conversion.

In Malaysia, biomass is sourced from residues from oil palm plantations such as the oil palm
trunks (OPT) and fronds (OPF) (Singh et al., 2013). Furthermore, the milling process to obtain
crude palm oil (CPO) in palm oil mills (POMs) produces solid palm oil mill residues (POMR),
such as empty fruit bunches (EFB), mesocarp fibres (MF) and palm kernel shell (PKS) and
also liquid residues in the form of palm oil mill effluent (POME) (Yusoff, 2006; Loh, 2017).
It is estimated that the oil palm sectors will generate approx. 100 million tonnes of solid POMR
and 100 million tonnes of POME by 2020 (Malaysia Innovation Agency, 2013) (Table 2.1).

Table 2.1: Standard Palm Oil Residues Extraction Rates (Yusoff, 2006; Loh, 2017)

Types of Residues Extraction Rate


POMR
EFB 22.0% of per tonne FFB processed
MF 13.5% of per tonne FFB processed
PKS 0.5% of per tonne FFB processed
POME 60.0% of per tonne FFB processed
Plantation
OPT 74.5 t / ha oil palm plantation
OPF 10.4 t / ha oil palm plantation

28
Utilization of the plantation residues is still limited due to long-term soil enhancement and
management consideration and, for that reason, only 50% of these residues are permitted to be
removed from the plantation area, as suggested by MPOB (Loh, 2017). On the other hand,
published research has provide evidence of the interesting potential and initiatives to enhance
the usability of POMR, specifically for electricity generation (Muis et al., 2010; Mahlia et al.,
2012; Ng et al., 2013; Hunpinyo et al., 2013). These studies indicate a substantial theoretical
potential to generate electricity from POMR and agree that the utilization of these residues for
electricity generation represents one of the most sustainable ways to increase the electricity
supply security, reduce dependency on fossil fuels and to minimise carbon emission from
electricity generation. Hypothetically, the total RE generation potential from POMR in 2012
for Malaysia was reported as 968 MW (Shafie et al., 2012). Umar, Jennings and Urmee (2014a)
indicated that the amount of available residues in 2020 are sufficient to achieve the 800 MW
target for the RE generation capacity for Malaysia while Sadhukhan et al. (2018) claimed that
34% of the national electricity demand could be supplied by the renewable electricity (RE)
generated from the available POMR in 2015. This aspirational claim, if translated into its actual
deployment, should improve the electricity supply constraints and help achieve Malaysia’s RE
generation target.

In addition to its high potential to increase the security of electricity supply, RE from POMR
is capable of reducing the GHG emission from the electricity generation sector. The Malaysian
electricity generation sector is predicted to emit approximately 800 million tonnes of GHG by
2020 if its fuel source remain as at present, an increase of 63% from emissions in 2000
(Mohamed and Lee, 2006; Sarkar et al., 2013). Sadhukhan et al. (2018) reported about 0.12 kg
CO2eq/kWh is avoided by RE from POMR relative to conventional fossil-based electricity
generation. This concurs well with simulation by Chanlongphitak et al., (2015) who assigned
a value of 0.15 kg CO2eq /kWh as the life cycle emission factor for RE from EFB. Several other
studies have also indicated lower emission factor for RE from POMR indicating that utilization
of this resources for electricity generation is more environmentally friendly (at least in this
regards) than fossil-based alternatives (Chiew and Shimada, 2013; Siregar et al., 2015; Beaudry
et al., 2017).

Unfortunately, despite these potentials, POMR are currently underutilized. MF and PKS are
mostly used as the fuel for the mill’s internal boilers while EFB and POME are used for non-
energy purposes with a minimal amount being returned to the plantation as fertilizer. The
remaining EFB and POME are treated as ‘waste’ with little or no economic value (Ali et al.,

29
2015). EFB are commonly discarded and left to biodegrade naturally since the banning of
incineration in 2000. In the past, POME has been ‘treated’ using open anaerobic ponding before
the effluent is discharge to the watercourses (Madaki and Seng, 2013); however, this is no
longer permitted (see below). Although natural biodegradation of EFB and especially open
anaerobic ponding treatment of POME are claimed as the most cost-effective and least
complicated way to dispose these residues/wastes from the mills, they have had adverse effects
on the environment (Wu et al., 2010). These treatment methods emit methane to the atmosphere
which has approx. twenty five times the GHG value than CO2 (Yacob et al., 2006, IPCC, 2007).
The LCA studies for palm oil production recognized that methane emission from open ponding
treatment contributes up to 52% of total GHG emission of the palm oil industry (Subramaniam
et al., 2010; Ng et al., 2011; Nasution, Wibawa and Noguchi, 2018).

Acknowledging the electricity generation potential and environmental disadvantages of open


ponding POME treatment method, in 2010, the Malaysian Government under the Economic
Transformation Programme has compelled mills to treat the POME in closed anaerobic digester
(AD) tanks and capture the resulting biogas (PEMANDU, 2010). The resulting biogas is
captured and either flared to CO2 or utilized in order to overcome the drawbacks of methane
emission to the atmosphere from the open anaerobic ponding. Consequently, all mills are
expected to have their closed AD processes in place by 2020 (Malaysia Innovation Agency,
2013).

According to Shamsuddin (2012), the electricity generation potential from the biogas extracted
from anaerobic digestion of POME in 2011 is 0.6 GW. Chin et al. (2013) emphasized that
utilizing biogas from POME for electricity generation could increase the renewable energy
share in the national electricity generation mix. Sadhukhan (2014) claimed that biogas from
sewage sludge could be used to generate heat and power at various scales and estimated that
0.92 m3 of natural gas can be saved from the use of 1 m3 of biogas. In similar vein, a laboratory
study by Hasanudin et al. (2015) concluded that the biogas extracted from 0.65 tonnes POME
are capable of generating 25.4 – 40.7 kWh of electricity. This study further estimated that
generating electricity from biogas could reduce GHG emissions from the palm oil milling
process by 110 – 175 kg CO2 / tonne of FFB processed. However, 58% of the mills in Malaysia
have chosen to flare the biogas (to convert the CH4 to CO2 and hence reduce its GHG emission
potential) to comply with the environmental requirement for biogas disposal (Loh, 2017). The
flaring of biogas, which has comparable combustible properties to natural gas, is considered as
resource inefficient and environmentally unsustainable under the Economic Transformation

30
Program (ETP) (PEMANDU, 2010; Loh et al., 2017a). The Palm Oil National Key Economic
Area (NKEA) has been introduced since 2010 to facilitate biogas recovery initiatives with an
aim to mitigate approx. 18 million tonnes CO2eq annually.

Whilst action has been and is being taken on methane emission from POME, to date, there has
been little discussion about control measures for the potential GHG emissions for EFB, and no
specific restriction has been imposed except for banning incineration of EFB since 2000
(Yusoff, 2006; Hashim and Ho, 2011; Elmer and Nygaard, 2014). EFB mostly were left to
naturally biodegrade and only minimal amount were returned to plantations for mulching and
as a soil enhancement agent (Menon, Rahman and Bakar, 2003; Teh, 2016). Hansen, Olsen and
Ujang (2012) indicate that improper EFB biodegradation management is a potential source of
methane and nitrous oxide emissins. Stichnothe and Schuchardt (2011) estimated that each
tonne of EFB left at the dumping site for biodegradation emits approx. 130 kg CO2eq to 980 kg
CO2eq.

2.2 Generating RE from POMR in Malaysia

The generation of RE from POMR involves a value chain consisting of residues acquisition,
RE generation (for use in POM operations; surplus for export to the grid or other local use) and
RE transmission and distribution (Figure 2.2). This value chain is sandwiched between two
highly competitive and profitable sectors (e.g. CPO production and conventional electricity
generation) as mention in Chapter 1. This situation has made the value chain management and
implementation process more complicated, rigid and restricted. According to Umar, Jennings
and Urmee (2013), the implementation of the POMR to RE generation value chain requires
meticulous operational and financial planning (Umar, Jennings and Urmee, 2013).

RE Transmission
Residues Supply RE Generation
and Distribution

Figure 2.2: POMR to RE Generation Value Chain (Umar, Jennings and Urmee, 2013)

The performances of the POMR to RE generation value chain has not been thoroughly
investigated. According to Sovacool and Drupady (2011), Shamsuddin (2012) and Mukherjee
and Sovacool (2014), the rather limited RE generation from POMR to date stems from the lack

31
of technical and techno-economic feasibility across the value chain. Meanwhile, a detailed
study on the value chain by Petinrin & Shaaban (2015) and Oh et al. (2017) found that there is
a physical mismatched between the chain structures which leads to misalignment of the goals
and objectives across the value chain. A major cause of such misalignment arose from the
inability and unwillingness of the PO millers to share information leading to an ambiguous
relationship through the chain structures (Darshini, Dwivedi and Glenk, 2013; Elmer and
Nygaard, 2014). Most of the plantations and mills in Malaysia are affiliated with major
developers with limited access to specific information; therefore, information sharing is not
possible (Umar, Jennings and Urmee, 2013). Some of the millers and plantation operators’
perceived information sharing as a business threat which leads to unhealthy business
competition (Oh et al., 2017).

In addition to the above relationship, other factors have been cited as constraints in the RE
value chain. These include inconsistent supply of POMR due to uncertainty of availability at
the mills, poor efficiency of the energy conversion technology, high CAPEX, limited access to
reliable financial assistance (government incentives and commercial loans) unfavourable FiT
compared with the cost of generation, and limited grid access (Sovacool and Drupady, 2011;
Shamsuddin, 2012; Mukherjee and Sovacool, 2014).

In view of the above, the process of generating RE from POMR is perceived to be high risk,
unfeasible and unprofitable (Yatim et al., 2017). Umar, Jennings and Urmee (2014b) reported
86% of the 85 palm oil millers surveyed have a pessimistic opinion on generating RE from
POMR and they prefer to use the excess POMR for other purposes. Furthermore, Umar,
Jennings and Urmee (2014b) highlighted that 90% of their respondents have included the
uncertainty over the feasibility of effectivePOMR to RE value chain in their consideration in
deciding whether to venture into the RE generation business. On the other hand, 75% of the
respondents expressed their readiness to collaborate with the government in achieving the RE
generation targets from palm oil biomass if the workability and profitability of RE generation
can be ascertained. Meanwhile, the responses from the millers interviewed during fieldwork in
the present study suggest that there is a need to conduct meaningful feasibility study with the
miller in order to attract their interest and confidence in the generation of RE from POMR
(Tenaga Sulpom, pers. comm., 27/09/2017; Havys Oil Mill Sdn. Bhd, pers. Comm.,
06/06/2016). In the light of the above, a pressing concern appears to be the need to evaluate
the feasibility of the overall value chain as part of building confidence and interest in the
process of delivering RE generation projects.

32
2.2.1 Residues Supply

The actual availability of POMR for RE generation is still somewhat vague, controversial and
inconsistent. This is because, in essence, the availability of the residues has been reported as
an aggregated value at the national level based on historical data or extrapolated theoretical
estimations (Sulaiman et al., 2011; Abdullah and Sulaiman, 2013). According to Batidzirai,
Smeets and Faaij (2012), the estimations based on such aggregated values have a high tendency
to be overestimates. Their study emphasized that the availability of residues is subjected to
their technical, market, sustainability and implementation constraints: therefore, the amount
that can be implemented in practice in a specific time frame is much lower than the anticipated
theoretical estimate (Figure 2.3) (Batidzirai et al., 2016; Hennig, Brosowski and Majer, 2016).
IRENA (2015) also pointed to the over-estimation concern when the estimation is made based
on the aggregated values of the residues availability. These two arguments are consistent with
the findings of Karaj et al. (2010), Perpiñá et al. (2009), Alfonso et al. (2009) and Singh,
Panesar and Sharma (2008). These studies concluded that the availability of biomass is
subjected to various constraints and, as such, only some portion of the biomass being
considered as collectible, could be utilized for energy purposes.

Theoretical Availability

Technical Availability

Economic Availability

Sustainable Availability

Implementation Availability

Figure 2.3: Relationship between the Theoretical Availability of Residues and Constraints on
Their Availability

The variation in the quantity of available POMR has adverse effects on the feasibility of
generating RE. In an on-site performance observation of POMR RE generation plants, Nasrin
et al. (2011) reported that four out of six plants were operated below their design capacity and

33
below their licenced capacity due to insufficient availability of residues. These findings
emphasize the need for realistic and accurate estimates of the availability of residues for RE
generation. However, while in-field data measurement such in Nasrin et al’s research and field
visits in this research can provide more reliable information about the availability of the
residues, it also found in conducting the present research that the majority of the millers were
unprepared/unwilling to share such information. This limits the availability of reliable data. In
the present study, a mixture of field data collected during the field visits and inputs from the
recent literature and sectors reports has been used in an attempt to take account of such issues.

2.2.2 Energy Conversion Technology

Renewable electricity generation from POMR is an immature development in Malaysia. The


still limited technical experience and expertise, means that this energy conversion technology
is vulnerable to the effects of technical failure (Shafie et al., 2012). Generating electricity from
biomass typically involves four energy conversion steps, converting the chemical energy in the
feedstock to thermal energy, using the thermal energy to raise steam, extracting the thermal
energy of the steam into the kinetic energy of the turbine and using the kinetic energy of the
turbine to generate electricity (Chiew & Shimada, 2013; Chanlongphitak et al., 2015). In this
light, even though it mimics conventional fossil-based electricity generation process, selecting
and operating the appropriate technologies is challenging due to differences in the chemical
composition of the residues (Meiller et al., 2017) and other factors.

Numerous published studies have described the functionality of the energy conversion
technologies – combustion, gasification, and pyrolysis - from various perspectives. Most
studies have been conducted on both simulation or prototype bases and Ahmed et al. (2015)
and Faaij (2006) note that there are different types of technologies that can be used with
different types of biomass depending on the end-use of the energy generated. In general,
combustion technology is the most mature and widely used followed by gasification and
pyrolysis (Chang, 2014; Samiran et al., 2016). Combustion involves a simple operation of
directly burning the biomass in the boiler. This process is less complicated and has simpler
maintenance requirement than the other thermochemical energy conversion technologies. As a
result, combustion technology has become the preferred technology to convert the chemical
energy content of biomass into the thermal energy (Tsai, 2012; Hosseini and Abdul Wahid,
2014b). It is the only widely proven technology that practically produces heat and power at
lower parasitic demand (Patel et al., 2011; Wan et al., 2016; Qian et al., 2017). Moreover, due

34
to the relative simplicity of the equipment structure and operation, combustion technology has
a reasonable installation and operation cost. Wan et al. (2016) appraise combustion technology
positively for having the highest technical, economic and environmental performance among
the other technologies. Similarly, Patel et al. (2011) and Patel, Lettieri and Germanà (2011)
further emphasised the superiority of combustion technology in converting the chemical energy
content of biomass into thermal energy. It has been reported that combustion technology
coupled with a backpressure steam turbine is the most common arrangement used in the
Brazilian sugarcane industry (Dantas, Legey and Mazzone, 2013). Recently, Ozturk et al.
(2017) acknowledge there has been a shift to use condensing extraction steam turbine (CEST)
to produce higher pressure and temperature steam for electricity generation n Brazilian
sugarcane mills.

Despite these appraisals, combustion technology is known to be sensitive to the moisture


content of the feedstock, with wet feedstock causing incomplete combustion (Yarnal and
Puranik, 2010; Miccio et al., 2014). Incomplete combustion increases the emission of carbon
monoxide and carbon particulates instead of carbon dioxide. Incomplete combustion also
exposes the boiler to the effects of slagging, fouling and erosion, and increases the formation
of ash (Panwar, Kaushik and Kothari, 2011; Perez-Jimenez, 2015). Mechanical (e.g. shredding
and pressing) and formal drying processes (e.g. rotary drum, fluidized bed) can reduce the
moisture content of the biomass and thus reduce the possibilities of incomplete combustion
(Brammer and Bridgwater, 2002; Gebreegziabher, Oyedun and Hui, 2013). In Malaysia,
approx. 80% of POMs use combustion technology for their internal boiler (Umar, Jennings and
Urmee, 2014b). Combustion technology is also used in both of the POMR-based power plants
in Perak and Selangor (Tenaga Sulpom pers. comm., 27/09/2017; SEDA pers. comm.,
21/08/2017).

Gasification technology is claimed to have a higher energy conversion efficiency than


combustion technology, but requires more stringent technical operation to convert the biomass
into a combustible gas that can be burnt to produce heat and steam (Caputo et al., 2005).
Fluidized bed gasifier have higher rate of feedstock gasification with shorter finishing time
with minimum formation of ash than conventional combustion system (Hosseini and Abdul
Wahid, 2014b; Mahinpey and Gomez, 2016). However, higher formation of tars due to low-
temperature condensation has been a major disadvantages of this technology (Tsai, 2012). Tsai
(2012) indicated the formation of tar lowers the heating value of the produced gas. Typically,
pre-treatment of the residues is required to reduce their moisture content prior to gasification

35
in order to reduce tar formation (Samiran et al., 2016). Sansaniwal, Rosen and Tyagi (2017)
concluded that further research and design studies are required to develop highly efficient
gasification reactors that successfully demonstrate the feasibility of the technology. It has also
been noted that transforming the energy content of the feedstock using gasification technology
is not yet widely deployed, and the technical and economic performance improvement is
therefore very uncertain (IRENA, 2018b). In this light, only four mills in Malaysia are using
gasification technology to sustain their energy consumption (Umar, Jennings and Urmee,
2014b).

The use of pyrolysis for converting biomass energy is in its infancy currently in Malaysia.
Pyrolysis is a subset of the gasification process which comprises converting the solid biomass
into liquid, solid and gaseous fuel in the absence of air (Alias, Ibrahim and Hamid, 2014). It is
worth nothing that despite having a similar energy content as crude oil, the pyrolysis liquid has
to be treated before it can be used for electricity generation. Similarly, both the solid and
gaseous fuel can be used to generate heat and steam for electricity generation after certain pre-
treatment (Johnson et al., 2009; Chen et al., 2017). This technology is still in the proving of
concept stage (i.e., technology readiness levels three (TRL 3)) (IRENA, 2012, 2018b).

Thus, combustion technology has been appraised to have convincing performance since
biomass-based power generation by this route has gone mainstream globally (Jenkins et al.,
1998; McIlveen-Wright et al., 2013). Bazmi, Zahedi and Hashim (2015) credited combustion
technology as having the lowest capital and operating cost in Malaysia. IRENA (2012) notes
that every technology has different performance and cost curves and that combustion is the
technology which is commercially available with the lowest anticipated capital cost. Due to
these circumstances, combustion technology appears the preferred route currently to convert
the energy content of biomass for electricity generation purposes and has been the focus in this
thesis.

2.2.3 RE Transmission and Distribution

The electricity supply industry in Malaysia works on a privately-owned vertically-integrated


utility system where the process of generation, transmission, and distribution of electricity is
owned and managed by the same utility (Hassan et al., 2008; Sovacool and Drupady, 2011).
Tenaga Nasional Berhad (TNB) operated under a single buyer market model before the
managed market model (3M) was introduced in early 1990s governed by the Electricity Supply
Act 1990. This new model in the electricity supply system was introduced to usher in the

36
national aspiration to reduce the country’s over-reliance on fossil fuel and to meet rising
demand for electricity (Hassan, 2009; Energy Commision, 2013; Ngadiron et al., 2015) in
conjunction with an announcement of renewable energy as the ‘fifth fuel’ after oil, gas, coal
and large-scale hydro. As a result, TNB is obligated to purchase RE from independent power
producers (IPPs) to significantly increase the share of RE in the electricity generation mix
(Hashim and Ho, 2011). Figure 2.4 shows the structure of Peninsular Malaysia’s electricity
supply system.

The introduction of 3M has resulted in a more regulated electricity supply where TNB holds
60% of the total generation capacity, and the IPPs have the provision to generate the remaining
40% (Jaafar, Kheng and Kamaruddin, 2003). Furthermore, the reformation of the electricity
market model has positively boosted the total generation capacity and provided competitive
electricity cost. However, the electricity supply system has become more complex involving
multiple stakeholders including the Energy Commission, national and state utility companies,
IPPs, co-generators, the Energy Planning Unit and the Ministry of Energy, Green Technology
and Water (Loh and Abdul Majid, 2016).

37
Energy Commission

Independent TNB Independent TNB


Generation Power Producers Generation Co-generators
Power Producers Generation

Transmission TNB TNB


Transmission Transmission

TNB Local Distributors TNB


Distribution Distribution Distribution

Customers Customers Customers

Figure 2.4: Peninsular Malaysia Electricity Supply System

38
There has been an increasing amount of research that argues that RE is discriminated against
in term of gaining access to the national grid. In 2004, Beck and Martinot described limited
grid accessibility as a traditional barrier for RE deployment and one that will remain a barrier
that needs to be addressed through the conscious action of multiple actors. Similar difficulties
in RE accessing the national grid to those experienced in Malaysia are also experienced in a
number of other countries e.g. United Arab Emirates (Al-Amir and Abu-Hijleh, 2013).
Denholm and Hand (2011) and Namuli et al. (2013) suggested that government intervention is
necessary to reconcile the grid accessibility issue for RE projects.

Petinrin and Shaaban (2015) emphasised that the divergence of interest between the three main
actors – the utility company, government and potential investors – has prevented a proper
coalition between them. Unbalanced burdens and responsibilities pertaining to new grid
installations or extensions have also been identified as factors leading to the failure to deploy
new grid-connected RE projects (Borhanazad et al., 2013; Asia-Pacific Economic Cooperation,
2014). One way to address the long-standing issue on the new grid installation and extension
is to establish a reasonable responsibility sharing between the electricity generators and the
grid operators (Hiroux and Saguan, 2010). In Malaysia, there is still no clear direction to resolve
this issue, and until today, everything related to grid connection is under the full responsibility
of the electricity generators. Additionally, a burdensome application and approval processes
for renewable energy power purchasing agreements (REPPAs) is a further reason reducing
interest in participating in RE generation (Abdul Malek et al., 2017; Loh et al., 2017a).

2.3 Barriers and Opportunities for RE Generation from POMR in Peninsular Malaysia

The generation of RE from POMR in Malaysia has been driven by two solid motivations:
energy security and environmental protection commitments. Success in implementing the RE
generation potential depends on the feasibility of the value chain and the policy and regulatory
‘environment’ in which such RE projects develop. Fiala, Pellizzi and Riva (1997) and Negro,
Alkemade and Hekkert (2012) stated that the speed, direction, and success of RE development
is highly dependent on the environment where it is implemented, rather than the deficiency of
the energy conversion technology adopted. According to these two studies, RE generation from
POMR has to be developed in a ‘conducive environment’. Even though the term ‘conducive
environment’ has been frequently used in the literature, there is still no fixed definition since
the term has been applied to suit the needs of each individual study. Thus, while a variety of
definitions has been suggested, the term ‘conducive environment’ is used throughout this thesis

39
to refer to a policy and regulatory environment which allows the RE to be treated as fairly as
the existing fossil-based electricity (Bakhtyar et al., 2013; Petinrin and Shaaban, 2015; Alam
et al., 2016).

The absence of such ‘conducive environment’ is considered to be as system failure that


required transformative change to reduce/remove barriers (Negro, Alkemade and Hekkert,
2012). Negro, Alkemade and Hekkert (2012) describe five types of the system barriers: market
structure, infrastructural, institution, interaction, and capabilities. Moreover, Trudgill and
Richards (1997) developed the Trudgill’s Framework for Analysis in which the barriers
relating to environmental and energy issues are more often regarded as non-technical barriers
and divide into six categories – agreement, knowledge, technology, economic, social and
political – the so-called ‘AKTESP’ barriers.

In the case of Malaysia, the lack of ability to be convincing on the feasibility of RE generation
and a strong recognition of a non-conducive implementation environment has led to the
sporadic performance and limited success of RE generation models in the country. The non-
conducive environment is caused by various barriers that are classified into three main
categories: operational, financial and institutional (Kardooni, Yusoff and Kari, 2016; Nizami
et al., 2017). In this light, a positive operational, financial and institutional environment has
allowed RE generation technology to grow rapidly in several European countries and the
United States of America. This growth is also driven by active participation and coordination
from various stakeholders (Staffas, Gustavsson and Mccormick, 2013; Abdmouleh, Alammari
and Gastli, 2015; Polzin et al., 2015).

Curnow, Tait and Millar (2010) argued that RE generation commonly faces a combination of
at least two barrier categories, and this duality has been difficult to overcome. Curnow’s
argument was based on a comprehensive review of RE generation projects in Asia which can
be considered the best representation of the current context of RE generation from POMR in
Malaysia. Most published works have described the combination of operational and financial
barriers as the main ones hindering the development of POMR to RE generation technology in
Malaysia which has largely remained stalled at the take-off phase (Mustapa, Leong and
Hashim, 2010; Kardooni, Yusoff and Kari, 2016). Similarly, in Indonesia, the efforts to expand
the implementation of POMR-based electricity generation has been restricted by financial and
operational constraints resulting in only a marginal contribution to the electricity generation
mix (Wicke et al., 2011; Setyawan, 2014).

40
Although the financial structure and requirements for RE generation have been sufficiently
documented, they are always found to be uncompetitive compared with conventional electricity
generation (Ujang, 2018). Due to high-risk perception on RE generation and lack of
understanding of the RE financial requirements, most financial institutions have enforced
complex loan application processes in addition to high-interest rate (Lidula et al., 2007; Maulud
and Saidi, 2012; Sen and Ganguly, 2017). For instance, a period of four to six months is needed
for the loan to be processed and an annual interest rate of 12% is imposed on the loan amount
(Tenaga Sulpom pers. comm, 27/09/2017; Havys Oil Mill Sdn. Bhd pers. comm. 06/06/2016).
Comparably, the interest rate imposed on the fossil-fuel based electricity generation is around
4.5% (Ujang, 2018). Held, Ragwitz and Haas (2006) suggested 6% as the maximum interest
rate to accelerate the RE generation development and countries like Thailand and Germany
have taken proactive measure to offer 4% interest rates to stimulate the RE sectors in those
countries (Oschmann, 2010; Gruning et al., 2012; Rahman, Saat and Wahid, 2016). In a similar
vein, the Malaysian Government introduced the Green Technology Financing Scheme in 2011
which offered a 2% rebate on the interest rate. However, this scheme was terminated in 2015
due to limited budget allocation and a low number of successful business cases (Oh et al.,
2017). Therefore, providing competitive financial structure to address the present financial
disadvantages of RE is essential. There are several successful RE financial structure, for
instance, the Energy Efficiency Revolving Fund (EERF) in Thailand (Gruning et al., 2012) and
Community Energy Initiatives in Germany (Lutz et al., 2017) which have shown encouraging
outcomes. For instance, 294 RE projects were financed by 13 public and local banks
participated under EERF in Thailand with total financial savings estimated to be US$ 169
million per year (Jue, Johnson and Vanamali, 2012). Therefore, a new financial structure for
RE projects in Malaysia is worthy of consideration. The new financial structure should seek
to: (1) create a mutual network agreement between the millers and commercial financial
institutions on the RE financial requirement; (2) stimulate the investment appetite among the
millers; (3) enhance the financiers understanding of the RE funding requirements; (Khor and
Lalchand, 2014; Mafakheri and Nasiri, 2014).

Umar, Jennings and Urmee (2013), Yatim et al. (2016) and Oh et al. (2017) pointed out the
need to improve the current RE institutional setting in Malaysia (e.g. policies, incentives,
regulatory and local advocacy assistance). According to these studies, the current institutional
settings are ineffective at reducing the uncertainties over the RE generation, creating a
substantial market for RE, increasing the accessibility to financial resources and allowing

41
profitable RE generation. One common example is that the current feed-in-tariff rates in
Malaysia have failed to fairly incentivise the RE generators and as a result, it has discouraged
the exploitation of the RE in the country has not been adequately simulated. Acknowledging
the existence of a comprehensive set of RE policies in Malaysia (see Appendix A), Chang,
Fang and Li (2016) highlighted that these are dissociated from one another. In Australia, such
disassociation between RE policies has led to a range of ad-hoc and often contradictory policy
pathways and has constrained the progress of RE generation deployment there (Warren,
Christoff and Green, 2016). Likewise in Malaysia, although there are various policies, action
plans, and national strategies for the development of RE, they have not yet been able to provide
clear policy direction or policy achievement pathways to guide the development progress
(Umar, Urmee and Jennings, 2017). In addition to policy disassociation, Yatim et al. (2016)
and Oh et al. (2017) acknowledged the policies were developed with little attentions given to
minimizing the techno-economic performance uncertainties.

A shift in the policy formulation approach in Malaysia is essential to unlock the potential and
promote the more sustainable development of RE generation. A shift from the norm of the
policy formulation approach to a more scientifically rigorous and context-sensitive manner has
been identified as paramount to open the windows of opportunity for RE development (Hatch,
2010; Yatim et al., 2016; Oh et al., 2017). For the case of RE generation from POMR, a valid
proof of the techno-economic performances and environmental benefits from such generation
is essential to establish a more fully-informed policy formulation in Malaysia. Such policy
formulation can be expected to enhance the feasibility of POMR to RE value chain and help to
provide a ‘conducive environment’ to promote wider development of POMR to RE projects.
Moreover, it allows more realistic setting of short- and long-term national RE targets from
POMR, and at the same time the better formulation of policy delivery and management
strategy. Therefore, a comprehensive techno-economic and LCA for POMR to RE generation
to examine the suitability and economic viability of POMR as a fuel for various scale RE
generation is warranted to complement the policy formulation before it can come into
realisation.

42
2.4 Summary

This review has highlighted the need for inter-linked solutions between the operational,
financial and institutional domains to support effective utilization of POMR for RE generation
in Peninsular Malaysia. The limited development to date of the conducive environment is, in
part, due to a shortage of scientific evidence showing the feasibility of POMR to RE generation
systems. It was observed that the two existing POMR to RE generation systems established in
Peninsular Malaysia are based solely on the experimental approach, and there is lack of proper
underlying modelling, design, operational requirements and guidance. Such information can
act to reduce the possibility for impractical RE generation projects by identifying a realistic
limit for residue requirements for generating profitable electricity. This allows the millers to
conduct first stage competency assessments by comparing their amounts of available residues
in their mills with the minimum requirement to secure a profitable electricity generation. In a
broader extent, this can help to facilitate the RE generation quota allocation, knowledge sharing
and efforts to avoid duplication among interested millers and other stakeholders. Furthermore,
this will lead to better planning of the resources, more streamlined coordination of the resource
supply chain, improvement in the economic performance of bioenergy, increasing emphasis on
the environmental advantages and provision of improved ability to verify the overall
performance. Importantly, such credible scientific and practical demonstrations will serve as
the basis to support the ongoing RE policy development initiatives.

In this respect, there is a need for a systematic study to bridge the gap between these
operational, financial and policy factors. This serves as a primary motivation for this research
work.

43
CHAPTER THREE

POMR-RES: A CASE STUDY FROM PENINSULAR MALAYSIA

The potential for generating renewable electricity from POMR has received policy support
from the Malaysian Government for almost two decades. However, uptake of the technology
is still relatively low. A significant issue that has dominated the discussion for many years is
how to translate the renewable electricity generation potential from POMR into actual
implementation. Although there are many appealing reasons and considerable opportunities to
venture into the POMR-based electricity generation business, there are also substantial barriers
that inhibit the realization of this potential

3.1 Scope of Research

The site-specific case studies carried out in this research focused on two major residues from
the palm oil milling process; empty fruit bunches (EFB) and palm oil mill effluent (POME).
Based on the total quantity generated by the palm oil sector, these residues have been
acknowledged as having considerable electricity generation potential: 521 MW and 320 MW
respectively (Ali, Daut and Taib, 2012). The present research has been designed according to
the current renewable energy outlook in the Peninsular Malaysia and takes account of the
availability of such residues and the respective energy conversion technology.
Several factors have led to the selection of Peninsular Malaysia as the study area as a defined
geographic part of Malaysia with an established palm oil industry. Appropriate access to the
high-level data required for the study is available since most of the related government agencies
such as MPOB, SEDA, and EC are located in Peninsular Malaysia. There is clear separation in
the electricity supply system between Peninsular Malaysia and East Malaysia especially
Sarawak. The electricity supply system in Peninsular Malaysia is under the governance of
TNB, Sabah Electricity Sdn Bhd (SESB) for Sabah and the Sarawak Electricity Supply
Corporation (SESCO) for Sarawak (Energy Commission, 2010; Institute of Strategic and
International Study, 2014) . Due to the separation of the electricity supply system, the feed-in
tariff (FiT) system is not applicable for Sarawak, and lastly, the target for renewable-based
electricity generation is only specified for Peninsular Malaysia and Sabah as mentioned in the
Introduction chapter (Economic Planning Unit, 2015).
The methodology applied in this research is appropriate for new on-site, off-site and
cooperative electricity generation system installations with direct combustion as the primary

44
energy conversion technology as suggested in the literature (McIlveen-Wright et al., 2013;
Hosseini and Abdul Wahid, 2014a). The on-site installation (OnS Case Study) refers to the
POMR-RES development being owned by and in the vicinity of existing mills and using EFB
and/or biogas feedstock supplied internally from the mill at no cost. For this OnS case, the
electricity and/or heat demand from the mills are supplied from POMR-RES, with surplus
electricity being exported to the grid. For the off-site installation (OffS Case Study), POMR-
RES is deemed to be located away from the mills and the EFB is considered to be transported
for 10 km from the supplying mills to the POMR-RES site. The generating plant is owned by
either the millers or independent investors. For the POMR-RES owned by millers, EFB
supplies are available for free, but with an additional EFB transportation cost. Conversely, for
those owned by independent investors, the EFB supplies have to be bought from the millers in
addition to the feedstock transportation cost. For OffS POMR-RES, all the electricity generated
is exported to the grid. On the other hand, cooperative installation involves a technology
sharing concept whereby the generation plant is shared by some existing mills and developed
in the most business-friendly location as suggested by Umar, Jennings and Urmee (2014b).
The methodology has been modelled broadly for small-, medium- and large-scale palm oil
mills in Peninsular Malaysia for an economic lifetime of 15 years. This approach and modelling
are suitable, however, for adaptation by other palm oil growing/process countries e.g.
Indonesia, Thailand, and Columbia.

3.1.1 Study Area

This research focuses on palm oil mills in Peninsular Malaysia. There were approx. 2.7 million
hectares of oil palm plantation across the peninsula in 2017 (Table 3.1) yielding an average of
1.6 tonnes fresh fruit bunch (FFB)/hectare (MPOB, 2017).

Table 3.1: Oil Palm Plantation Area (in ha) in Peninsular Malaysia
State Mature Immature Total
Johore 682,624 66,236 748,860
Pahang 641,876 99,619 741,495
Perak 360,501 45,968 406,469
Negeri Sembilan 162,634 22,181 184,815
Terengganu 146,561 24,987 171,548
Selangor 128,058 9,725 137,783

45
Kelantan 118,090 40,220 158,310
Kedah 82,421 5,117 87,538
Melaka 52,322 5,050 57,372
Perlis 617 43 660
Total 2,388,574 319,839 2,708,413

These plantations are owned by private estates (61%), government agencies (17%) and
independent smallholders (22%) (MPOB, 2017). There are 235 mills operated in Peninsular
Malaysia with Pahang and Johore having the most as shown in Table 3.2. As observed, there
are no palm oil mills in the capital city of Malaysia, Kuala Lumpur, and in the smallest state
on the peninsula, Perlis. This is due to Kuala Lumpur being heavily urbanized, and to Perlis
having limited space for palm oil plantation due to extensive sugar cane plantations.

Table 3.2: Palm Oil Mills According to State in Peninsular Malaysia 2015

State Number of Mills


Pahang 67
Johore 56
Perak 40
Selangor 22
Negeri Sembilan 15
Terengganu 13
Kelantan 10
Kedah 7
Malacca 3
Pulau Pinang 2
Total 235

The mills are scaled based on the FFB processing capacity which varies from 10 t/hr to 120
t/hr with average operating hours of 3500 hours annually (Nasrin et al., 2011). Mills with
processing capacity from 10 t/hr to 40 t/hr are classified as small-scale, 41 t/hr to 80 t/hr as
medium-scale and 81t/hr to 120 t/hr as large-scale. Table 3.3 shows the tabulation of the small-
, medium- and large-scale mills in Peninsular Malaysia. Most of the mills (59%) are classified
as small-scale, and only about 6% are large-scale mills.

46
Table 3.3: Palm Oil Mills According to Scale in Peninsular Malaysia

Scale Number of Mills


Small (10 – 40 t FFB/hr) 139
Medium (41 – 80 t FFB/hr) 81
Large (> 80 t FFB/hr) 15
Total 235

3.1.2 Types of Residues

Malaysia is the second largest producer of CPO globally after Indonesia and the world’s largest
CPO exporter. In 2014, the total export of CPO was reported at US$11 billion and continued
to register positive growth to US$ 16 billion in 2016, making CPO industries the highest
contributor to the exports of agricultural goods accounting for 9% share of total exports in the
year (MITI, 2014, 2015). The wet milling process of CPO, in summary, involves seven
sequential steps and general wet milling process of one tonne CPO is shown in Figure 3.1. The
production of one tonne CPO requires 5.2 tonnes of fresh fruit bunches (FFB) and generates
about 5.2 tonnes of residues (Abdullah and Sulaiman, 2013; Loh, 2017).

FFB
Mill
Natural
Sterilization Biodegradation

Electricity
Stripping EFB Mulching

Turbine LP Steam Fertilizer


Digestion

Boiler
Pressing POME

Fibre Separator
MF Anaerobic
Digestion

PKS Nut Cracker

Oil Clarification CPO

Figure 3.1: Wet Milling Process of CPO (Abdullah and Sulaiman, 2013; Loh, 2017)

47
The production of CPO in 2015 generated roughly 94.2 million tonnes of POMR which are the
combinations of EFB, MF, PKS and POME as presented in Table 3.4 (Energy Commission,
2016; Sadhukhan et al., 2018).

Table 3.4: POMR Availability in 2015 (Energy Commission, 2016; Sadhukhan et al., 2018).

Residues Availability (million tonnes)


EFB 21.24
MF 12.00
PKS 5.54
POME 55.40

These large quantities of residues have been perceived as a ‘wastes’ and are currently
underutilized, bringing little or no economic value (Ali et al., 2015). The MF and PKS are
currently used as the feedstock for the boilers in the mills for internal heat and electricity
consumption. However, the process has been made intentionally of a lower efficiency more as
a means of disposal of these residues than as a way of optimally using them to generate more
electricity and to curb carbon emission (Ali et al., 2015; Loh, 2017). Although operating at this
lower efficiency, most mills are considered as heat and electricity self-sufficient. This gives the
mills clear economic advantages as they avoid the cost of using other fuels to generate heat and
for buying electricity from the grid. In 2013, 396 MW off-grid electricity generation capacity
from MF and PKS was installed in the mills in Peninsular and East Malaysia (Abu Bakar,
2015).
No specific utilization routes have been designated by the government and industry for EFB
and POME (Abdullah and Sulaiman, 2013). With an indication that the palm oil industries will
continue to grow rapidly to meet the global demand for CPO, the associated amounts of EFB
and POME will pose an ongoing waste management issue to the mills. Saswattecha et al.
(2015) and Saswattecha et al. (2017) identified that the main source of CH4 emission from palm
oil production was from the inappropriate management of EFB and POME. Therefore, urgent
measures to overcome this issue are very much needed (Visvanathan and Chiemchaisri, 2006;
Fallis, 2013; Lam et al., 2013). A more efficient utilization pathway for EFB and biogas is
essential to maximize the potential of these resources and at the same time to reduce the effect
on the environment and society.

As shown in Figure 3.1, both of these residues have the least financial contribution to the
routine operation of the mills. The EFB has commonly been dumped to undergo natural

48
biodegradation since the ban on the incineration process by the Malaysian Government in 2000,
with a minimal amount being returned to the plantation area for mulching and as a bio-fertilizer
(Menon, Rahman and Bakar, 2003; Teh, 2016). The estimated amount of GHG emission for
the natural biodegradation of EFB varies depending on the anaerobic condition of the EFB
piles. With an assumption that biogenic CO2 is not taken into account (had been absorbed
during the growth phase) and only 49% of the carbon in EFB is biodegradable while the
remaining is converted to methane (and some of the N to N2O) (IPCC, 2006), at 10% (best case
scenario), 50% (baseline scenario) and 80% (worst case scenario) anaerobic condition, about
130 kg CO2eq, 650 kg CO2eq and 980 kg CO2eq are emitted from each tonne of EFB (Stichnothe
and Schuchardt, 2011). Although mulching of EFB residues is claimed to provide indicative
CO2 emission reductions, over-utilization of EFB for this purpose has about an equal amount
of CO2eq emission as that from natural biodegradation (Ministerie van Economie Landbouw &
Innovative, 2013). Approximately 35 t – 45 t of EFB is required for mulching one hectare of
plantation each year (Menon, Rahman and Bakar, 2003). It is estimated that over-utilization of
EFB for mulching emits about 4.7 tonnes CO2eq /ha/year and emission from the transportation
of the EFB to the plantation further increases the carbon footprint of the mulching activities
(Foo et al., 2013).

The most popular option for POME treatment has been so-called ‘open ponding’ approach as
it is the cheapest and least complicated treatment. This is borne out from review of the literature
and in personal communications between the author of this thesis and various millers.
However, the methane produced is a major problem with this method as it released directly
into the atmosphere (Madaki and Seng, 2013). The average emission from an open pond in
Felda Serting Palm Oil Mill with a FFB processing capacity of 54 t/hr has been reported as
1043 kg CO2eq /day (Yacob et al., 2006). Furthermore, Ahmed et al. (2015) estimated the
emission value of an open ponding treatment to be 1225 kg CO2eq /tonne CPO produced.

A further environmental concern over open ponding is the amount of space used to
accommodate a complete open ponding POME treatment, which usually consists of a de-oiling
tank and a series of ponds including acidification, anaerobic and aerobic ponds (Tong and
Jaafar, 2004; Yacob et al., 2006; Aziz et al., 2017). Chin et al. (2013) emphasized that the
number of ponds and space required for the treatment ponds is dependent on the FFB
processing capacity of the mills. The typical size of each treatment pond based on the
measurement made by Poh and Chong (2009) for a 54 t/hr FFB processing capacity mill is 60m
× 30m × 6m (length × width × depth) - the size of half football pitch.

49
Due to increasing environmental concerns, since 2010 the Malaysian Government has
compelled mills to treat the POME in closed anaerobic digester (AD) tanks and capture the
resulting biogas to overcome the drawbacks of open anaerobic ponding. This treatment method
for the POME started to emerge since the Department of Environment enforced the
Environmental Quality Act 1974 which requires the POME to meet four regulatory liquid
discharge limits: 1. biological oxygen demand should not be more than 100 mg/l; 2. suspended
solids content need to be less than 400 mg/l; 3. measurement of oil and grease should less than
50 mg/l; 4. pH value of the discharges need to be between 5 to 9 (Department of Environment,
1974).

The Malaysian Government’s enthusiasm to better manage the POME using closed AD tank
has received encouraging responses from the millers. Although the resulting biogas is now
being captured, it is flared to minimize the environmental impact and eliminate the unnecessary
auxiliary waste handling process (Basiron and Weng, 2004; Yeoh, 2004). According to Loh
(2017), 58% of the mills choose to flare the biogas as a short-cut approach to discard the biogas
and accord with the environmental requirement for biogas disposal. While flaring the biogas
reduces CH4 emission, it releases sulphur dioxide (SO2) which reacts quickly with water and
air to form sulphuric acid (Beér, 2000). Formation of large amounts of sulfuric acid increases
the potential for acid rain which can also severely impact the environment and society.

From the life cycle perspective, utilizing the EFB and POME for electricity generation appears
an attractive route for utilising residues to give a positive impact on both the environment and
the economy (Seabra and Macedo, 2011; Chiew and Shimada, 2013; Chanlongphitak et al.,
2015; Beaudry et al., 2017). However, there is still a need for an insightful evaluation of the
actual quantity of electricity generation potential from both residues, the economic feasibility
and investment profitability of such electricity generation, and environmental benefit from
utilizing these residues for electricity generation, especially in Peninsular Malaysia.

3.2 Research Framework Summary

The overall purpose of this study is to undertake the evaluation described above to provide
evidence and support for the translation of sources of electricity generation from EFB and
POME that can be positively identified as sustainable into implementation. Figure 3.2
illustrates the research framework established to achieve the aims and objectives described in
Chapter 1. The research began with a comprehensive literature review to identify the specific

50
research gaps, define the research scope and design the research framework. Next, the data
gathering process was undertaken through visiting selected mills, conducting interviews with
relevant stakeholders and clustering the pre-existing data available from credible sources, for
instance, official reports, policies, websites and academic publications. Two different data sets
were collected; palm oil mill statistical and spatial information and policy, regulatory and
institutional information.
The POMR-RES power and CHP plant systems were modelled using the information gathered
to assess the overall potential of the POMR for electricity or combined electricity and heat
generation. The power plant was modelled on maximising electricity generation while CHP
was modelled for simultaneous electricity and heat production.
Three overall scenarios were used for plant location: i) plants were installed on-site, ii) plants
installed off-site and iii) plants operated on a co-operative basis with installation at one of the
partner’s site. In the literature, there seems to be no general definition of these three types of
installation scenarios (Wood and Rowley, 2011; Sadhukhan, 2014). Generally, an on-site
installation is defined as installing the energy plant within the vicinity of the palm oil mills,
whereas off-site installation describes the energy plant is installed at a place away from the
mills (as implied in the name). The co-operative installation, on the other hand, was conceived
as a potential installation scenario through which a partnership between several smaller mills
might seek to overcome economy-of-scale limitations with the energy plant, installed at the
most business-friendly location to which POMR feedstocks could be supplied from other, local
partners.
The technical, techno-economic and environmental performance and feasibility of these three
installation scenarios was based on the specific criteria and assumptions set out briefly below:
1. The Technical performance and feasibility was based on:
1. Electricity generated is sufficient to accommodate the combined parasitic
load of a) the electricity generation process itself, and b) the mill’s
operational electricity demand. Surplus electricity is exported to the
Peninsular Malaysia National Grid via a grid connection.
2. Partial heat demand of the mill is supplied by the heat extracted from the
back pressure steam turbine.
2. The Techno-economic performance and feasibility was based on:
1. Positive ROI and NPV.
2. PP and break-even point within the economic life of the plants.

51
Sensitivity analysis was conducted to assess the effect of changing parameters on the cost and
revenue functions of the energy plant and its subsequent impacts on the ROI and NPV as well
as PP and BEP.
Therefore, in the study the suitability of POMR for electricity generation was first rationalized
based on technical and techno-economic features. This was followed by an ‘optimization’
phase in which, the scale of the energy plant and its output were evaluated against two
optimization objectives which are ensuring ROI by 20% and positive NPV.
3. The environmental performance and feasibility was conducted on the optimum size
POMR-RES to quantify the carbon footprint from every megawatt-hour of electricity
generated. The global warming potential (greenhouse gas balance as kg CO 2eq) were
compared with the current Malaysian electricity grid average and the GHG balance
from avoided of conventional EFB and/or biogas management was also determined.
The performance of the optimum size POMR-RES was also reviewed with regard to the
levelised cost of electricity, the number of eligible mills considered to be at an appropriate scale
to participate as independent power producers, POMR utilization rate and overall contribution
towards the Malaysian Government’s RE policy target. By identifying the optimum size
POMR-RES and with consultation with a number of stakeholders, implementation guidelines
for the millers, prospective investors and other stakeholders were developed. It is hoped that
an outcome of this study may be a reduction in the uncertainty and potential complexity for
developers in designing, operating and managing the POMR-RES. This may therefore make a
positive contribution towards achieving Malaysia’s national target for additional RE power
supplies and subsequently contribute to improved global sustainability.

52
To translate the renewable electricity generation potential from palm oil mill residues to focus on empty
Aim fruit bunches (EFB) and biogas into actual implementation

Design an economically feasible operating condition for the palm oil mill residues to renewable
electricity system (POMR-RES)
Study
Area

Peninsular Malaysia
Resources/Methods

Field Visit to the mills Pre-existing statistical data


Data

Interviews, govt. agencies, industries bodies etc

Palm Oil Mill Statistical and


Spatial Information:-
Data Collected

Number of Mills, Location of Policy, Regulatory and Institutional


Mills, Processing Capacity, Information:-
Operation Hour, Mill electricity RE Policy Target, Policy Roadmap,
and heat demand, Amount of Available Incentives and
available EFB and POME, In- Assistance, Technology
house proximate and ultimate Availability
parameters of EFB and
POME/biogas
Plant under
Study

POMR-RES Power Plant POMR-RES CHP Plant


Installations
Mode of

On-Site Off-Site Co-op


Assessment
Integrated

Technical Techno-Economic

Sensitivity Analysis
Residues Suitability
Assessment

EFB Biogas

53
Optimization Objectives:
Optimization
Process

1. POMR-RES providing 20% ROI in order to be classified as


economically feasible
2. POMR-RES providing 75% ROI to be to be consistent with
conventional palm oil cultivation.
Environmental

Assessment

1. GHG emission factor of the electricity from POMR-RES


Impact

2. Carbon reduction from POMR-RES


3. Carbon emission avoidance from conventional EFB and biogas
disposal
Performance

1. Cost of Per Unit Electricity


Indicator

2. POM Eligibility
3. POMR Utilization Rate
4. Contribution of roll-out towards RE Policy Target
Implementation &

Standard Implementation Guidelines for POMR-RES


Execution

Figure 3.2: Research Framework

3.3 Data Collection Approach

Field visits to selected palm oil mills, semi-structured interviews with personnel from relevant
government agencies (i.e. MPOB, SEDA, EPU, TNB, UNDP Malaysia) and industries bodies
(i.e. RSPO, MyBiomass, MBIC) and on-going data collection from literature were the methods
employed to obtain the data required for this research. The researcher visited several selected
palm oil mills in the state of Malacca, Johore, and Pahang to closely observe routine operations
and to conduct a spatial study of the mills. The primary purpose of the visit was to compile
first-hand data required to model the POMR-RES including the real-time availability of EFB

54
and biogas in the mills, in-house proximate and ultimate parameters of the residues, mills
processing capacity, operating hours, and mill electricity and heat demand to run their routine
operation. Besides the quantitative operational data, the spatial information about the mills was
also desired. Information on the existence of the infrastructure such as roads, grid connection
and vacant space to install POMR-RES was also obtained.

The visits to the mills were conducted upon receiving official approval from the millers, which
were obtained after the researcher had sent an email with detail description on the purpose of
the visit, the goal of the study, and the types of data needed and how it shall be used. A copy
of a data checklist form designed specifically for this data collection (see Appendix B-1) with
a data privacy and non-disclosure was attached to the email to expedite the approval process.
Out of ten emails that were sent to ten different mills for the first field visit in July 2015, only
two millers responded to the request. Only one miller granted access to their premises and
allowed data on EFB and biogas availability taken from their production reports to be used in
this study. The second field visit was conducted in February 2016. For this, twenty-one millers
were contacted; however, only two millers responded and agreed to provide access to
production data. The third attempt to gain access to the mills in June 2016 did not receive a
positive response from the millers.

The reasons behind the unfavourable response and unwillingness of the millers to share the
information required are understandable. Most of the mills are affiliated to well-established
parent companies that have specific and stringent data management procedure. This gives the
millers limited authority to share data with other parties. Furthermore, the millers perceived
information sharing as a potential business threat that may affect the performance of their
business. The most obvious effects would be for instance market competition, stealing ideas
and reduction in sales. It is also noted that there may also reluctance to share data due to
possible weaknesses in proper data management procedure.

A series of formal, semi-structured interviews and discussions was also conducted with the
respective personnel from the relevant government agencies to seek for any accurate pre-
existing data to be used for modelling the POMR-RES. The interview and discussion sessions
were conducted upon receiving official invitation from the agencies, which were obtained after
the researcher had sent an official letter from the university that acknowledged the purpose of
the interview and the data confidentiality concern. As is common practice, all the interviewed
personnel chose not to disclose their identity, affiliation and job title.

55
The interview sessions were also aimed to better understand the future direction of renewable
electricity generation from the viewpoint of policy makers, authorities and stakeholders. Semi-
structured interviews were used to allow those personnel to share the information on their own
terms besides encouraging a two-way communication between the interviewer and those being
interviewed. Although in essence semi-structured interviews can provide reliable and
comparable qualitative data, the data collected during these interview sessions was somewhat
mixed and inconsistent. Further communication via email between the researcher as the
interviewer and the personnel was used to clarify the data inconsistency issues and was largely
successful at resolving such issues.

Data were also collected from the extensive literature review of the relevant academic
publications, official reports, and relevant policies. These provided useful source for
combination and verification/triangulation with the data collected from the field visits and
interviews.

56
CHAPTER FOUR

MODELLING OF POMR-RES

This chapter presents the modelling approach of POMR-RES in two different configurations:
1. POMR-RES power plant that emphasises on maximizing electricity generation capacity; 2.
POMR-RES CHP plant that produces heat and electricity simultaneously. The following
sections detail the modelling procedure and the outcomes from the modelling of the two
POMR-RES configurations.

4.1 Introduction

POMR-RES was modelled based on the established waste-to-energy scheme that converts
waste to electricity and useful heat (Yilmaz and Selim, 2013; Ozturk et al., 2017). The POMR-
RES power plant and POMR-RES CHP plant models are shown in Figure 4.1 and Figure 4.2
respectively. The system consists of a boiler, back pressure steam turbine, condenser and feed-
water pump. The system works in a fundamentally similar way to the steam-based power plant
examined by Sadhukhan et al. (2009) and Wan et al. (2016) where the steam is generated by
combusting the EFB and biogas under full combustion condition to release the heat energy
from the residues. The heat energy increases the temperature of the water in the boiler and turns
it into a high-pressure superheated steam (HPSHS) at 50 bar, 500°C (a higher temperature
HPSHS generation is possible; temperature of HPSHS has been kept at this level because of
practical consideration (Sadhukhan et al., 2009)). The steady flow HPSHS is piped to the back
pressure steam turbine (BPST). In the adiabatic condition, BPST converts the heat energy to
kinematic energy by rotating the turbine blades and shaft to generate electricity. In the POMR-
RES power plant, the exiting steam of BPST (~1 atm pressure, 105°C) is cooled in the
condenser and pumped to the boiler as boiler fed water (BFW), and is returned to the boiler for
continuous steam generation. The flue gas from the boiler at a temperature of 120°C is released
through the chimney to the atmosphere (Wan et al., 2016). The bottom ash of the boiler is
collected and assumed to be utilized for other useful purposes (Abdullah, Sulaiman and
Gerhauser, 2011; Kim and Lee, 2011).

In contrast, for the POMR-RES CHP plant, the exiting steam at medium pressure (~2.5 bar,
145 °C) is channelled to the sterilizer in the mill to cook the FFB. The lower pressure steam is
condensed into BFW. In Wan et al.’s (2016) study, main-stream steam extraction approach has

57
been used, where the HPSHS from the boiler is supplied to the mill. This approach, however,
has been challenged by the millers for two reasons: first, the mill does not require high-pressure
heat to cook the FFB; therefore, auxiliary equipment such as a pressure reducing valve are
required; second, the portion of HPSHS being channelled to the mill is more useful to drive the
turbine to maximize the electricity generated (Tenaga Sulpom pers. comm., 27/09/2017). From
the thermodynamic perspective, HPSHS has a low heat transfer rate which makes it unsuitable
for heating purposes; the de-superheated process is required to bring the steam to saturation
point (Simões-moreira, 2012).

Figure 4.1: POMR-RES Power Plant

Figure 4.2: POMR-RES CHP Plant

As a resource-driven energy generation plant, the amount of output energy generated is


estimated based on the flow rate of the feedstock, the energy content of the feedstock and the
efficiency of the energy conversion process (Mckendry, 2002; Patel et al., 2011). In that sense,

58
modelling the POMR-RES relies on the availability and consistency of EFB and biogas flows
and their energy content (or calorific value) (Hussain et al., 2006; Gebreegziabher et al., 2014;
Wu et al., 2017).

4.2 POMR-RES Modelling

POMR-RES was modelled taking into consideration, the complete energy equation using three
different approaches: first, mathematical modelling for the power plant; second, the
combination of mathematical and computer-aided process simulation for the CHP plant; third,
Malaysia Palm Oil Mill Standard Energy Calculation Method established by MPOB (Wang et
al., 2009; Ng et al., 2011; Wan et al., 2016). The POMR-RES modelling is expected to provide
a technical dimension of the systems by providing few core technical output parameters. These
included the amount of extractable heat from the combustion process, the accumulated amount
of HPSHS generated, and the electricity generation potential from the expansion of HPSHS.
These parameters were used as the comparators for the technical feasibility assessment and as
the basis to estimate the electricity generation cost and emission associated with it.
Both the POMR-RES power plant and CHP plant were modelled identically regarding the
hardware arrangement and operating principles except for the turbine’s outlet steam setting.
For the power plant model, the exiting stream of BPST was set at one atmospheric pressure
and the temperature of 105°C that represents a condensed water condition (Holmberg,
Ruohonen and Ahtila, 2009). The exiting stream for the CHP plant was set at 2.5 bar, 145°C
to accommodate the heat requirement for the palm oil mills (Tenaga Sulpom pers. comm.,
27/09/2017). The primary goal of modelling POMR-RES in two different configurations is to
study comparison of the technical, economic and environmental performances between them.
From this, the future developer could have better insights into how the system will work, what
to expect from the system and which system is in a good agreement with their objectives for
installation to which site.

4.2.1 Mathematical Modelling

A spreadsheet of the POMR-RES power plant predictive model was developed to estimate the
electricity generation potential from the combustion of the semi-dry and biogas as shown by
Sadhukhan et al. (2009). The sequential steps of estimating the electricity generation potential
in this research follow the thermodynamic principle of steam-based power plant suggested in

59
many applied thermodynamic analyses (Borgnakke and Sonntag, 2009; Sarkar and
Bhattacharyya, 2012). Patel et al. (2011) and Wan et al. (2016) have used this approach in their
study to assess the performance of biomass-based electricity and heat generation at different at
different scales. The modelling process start with the prediction of the amount of heat extracted
from the combustion process (𝑄𝐶 ) using Eq. (1)

𝑄𝐶 = 𝑚𝑓 × 𝐶𝑉𝑓 × 𝑛𝑏𝑜𝑖𝑙𝑒𝑟 (1)

where 𝑚𝑓 is the flow rate of the feedstock in kilogram per second (kg/s), 𝐶𝑉𝑓 is the calorific
value of the feedstock in mega Joules per kg feedstock (MJ/kg), and 𝑛𝑏𝑜𝑖𝑙𝑒𝑟 is the thermal
boiler efficiency. The flow rate and calorific value of EFB were used for EFB-based POMR-
RES while the corresponding values for biogas were used in biogas-based system.
In the present research, the EFB is considered as semi-dry EFB with the moisture content of
45%. The 15% moisture content reduction is anticipated from mechanical drying process where
the EFB is shredded into short fibres before pressed to squeeze out the remaining oil and water
mixture (Tenaga Sulpom pers. comm., 27/09/2017). The mass of the semi-dry EFB decreased
by 23% from the mass of the raw EFB after the mechanical drying process (Luk et al., 2013).
Most of the literature suggested the CV of the dry EFB (18.8 MJ/kg) with little or no mention
of the CV of semi-dry EFB (Sulaiman et al., 2010; Luk et al., 2013; Loh, 2017).
Chanlongphitak et al. (2015) quoted the CV of EFB at the collection point as 8.00 MJ/kg and
Abdul Malek et al., (2017) used 8.80 MJ/kg as the CV in their analysis. In this modelling, a
CV of 8.5 MJ/kg semi-dry EFB obtained from the fieldwork was thus used (Tenaga Sulpom
pers. comm., 27/09/2017).
For the biogas, the mass amount of CH4 available was identified considering that only CH4
content of the biogas was combustible (Yeoh, 2004). The suggested chemical oxidation
demand value (COD) of 51,000 mg/l and methane conversion factor of 0.25 kg CH 4 /kg COD
were used to identify the methane availability (Chin et al., 2013). The CV of CH4 is quoted in
a range of 50 MJ/kg - 55 MJ/kg (Chin et al., 2013). The minimum CV of CH4 was used
accordingly. The efficiency of AD process was assumed at 80% (Tong and Jaafar, 2004; Yacob
et al., 2006). It is understood that the amount of CH4 recovered is subjected to 0.1 methane
leakage factor (UNFCC, 2017).
The average thermal efficiency of the boiler derived from the previous technological
assessment studies of the biomass-based power plant can reach up to 80% to 90% (Gómez et
al., 2006; Thornley et al., 2009). In this study, a conservative 80% boiler efficiency was applied

60
to all systems. Note that, with this selection of parameters, the heat released from the
combustion process is considered as the maximum heat hence the steam that is generated.
By substituting Eq. (1) into Eq. (2), the mass flow rate of the HPSHS (𝑚𝐻𝑃𝑆𝐻𝑆 ) can be
expressed in the latter.
𝑄𝐶 = 𝑚𝐻𝑃𝑆𝐻𝑆 [𝐶𝑝 (𝑇𝑠𝑎𝑡 − 𝑇𝐵𝑊𝐹 ) + ∆ℎ𝑣𝑎𝑝 + (ℎ𝑠𝑢𝑝 − ℎ𝑣 )] (2)

where 𝐶𝑝 is specific heat capacity of water, 𝑇𝑠𝑎𝑡 is the saturated temperature of steam, 𝑇𝐵𝑊𝐹 is
the temperature of boiler feed water, ∆ℎ𝑣𝑎𝑝 is the standard enthalpy of vaporization of water,
ℎ𝑠𝑢𝑝 is the specific enthalpy of HPSHS and ℎ𝑣 is specific enthalpy of saturated steam. In this
research, the value extracted from the applied thermodynamic steam table as below were used
as follows (Wan et al, (2016)):
i. 𝐶𝑝 = 4.187 kJ/kg°C
ii. 𝑇𝑠𝑎𝑡 = 264 °C at 50 bar.
iii. 𝑇𝐵𝑊𝐹 = 105 °C
iv. ∆ℎ𝑣𝑎𝑝 = 1639.6 kJ/kg
v. ℎ𝑠𝑢𝑝 = 3433.7 kJ/kg at 50 bar.
vi. ℎ𝑣 = 2794.2 kJ/kg at 50 bar
Next, the HPSHS generated on the annual basis (𝐻𝑃𝑆𝐻𝑆𝐴𝐶𝐶 ) is determined from the heat
released from combustion by applying Eq. (3).
𝐻𝑃𝑆𝐻𝑆𝐴𝐶𝐶 = 𝑚𝑆 × 60 × 60 × 𝑂𝑝𝑒𝑟𝑎𝑡𝑖𝑜𝑛 𝐻𝑜𝑢𝑟 (3)
The operation hours of POMR-RES are fixed at 8000 hours (Sadhukhan, Ng and Hernandez,
2014; Abdul Malek et al., 2017). The installed electricity generation capacity (𝐸𝐶𝐴𝑃 ) in mega-
watt (MW) is estimated using Eq (4) and the electricity generation efficiency (𝑛𝑒𝑙𝑒𝑐 ) is fixed at
35% (Eurelectric, 2003; Rosen, Le and Dincer, 2005; Alves et al., 2015)
𝑂𝑝𝑒𝑟𝑎𝑡𝑖𝑜𝑛 𝐻𝑜𝑢𝑟𝑠
𝐸𝐶𝐴𝑃 = 𝐻𝑃𝑆𝐻𝑆𝐴𝐶𝐶 × ℎ𝑠𝑢𝑝 / ( ) × 𝑛𝑒𝑙𝑒𝑐 (4)
3600

The installed electricity generation capacity (𝐸𝐶𝐴𝑃 ) is also estimated by fixing the specific
steam consumption at 3 tonnes of steam per MW capacity (Summers, 1971; Ganapathy, 1996;
Bartnik, Buryn and Hnydiuk-Stefan, 2016).
Lastly Eq. (5) is used to estimate the potential amount of electricity generated (𝐸𝐺𝐸𝑁 ) in
megawatt-hour (MWh) from the installed capacity established above.
𝐸𝐺𝐸𝑁 = 𝐸𝐶𝐴𝑃 × 𝑂𝑝𝑒𝑟𝑎𝑡𝑖𝑜𝑛 𝐻𝑜𝑢𝑟𝑠 (5)

61
4.2.2 Computer-Aided Simulation Modelling

The ASPEN Plus ® process simulation software package is used to simulate the POMR-RES
CHP plant. The aim of the simulation process is to estimate the electricity generation potential
and heat extraction rate from the BPST (Magnusson, 2011; Hunpinyo et al., 2013). This
approach has been widely adopted in modelling the biomass CHP plant, allows flexibility to
modulate the operational setting of the plant according to the specific system requirement (Ng
and Sadhukhan, 2011; McIlveen-Wright et al., 2013). The process flow diagram for the CHP
plant was thus developed and simulated. The flow diagram comprises a RGibbs reactor acts as
the boiler for the generation of the HPSHS, isentropic turbine to drive the electricity and heat
generation and heat exchanger for heat transfer. The RGibbs reactor choice in ASPEN Plus ®
works on the principle of minimum Gibbs free energy change across the reaction system, to
estimate the product composition, based on the specifies product components. The chemical
compositions of EFB and biogas were considered as the input to the RGibbs reactor together
with sufficient air to promote full combustion condition. The chemical compositions of both
feedstocks and the amounts of air required are specified using spreadsheet yield models based
on the proximate and ultimate analyses of the feedstock (Sadhukhan et al., 2009).

There is limited chemical composition analysis for semi-dry EFB found in the literature.
Therefore, the proximate and ultimate analyses of dry EFB in Table 4.1 are used. 45% water
by mass are included to replicate the moisture content of the semi-dry EFB (Hussain et al.,
2006; Alias, Ibrahim and Hamid, 2014). The proximate and ultimate analyses of biogas are
shown in Table 4.2. The turbine is simulated under adiabatic condition receiving HPSHS (50
bar, 500°C) for simultaneous electricity and heat generation. The discharge pressure and
temperature for the BPST are set at 2.5 bar and 145°C respectively. An isentropic efficiency of
80% is considered for the turbine model to be able to estimate the power and heat generation
from the turbine. Note that the indicated horsepower of the turbine and the heat duty of the heat
exchanger were recorded as the electricity and heat generation potential of the CHP
respectively.

62
Table 4.1: Ultimate and Proximate Analysis of EFB (Hussain et al., 2006; Alias, Ibrahim and
Hamid, 2014)
Ultimate Analysis wt %
Carbon (C) 48.71 ± 5.08
Oxygen (O) 48.18 ± 1.26
Hydrogen (H) 7.86 ± 3.09
Nitrogen (N) 0.25 ± 0.11
Proximate Analysis wt %
Moisture Content 67.00 ± 1.41
Volatile Matter 87.04 ± 0.42
Ash 4.60 ± 0.50
CV 18.80 ± 0.74

Table 4.2: Ultimate and Proximate Analysis of Biogas (Yeoh, 2004; Yacob et al., 2006)
Ultimate Analysis wt %
Methane (CH4) 65
Carbon Dioxide (CO2) 35
Proximate Analysis wt %
Moisture Content 0
CV 50

4.2.3 Malaysia Palm Oil Mill Standard Energy Calculation Method

The electricity generation potential is also estimated using the standard calculation method
established quite recently by MPOB. This standard is formulated according to the Biomass
Fact Sheet for the Malaysian Palm for Malaysia Palm Oil Industry (Ng et al., 2011). Due to its
simplicity, this method has been widely used in several studies to reckon the utilization of the
POMR for electricity generation at national level (Shuit et al., 2009; Shafie et al., 2012; Chin
et al., 2013; Aghamohammadi et al., 2016; Md Jaye, Sadhukhan and Murphy, 2016). These
studies have successfully projected the high potential of generating electricity from EFB and
biogas. This method, however, does not define the type of energy conversion technology used,
the efficiency level implied as well as the probability to supply the heat to the mills. Therefore,
the calculation method in Table 4.3 and 4.4 for EFB and biogas is considered valid for
combustion technology with electricity generation potential quoted at the rate of 21%- 25%
from the extractable energy from biogas and EFB.

63
Table 4.3: Estimation of Electricity Generation from EFB (Ng et al., 2011)

No Parameters (unit) Calculation


1. Extractable Energy from EFB (MJ/year) Amount of available EFB × Calorific
Value of EFB
2. Estimated Output Energy (MJ/year)a 0.25 × Extractable Energy from EFB
3. Estimated Generated Electricity (MWh/year)b Estimated Output Energy / 3600
4. Electricity Generation Capacity (MW) Estimated Generated Electricity /
Plant Operation Hours
Notes:
a b
Estimated output energy at 25% thermal efficiency 1 MWh is equal to 3600 MJ

Table 4.4: Estimation of Electricity Generation from Biogas (Ng et al., 2011)

No Parameters (unit) Calculation


1. Extractable Energy from Biogas (MJ/year) Amount of available biogas × Calorific
Value of biogas
2. Estimated Output Energy (MJ/year)c 0.21 × Extractable Energy from Biogas
3. Estimated Generated Electricity (MWh/year)b Estimated Output Energy / 3600
4. Electricity Generation Capacity (MW) Estimated Generated Electricity /
Plant Operation Hours
Notes:
c
Estimated output energy at 21% thermal efficiency

4.3 Data Observation

This section discusses the data collected for this research from the fieldwork (i.e. visits to the mills and
semi-structured interviews) and literature review. The researcher visited three different palm oil mills
at various scales in Johore (Mill A), Melaka (Mill B) and Pahang (Mill C). The researcher also
interviewed seven authorities’ personnel from MPOB, MBIC, EPU, SEDA and TNB. As mentioned in
Section 3.3, there were difficulties gaining access to the mills, consequently to the mill’s production
data, due to the miller’s concern over the data sensitivity. The nature of the palm oil business itself
restricts the data transaction process as most of the mills and plantations are affiliated to the major
developers who give them a limited authority to disclose the information (Umar, Jennings and Urmee,
2013). Furthermore, from the interview conducted, it was found that the authorities’ personnel are not
prepared to share the information required as they are bounded with the data privacy restriction,
bureaucracy as well as data ownership concern. The data scarcity issues have been described as one of
the limitations and challenges to conduct the biomass-based energy generation research that required
further rectification (Yatim et al. 2016).

Due to this known reason, despite of exhaustive efforts put during the data collection process, the
available data for modelling the POMR-RES is still limited. Therefore, the researcher opts to use a

64
reasonably accurate mixture of data gathered from the fieldwork, review of literature and other credible
resources throughout. Most of the data used either from fieldwork or literature review have been verified
by the millers and qualified personnel from the relevant agencies.

This section discusses the data collected for this research from the fieldwork (i.e. visits to the
mills and semi-structure interviews) and literature review. A reasonably accurate mixture of
data gathered from the fieldwork, literature review and other credible resources is used
throughout. Most of the data used either from fieldwork and literature review have been
verified by the millers and qualified personnel from the relevant agencies. The researcher
visited three different palm oil mills at various scales in Johore (Mill A), Melaka (Mill B) and
Pahang (Mill C). There were difficulties gaining access to the mills, consequently to the mill’s
production data, due to the miller’s concern over the data sensitivity. The nature of the palm
oil business itself restricts the data transaction process as most of the mills and plantations are
affiliated to the major developers who give them a limited authority to disclose the information
(Umar, Jennings and Urmee, 2013).

The tabulated data obtained from the fieldwork are given in Table 4.5. The data were based on
the mills’ production reports for the year 2014 and 2015.

Table 4.5: Production Profile of Visited Mills

Data Types Mill A Mill B Mill C


Processing Capacity (t FFB/hr) 27 35 90
Daily Operating Hours (hrs/day) 16 14 20
Yearly Operating Hours (hrs/yr) 4992 3360 6240
Operating Day (day/yr) 312 240 312
FFB processed(t/yr) 134,800 115,700 561,600
EFB produced (t/yr) 29,700 27,800 123,500
POME produced (m3/yr) 96,200 69,400 393,120
Biogas produced (m3/yr) 47 N/A N/A
Calorific Value of EFB (MJ/kg) N/A 19 N/A
Calorific Value of biogas (MJ/kg) N/A 55 N/A

The data revealed significant inconsistencies in all the measured parameters. These
inconsistencies are due to a number of factors: 1. different operation procedures between the
mills; 2. limited deliveries of FFB from the estate; 3. seasonal FFB pruning; 4. unexpected
machine failures; 5. scheduled maintenance routine for the mill; 6. assumption that both
residues are ‘waste’ and so little attempt to maintain records of residues availability. As these
factors increased the likelihood of the mills not operating at the designated processing capacity,

65
it was difficult to quantify the actual availability of the residues (pers. comm., 13/07/2015 –
26/06/2016).

To establish an appropriate POMR-RES model, it was necessary in this study to estimate the
available EFB and POME. The information gathered from literature review and interviews
suggested that the availability of EFB and biogas was commonly estimated using standard
biomass to product ratio predicated by MPOB (Yusoff, 2006; Ng et al., 2011; Chin et al., 2013;
Loh, 2017). It is also assumed that the residues are readily available over the lifespan of the oil
palms which is typically 25 years (Yusoff, 2006). The complete procedures to estimate the
availability of EFB and biogas are presented in Table 4.6 and Table 4.7 respectively.

Table 4.6: Estimation of EFB Availability (Ng et al., 2011)

No Parameters (unit) Correlation


1. Estimated EFB available (tonne/year)a 23% × FFB processed
2. Estimated Dry EFB available (tonne/year)b 35% × Estimated EFB available
Notes:
a b
EFB is weight at 60% moisture content EFB is weight at 11% moisture content

Table 4.7: Estimation of Biogas Availability (Ng et al., 2011)

No Parameters (unit) Correlation


1. Estimated POME available (tonne/year)c 0.65 × FFB processed
2. Estimated Biogas Production (m3/year)d 28 m3 × Estimated POME available
Notes:
c
Each tonne FFB processed in the mill generated 0.65m3 POME
d
Each tonne of POME produced 28 m3 of biogas

However, the correlations for EFB gave higher values than the actual observations from the
fieldwork. The proportion of the EFB that was practically available was only about 15% to
16% instead of 22% to 23% shown in the literature (Loh, 2017; Wu et al., 2017). This is found
due to ongoing improvement in the mills’ operation efficiency and awareness of the millers to
minimize the residues produce from the milling process. Similarly, for POME, some mills
produced more than 0.8-tonne effluent from processing one tonne of FFB. There were no
formal measurements taken for biogas or methane so far (pers. comm., 13/07/2015 –
26/06/2016).

All of these inconsistencies resulted in the actual availability of the residues not correctly
specified leading to lack of consensus on how to model the POMR-RES, which made
replication of the model difficult and slowed the momentum to scale up the number of

66
installations (Abdullah, Sulaiman and Gerhauser, 2011; Teh, 2016). In this context, this chapter
aims to establish an appropriate estimation of available EFB and biogas in the mills according
to their fresh fruit bunch processing capacity. To date, far too little attention has been paid to
quantify the actual availability of the residues in the mills where most of the quantification
made so far are based on the aggregate values at the national level (Sulaiman et al., 2010;
Abdullah and Sulaiman, 2013; Lim and Biswas, 2015).

The 15% EFB fractions were used to estimate the availability of EFB. The availability of
POME was evaluated using the ratio of 0.65 m3 to one tonne of FFB processed. The operating
hours for the mill was fixed at 8000 hours annually. Figure 4.3 presents the range of potential
availability of EFB and biogas in the mill according to their FFB processing capacity and the
tabulated statistical data is available in Appendix B-2. Note that, in this chapter, the term
biogas is used to refer to only the methane content of the biogas produced from the AD of the
POME. From the graph in Figure 4.3, it can be seen that higher availability of EFB and biogas
observed in larger-scale mills in Peninsular Malaysia.

120
110
100
FFB Processing Capacity (t/hr)

90
80
70
60
50
40
30
20
10

0 200 400 600 800 1000 1200


k t/yr

FFB Processed EFB Ganerated Biogas Generated

Figure 4.3: POMs EFB and Biogas Production Profile

67
4.4 POMR-RES Power Plant

4.4.1 EFB-based POMR-RES Power Plant

The mathematical modelling of the EFB-based POMR-RES power plant is discussed in this
section. The aim of modelling the power plant is to estimate the electricity generation potential
from the semi-dry EFB available in the mills. The electricity generation potential and the total
amount of electricity generated from the available EFB annually are estimated using Eq.1 to
Eq.5. The modelling spreadsheets are attached in Appendix B-3, and the results are presented
in Table 4.8.

It is evident that as a resource-driven system, the amount of available EFB is the most
influential factor in the electricity generation. The linear relationship between the resource
availability and the electricity generation potential was observed. This result also indicates that
all palm oil mills in Peninsular Malaysia are capable of generating electricity from semi-dry
EFB. With an electrical efficiency of 35%, the range of the electricity generation potential is
found to be in between 0.9 MW and 10.7 MW. However this electricity generation potential is
lower than the estimation made by Abdul Malek et al. (2017). This capacity does not exceed
the maximum size (30.0 MW) to be eligible as the feed-in approval holder (FiAH). As one of
the FiAH, the mill is entitled to apply the renewable energy quota to generate electricity from
biomass offered by SEDA. Ahmad and Tahar (2014) point out that with proper selection of
energy conversion technology and better articulation of FiT mechanism, EFB would emerge
as a favourable alternative resources to generate electricity in Peninsular Malaysia.

68
Table 4.8: Mathematical Modelling of EFB-based Power Plant
Processing Capacity (t/hr) 10 20 30 40 50 60 70 80 90 100 110 120
Extractable Heat (MW) 2.20 4.40 6.60 8.70 10.90 13.00 15.30 17.50 19.60 21.80 24.00 26.20
HPSHS Generated (k t/yr) 21.40 42.70 64.00 85.30 107.00 128.00 149.30 170.70 192.00 213.40 234.70 256.00
Electricity Generation 0.90 1.80 2.70 3.60 4.40 5.30 6.20 7.10 8.00 8.90 9.80 10.70
Potential (MW)
Electricity Generated 7.10 14.20 21.40 28.50 35.60 42.70 49.90 57.00 64.10 71.20 78.40 85.50
(GWh/yr)

69
4.4.2 Biogas-based POMR-RES Power Plant

The electricity generation potential for biogas was calculated according to the mathematical
modelling procedures described in Section 4.3.1. The linear relationship between the mass
flowrate of the biogas and the electricity generation potential was asserted and shown in Table
4.9. In view of the results obtained, biogas has lower electricity generation capacity than EFB
over the whole range of the mill processing capacity. This is associated with the consideration
that only methane composition in the biogas is combustible (Sadhukhan, 2014). The highest
electricity generation capacity was estimated at 3.60 MW with 29.20 GWh of electricity
generated annually using available biogas from the mills with 120 t/hr processing capacity. The
findings of the current research are consistent with those by Mohd Ghazi and Muhammad Nasir
(2017) that also showed lower amount of electricity generated from biogas. These findings
further supported the research by Yoshizaki et al. (2013) and Chin et al. (2013), which found
the 54 t/hr, and 60 t/hr mills have the potential to produce 8.2 GWh and 13.26 GWh of
electricity per year. Elsewhere, Ahmad and Tahar (2014) have concluded that although the
electricity generation potential from biogas is less than EFB, there is an accumulation of 250
MW electricity generation potential from biogas to be explored in diverse areas around the
country.

70
Table 4.9: Mathematical Modelling of biogas-based Power Plant
Processing Capacity (t/hr) 10 20 30 40 50 60 70 80 90 100 110 120
Extractable Heat (MW) 0.80 1.50 2.20 3.00 3.70 4.50 5.20 6.00 6.70 7.50 8.20 9.00
HPSHS Generated (k t/yr) 7.30 14.60 21.90 29.20 36.50 43.80 51.10 58.40 65.70 73.00 80.20 87.50
Electricity Generation 0.30 0.60 0.90 1.20 1.50 1.80 2.10 2.40 2.70 3.00 3.30 3.60
Potential (MW)
Electricity Generated 2.40 4.90 7.30 9.70 12.20 14.60 17.00 19.50 21.90 24.40 26.80 29.20
(GWh/yr)

71
4.5 MPOB Standard Calculation Method

The results of the electricity generation potential from EFB and biogas using the standard
calculation method are shown in Table 4.10 and 4.11 respectively. Altogether, it can be seen
that the results have a similar trend as the mathematical modelling. The electricity generation
capacities from EFB obtained from this method are consistent with the results obtained from
the mathematical modelling. For instance, both methods estimated the electricity generation
capacity of 5.3 MW for 60 t/hr mills. Differently, for biogas, the electricity generation
capacities obtained from this method were 37% lower than the mathematical model. The
variation in capacity between both methods can be seen graphically in Figure 4.4.

12

10
Electricity Generation Potential (MW)

0
10 20 30 40 50 60 70 80 90 100 110 120
Mill Processing Capacity (t/hr)

EFB Math Modelling EFB MPOB Biogas Math Modelling Biogas MPOB

Figure 4.4: Variation between Mathematical and MPOB Modelling of POMR-RES Power
Plant

The difference in the estimated electricity generation capacity from biogas between the two
methods is due to lower thermal efficiency settings of 21% while estimating the output energy
from the combustion process. Combustion and electrical efficiency of 80% and 35%
respectively were applied in the mathematical modelling. The difference in parameters setting
used for methane conversion process is also believed to be the potential contributing factor to
this inconsistency although it has not been mentioned in Table 4.4.

72
Table 4.10: MPOB Modelling Method of EFB-based Power Plant
Processing Capacity (t/hr) 10 20 30 40 50 60 70 80 90 100 110 120
Extractable Heat (PJ/yr) 0.10 0.20 0.30 0.40 0.50 0.61 0.71 0.81 0.92 1.02 1.12 1.20
Output Energy (PJ/yr) 0.02 0.05 0.07 0.10 0.13 0.15 0.17 0.20 0.23 0.25 0.28 0.30
Electricity Generation 0.9 1.80 2.70 3.50 4.40 5.30 6.20 7.10 8.00 8.85 9.70 10.60
Potential (MW)
Electricity Generated 7.10 14.20 21.30 28.30 35.40 42.50 49.60 56.70 63.75 70.80 77.90 85.00
(GWh/yr)

Table 4.11: MPOB Modelling Method of biogas-based Power Plant


Processing Capacity (t/hr) 10 20 30 40 50 60 70 80 90 100 110 120
Extractable Heat (MW) 0.03 0.05 0.09 0.12 0.14 0.18 0.21 0.23 0.26 0.30 0.32 0.35
HPSHS Generated (k t/yr) 0.01 0.02 0.03 0.04 0.05 0.06 0.06 0.07 0.08 0.09 0.10 0.12
Electricity Generation 0.20 0.40 0.60 0.90 1.10 1.30 1.50 1.70 1.90 2.10 2.30 2.60
Potential (MW)
Electricity Generated 1.70 3.40 5.10 6.80 8.50 10.50 11.90 13.60 15.30 17.00 18.70 20.40
(GWh/yr)

73
4.6 POMR-RES CHP Plant

4.6.1 EFB-based POMR-RES CHP Plant

A notably lower electricity generation potential was obtained from the simulation of the CHP
plant when the pressure and temperature of the exit steam from the BPST was set to 2.5 bar,
145°C. The simulation results of the electricity generation potential and the mass of the heat
produced in the CHP plant are presented in Table 4.12. The significant reduction in the
electricity generation potential by ~65% was resulted compared to the electricity generation
potential from the POMR-RES power plant. The electricity generation potential varies from
the smallest scale of 0.4 to 5.3 MW with the heat generation potential for the CHP plant being
estimated as about 1.9 – 21.2 MW which is equivalent to 18,400 – 220,000 tonnes of medium
pressure steam. The electrical efficiency of the CHP was measured at 14%-15%, and the overall
plant efficiency reached 80%. These results agree with those of Badami and Mura (2009) for a
wood waste-based CHP plant. However, it is found to be higher than the estimation made by
Wu et al. (2017) as shorter plant operation hour was used in their study.

The lower electricity generation of the CHP plant reflects the second law of thermodynamics
for the steam cycle which describes that in an adiabatic condition, the magnitude of work
produced by the turbine is determined by the enthalpy difference between the inlet and outlet
ports (Borgnakke and Sonntag, 2013). A higher magnitude of work is produced for a greater
enthalpy difference in the steam turbine (Ali, Tesfamichael and Hassan, 2014; Gebreegziabher
et al., 2014; Wu et al., 2015). Simoes-moreira (2012) in his investigation of the thermodynamic
performance of power plants demonstrated that extracting saturated steam from the steam
turbine for the heating process has a strong influence on the enthalpy difference in the turbine.
Later, an in-depth study for the small-scale CHP system by Geete and Khandwawala (2013),
reached the same conclusion. Alterations in the enthalpy setting of the turbine inlet and outlet
ports also alter the output power and heat extraction rate.

74
Table 4.12: Computer-Aided Modelling of EFB-based CHP Plant
Processing Capacity (t/hr) 10 20 30 40 50 60 70 80 90 100 110 120
Electricity Generation 0.40 0.90 1.30 1.70 2.20 2.70 3.10 3.50 3.90 4.50 4.80 5.30
Potential (MW)
Electricity Generated 3.20 7.20 10.40 13.60 17.60 21.60 24.80 28.00 31.20 36.00 38.40 42.40
(GWh/yr)
Heat Generation Potential 1.90 3.60 5.40 7.20 9.00 11.00 12.50 14.50 16.20 18.20 19.80 21.70
(MW)
Heat Available (k t/yr) 18.40 36.80 54.40 72.80 91.20 111.20 125.60 146.40 164.00 184.00 200.00 220.00

Table 4.13: Computer-Aided Modelling of biogas-based CHP Plant


Processing Capacity (t/hr) 10 20 30 40 50 60 70 80 90 100 110 120
Electricity Generation 0.04 0.08 0.14 0.18 0.23 0.28 0..33 0.37 0.41 0.46 0.50 0.55
Potential (MW)
Electricity Generated 0.32 0.64 1.10 1.40 1.84 2.24 2.64 2.96 3.28 3.68 4.00 4.40
(GWh/yr)
Heat Generation Potential 0.20 0.40 0.60 0.75 0.95 1.10 1.34 1.50 1.70 1.90 2.10 2.25
(MW)
Heat Available (k t/yr) 1.92 4.00 5.80 7.60 9.60 11.40 13.60 15.20 17.20 19.20 20.80 22.80

75
4.6.2 Biogas-based POMR-RES CHP Plant

The simulation result of the biogas-based CHP plant can be seen in Table 4.13. The tabulation
shows that the highest electricity generation potentials in the CHP plants is 0.5 MW. Similar
to the earlier discussions in Section 4.7.1, setting the exiting steam from the turbine at higher
pressure and temperature lowered the electricity generation potential from the CHP plant. With
the heat generation potential of 0.2 to 2.25 MW, the CHP plant supplied about 1,920 tonnes to
22,800 tonnes of medium pressure steam to the mills.

The lower electricity and heat generation potential obtained shows the unsuitability of biogas
as the feedstock for the CHP plant as described by Hosseini and Wahid (2014a). The flame
produced from combusting the biogas is found to be unstable and insufficient to increase the
temperature of the steam. Hosseini and Wahid (2014a) highlighted the need to install the
flameless combustor, despite being costly to produce more uniform temperature and pressure
steam. In other study, Qian et al. (2017) raised the concern on the low combustible properties
of biogas due to the high CO2 content. The competency of biogas for combustion process was
studied extensively by Yingjian et al. (2014) which conclusively stated the incompatibility of
biogas to be used in CHP plant. They suggested to use the biogas for single generation of either
electricity or heat.

4.7 Biomass Dryer

Drying the EFB is one of the suggested pre-treatment measures to enhance the performance of
the combustion process. Drying the EFB reduces the moisture content and increases its energy
density and subsequently the overall heat recovery efficiency. These are achievable through
the elevation of the CV of EFB and effectiveness of the combustion process (Yarnal and
Puranik, 2010; Luk et al., 2013). For instance, by reducing the moisture content of EFB from
60% to 11%, the CV is expected to increase from 8.5 MJ/kg to 18.8 MJ/kg and the efficiency
of the combustion process is expected to reach to 90%. Drying the biomass also has been
appraised to have distinct environmental benefit by reducing the emission from the incomplete
combustion (Jenkins et al., 1998; Kampa and Castanas, 2008).

An additional installation of the hot air dryer is required to perform the drying process (Song
et al., 2012; Gebreegziabher et al., 2014). Drying the EFB is however considered as energy-
and cost-intensive process (Jirjis, 1995; Xu and Pang, 2008; Pang and Mujumdar, 2010).
Utilizing the waste heat such as flue gas and hot water, and extracting the superheated steam

76
from the existing industrial process as the drying medium could substantially reduce the energy
expenditure of the drying process (Fagernäs et al., 2010; Li et al., 2012). Besides the concerns
over the energy and cost intensiveness, space requirement is another consideration for dryer
installation. Approximately 15 m2 of space are needed for every 100 kW of heating requirement
(Treco Green Heat n.d.).

The rotary drum dryers, conveyors dryers, and cascade dryers are amongst the common types
of dryer used for biomass. These types of dryer have been proven to have a superior drying
effect in addition to easy installation, operation and maintenance requirements (Amos, 1998;
Iguaz et al., 2003). Vigants et al. (2015) in their study classified the dryer according to its
drying medium. Ståhl et al. (2004) suggested that rotary drum dryers using the heat from the
combustion process is far more cost-effective and technically sound, therefore better adapted
for drying the biomass.

Three main approaches can be adapted to model the biomass dryer in the literature: dryer
enthalpy balance and kinematics (Gebreegziabher, Oyedun and Hui, 2013); pinch analysis
(Gebreegziabher, Olajire and Yu, 2014); Aspen Plus simulation (Luk et al., 2013; Aziz, Oda
and Kashiwagi, 2015). The main aim of modelling the dryer is to analyse the energy
requirement of the drying process, to scale the size of the dryer and to evaluate the
improvements in the overall performance of the energy plant integrated with the dryer process.
Recently, a large volume of published literature has promoted the advantages of biomass dryer
from the technical and environmental perspective nonetheless most of them have omitted the
economic and spatial provision concerning the dryer installations. This section investigates the
technical feasibility of hot air dryer in lowering the moisture content of the EFB by estimating
the heat requirement and electricity generation after dryer integration.

4.7.1 Biomass Dryer Technical Modelling

The modelling of a rotary dryer for EFB was undertaken to study its enthalpy balance and
kinematics (Gebreegziabher, Oyedun and Hui, 2013). The total heat required to dry the biomass
in MW (𝑄𝑑𝑟𝑦𝑒𝑟 ) is the summation of the sensible heat(𝑄𝑆 ), and latent heat (𝑄𝐿 ) as shown in
Eq. (6):

𝑄𝑑𝑟𝑦𝑒𝑟 = 𝑄𝑆 + 𝑄𝐿 (6)

Where the 𝑄𝑆 and 𝑄𝐿 both measured in MW are deduced from Eq. (7) and Eq. (8), respectively.

77
𝑄𝑆 = [𝑀𝑤,𝐸𝐹𝐵 × 𝐶𝑃,𝑤 × (𝑇𝑓 − 𝑇𝑖 ) + (𝑚𝑑𝑎𝑓 + 𝑚𝑎𝑠ℎ ) × 𝐶𝑝,𝐸𝐹𝐵 × (𝑇𝑓 − 𝑇𝑖 )]/1000 (7)

𝑄𝐿 = [𝑀𝑤,𝑒𝑣𝑝 × 𝐶𝑃,𝑤 × ((𝑇𝑓 + 10℃) − 𝑇𝑖 ) + ∆𝐻𝑣𝑎𝑝 ] /1000 (8)

Here, 𝑀𝑤,𝐸𝐹𝐵 is the mass of water remaining in EFB, 𝐶𝑃,𝑤 is the specific heat of water, 𝑇𝑖 is
the inlet temperature of solid stream of a dryer, 𝑇𝑓 and is the outlet temperature of solid stream
of a dryer, 𝑚𝑑𝑎𝑓 is the mass flow rate of fuel in dry and ash free basis, 𝑚𝑎𝑠ℎ is mass flow rate
of ash, 𝐶𝑝,𝐸𝐹𝐵 is specific heat capacity of EFB, 𝑀𝑤,𝑒𝑣𝑝 is the water in the EFB evaporation rate
and ∆𝐻𝑣𝑎𝑝 is the latent heat of vaporization. The 𝑀𝑤,𝐸𝐹𝐵 and 𝑀𝑤,𝑒𝑣𝑝 are obtained using Eq. (9)
and (10).

(𝑚𝑑𝑎𝑓 + 𝑚𝑎𝑠ℎ ) × 𝑀𝑐𝑓


𝑀𝑤,𝐸𝐹𝐵 = (9)
(1 − 𝑀𝑐𝑓 )

(𝑚𝑑𝑎𝑓 + 𝑚𝑎𝑠ℎ ) × 𝑀𝑐𝑖


𝑀𝑤,𝑒𝑣𝑝 = (10)
(1 − 𝑀𝑐𝑖 + 𝑀𝑤,𝐸𝐹𝐵 )

It is assume that the moisture content of EFB is reduced from the initial moister content (𝑀𝑐𝑖 )
of 60% to final moisture content (𝑀𝑐𝑓 ) of 11% through the drying process. It is then further
assumed that the mass flowrate of EFB (kg/s) on dried and ash free basis (𝑚𝑑𝑎𝑓 ) and mass
flow rate of ash (kg/s) are obtained using Eq. (11) and Eq. (12).

𝑚𝑑𝑎𝑓 = (35% × 𝑅𝑎𝑤 𝐸𝐹𝐵) / 𝑂𝑝𝑒𝑟𝑎𝑡𝑖𝑜𝑛 𝐻𝑜𝑢𝑟𝑠 /3600 (11)

𝑚𝑎𝑠ℎ = (5% 𝑤. 𝑡 𝑜𝑓 𝑅𝑎𝑤 𝐸𝐹𝐵) / 𝑂𝑝𝑒𝑟𝑎𝑡𝑖𝑜𝑛 𝐻𝑜𝑢𝑟𝑠 /3600 (12)

In this case, the EFB was collected after the stripping process, and no mechanical drying was
required. The other modelling parameters from Luk et al. (2013) and Gebreegziabher, Oyedun
and Hui (2013) were used to complete the dryer modelling as follows:

1. Specific heat capacity of water (𝐶𝑃,𝑤 ) = 4.186 kJ/kg


2. Inlet temperature of solid stream of a dryer (𝑇𝑖 ) = 298 K
3. Outlet temperature of solid stream of a dryer (𝑇𝑓 ) = 373 K
4. Specific heat capacity of EFB (𝐶𝑃,𝐸𝐹𝐵 ) = 1.40 kJ/kg
5. Latent heat of vaporization (∆𝐻𝑣𝑎𝑝 ) = 2357.62 kJ/kg

A spreadsheet for the hot air dryer enthalpy balance and kinematics model as in Appendix B-
4 was developed to estimate the energy required to dry the available EFB from each mill. The

78
EFB-based POMR-RES was re-modelled using Eq. (1) to Eq. (5) to observe the technical
dimension variations from using semi-dry to dried EFB for electricity generation (see
Appendix B-5). The maximum CV value of EFB (18.80 MJ/kg) and combustion efficiency
(90%) were used to reckon the dried EFB. The heat requirement of the drying process was
added to the parasitic load of the electricity generation process. Similarly, the process flow
diagram for the CHP plant was also developed and simulated.

4.7.2 Dryers heat requirement

Table 4.14 presents the estimated heat requirement for drying the EFB to reduce its moisture
content from 60% to 11%. The tabulated results indicate that the heat required to dry the EFB
is ranging between 0.6 – 5.30 MW. These estimations are found to be different from the dryer
heat requirement reported by Aziz, Oda and Kashiwagi (2015) and Xu and Pang (2008). This
could be due to the different biomass used in their studies. However, the heat requirement to
dry EFB estimated in this study correlate favourably with Luk et al. (2013) who experimentally
assessed the heat requirement for drying the EFB. Due to limited study on the analysis of the
heat requirement for drying the EFB, the estimations in this thesis offer a compelling case to
extend the knowledge on the possibility to integrate the dryer to the EFB-based electricity
generation plant.

Table 4.14: Dryers Heat Requirement

Mill Processing Capacity Amount of EFB to be Dried Dryer Heat Requirement


(t/hr) (t/yr) (MW)
10 12,000 0.60
20 24,000 1.30
30 36,000 1.80
40 48,000 2.30
50 60,000 2.80
60 72,000 3.20
70 84,000 3.70
80 96,000 4.00
90 108,000 4.40
100 120,000 4.70
110 132,000 5.00
120 144,000 5.30

79
4.7.3 Dried EFB-based POMR-RES Power Plant

The mathematical modelling of the dried EFB power plant is presented in this section with an
aim to evaluate the electricity generation potential from the dried EFB. Similar modelling
approach as in Section 4.3.1 is employed using the CV of dried EFB (18.8 MJ/kg) and a boiler
efficiency of 90% to combust the dried EFB (Loh, 2017). In this section, the EFB is considered
to be undergoing a thorough drying process to reduce its moisture content to 11%. The mass
of the dried EFB is 35% of the mass of the raw EFB. It is assumed that the drying process of
EFB is carried out in two scenarios where the heat requirement is attained from (i) the heat of
the combustion process; and (ii) flue gas of the combustion process.

Table 4.15 presents the outcomes from modelling the dried EFB power plant when the heat
from the combustion process is used to dry the EFB. The electricity generation potential from
dried EFB varies from 0.80 MW to 10.10 MW. Note that this electricity generation potential
is found to be lower than the electricity generation potential of the semi-dried EFB in Table
4.8. Next, Table 4.16 provides the modelling outcomes from the power plant when the flue gas
of the combustion process is used to dry the EFB. In this case, higher electricity generation
potential is anticipated which varies from 1.00 MW to 12.10 MW.

From these results, the installation of the biomass dryer appears to be sensible in two situations:
1. the efficiency of the boiler without the dryer is lower than 70%; 2. the provision of auxiliary
heat resources to dry the EFB. From the mathematical modelling results, semi-dry EFB is in a
good position to be used as the feedstock without having to undergo the drying process (Song
et al., 2012; Gebreegziabher, Oyedun and Hui, 2013). Although drying the EFB increases its
energy content, Luk et al. (2013) argued the energy intensiveness of the drying process. Li et
al. (2012) found the parasitic load of the plant with a dryer to be higher than the energy
generated, making it technically infeasible. In theory, allowing EFB to undergo intentional
drying process can increase the overall system efficiency by 10% (Gebreegziabher, Oyedun
and Hui, 2013) while the installation cost of the industrial dryer leads to higher project cost
(Haque and Somerville, 2013). Due to this consideration, installation of a dryer to the biomass-
based electricity generation required further technical exploration and justification (Jorgenson,
Gilman and Dobos, 2011). However, some of the operational advantages of drying the biomass
mainly EFB can’t be assessed by energy balance and kinematic models. Drying the biomass
minimizes the boiler ignition problems such as higher ignition temperature requirement
(Sultana, Kumar and Harfield, 2010; Alves et al., 2015). Further to that, the possibility of

80
incomplete combustion is ultimately removed, lessening the effect of slagging and fouling,
formation of ash and corrosion and erosion problems in the boiler (Demirbas, 2005; Khan et
al., 2009; Saidur et al., 2011).

81
Table 4.15: Dried EFB Power Plant Mathematical Modelling with Steam Extraction for Drying Process

Processing Capacity (t/hr) 10 20 30 40 50 60 70 80 90 100 110 120


Available Heat (MW) 1.90 3.80 5.80 7.70 9.80 11.90 14.00 16.00 18.20 20.40 22.50 24.71
HPSHS Generated (k t/yr) 18.30 37.50 56.80 76.00 96.20 116.40 136.60 156.90 176.00 199.30 220.00 241.70
Electricity Generation 0.80 1.60 2.40 3.20 4.00 4.90 5.70 6.50 7.40 8.30 9.20 10.10
Potential (MW)
Electricity Generated 6.10 12.50 18.90 25.40 32.10 38.90 49.90 52.40 59.40 66.50 73.60 80.70
(GWh/yr)

Table 4.16: Dried EFB Power Plant Mathematical Modelling without Steam Extraction for Drying Process

Processing Capacity (t/hr) 10 20 30 40 50 60 70 80 90 100 110 120


Available Heat (MW) 2.50 4.90 7.40 9.90 12.30 14.80 17.30 19.70 22.20 24.70 27.10 29.60
HPSHS Generated (k t/yr) 24.10 48.30 72.40 96.50 120.70 144.80 168.90 193.10 217.20 241.30 265.40 289.60
Electricity Generation 1.00 2.00 3.00 4.00 5.00 6.00 7.00 8.00 9.00 10.10 11.10 12.10
Potential (MW)
Electricity Generated 8.10 16.10 24.20 32.20 40.30 48.30 56.40 64.40 72.50 80.60 88.60 96.70
(GWh/yr)

82
4.8 Conclusion

In conclusion, this chapter has scientifically estimated the annual availability of semi-dry EFB
and biogas in the mill depending on their FFB processing capacity. From these annual
availability estimations, the full electricity generation profiles of the POMR-RES power plant
and CHP plant were established using three modelling methods. Three sets of data were used
in the modelling: from fieldwork visits to three different mills; from semi-structured interviews
with relevant agencies; and from the literature review. However, these data are subjected to a
number of limitations. The data obtained from the fieldwork visits were found to be
inconsistent and incomplete. Furthermore, the data collected from literature review fails to take
few practical considerations into account. They were reported as an aggregated value at the
national level, rather than site-specific value. Due to these limitations, there is no detail report
on the actual availability of the EFB and biogas has been produced. Consequently, no
consensus for the suggested method for modelling the POMR-RES. The standard biomass to
product ratio approach coupled with MPOB standard calculation method was found to be the
common method used to estimate the available residues and for modelling the POMR-RES so
far.

From the analytical residues availability estimation conducted in this study, it was revealed that
the available EFB and biogas to be used for electricity generation were subjected to the multiple
technical constraints; therefore, it is much lower than what has been estimated theoretically. It
was revealed that the availability of EFB to be used for electricity generation (semi-dry EFB)
is only about 50% of its theoretical existence. Figure 4.5 illustrates the representation of the
availability of EFB based on the 50 t/hr FFB processing capacity mill that operates for 8000
hours.

83
Tonnes EFB
Theoretical Availability
23 % of FFB Processed
92,000 tonnes of wet EFB
(60% moisture content)

Technical Availability
15 % of FFB Processed
60,000 tonnes of wet EFB
(60% moisture content)

Actual Availability
77% of Technical
Availability
46,200 tonnes of semi-dry
EFB
(45% moisture content)

Actual Availability
35% of Technical
Availability
Year 21,000 tonnes of dried EFB
(11% moisture content)

Figure 4.5: Actual EFB Availability

Similarly, although a massive amount of available POME was reported, the amount of biogas
generated from the anaerobic digestion process is subjected to the conversion rate of the
organic content in POME. Further to that, only methane content of biogas was considered as
combustible; therefore, the combustion rate from biogas is lower than what was expected in
theory.

Based on a realistic estimation of available residues, the mathematical modelling method was
used to model the POMR-RES power plant and the combination of mathematical and
computer-aided process simulation for the CHP plant. The semi-dry EFB-based power plant
has the electricity generation potential ranging from 0.90 MW - 10.70 MW while the biogas-
based power plant from 0.30 MW - 4.10 MW. Differently, co-generating the heat has
compromised the potential to generate electricity in CHP plant. The electricity generation
potential from EFB has reduced to 0.40 MW- 5.30 MW. The reduction trend was also observed
for biogas. The electricity generation potential for biogas-based CHP plant was estimated at
0.04 W - 0.55 MW. Approximately 18,400 – 220,000 tonnes of medium pressure steam was
available in the EFB-based CHP plant while about 1,920 – 22,800 tonnes was available in
biogas-based CHP plant to be used for cooking the FFB.

84
The dried EFB-based power plant was modelled to evaluate the electricity generation potential
from the dried EFB. Additional about 0.60 MW to 5.30 MW heat energy was required for
drying the EFB. This heat energy demand was fulfilled either with the heat extracted from the
combustion process or the flue gas. It was demonstrated that the utilization of flue gas in the
drying process had increased the electricity generation potential of EFB. The electricity and
heat generation potentials established in this chapter will be used to assess the feasibility of the
POMR-RES in Chapter 5.

85
CHAPTER FIVE

TECHNICAL AND TECHNO-ECONOMIC FEASIBILITY ASSESSMENT OF


POMR-RES

This chapter provides detailed explanations of procedures to conduct the technical and techno-
economic feasibility assessment for POMR-RES. This chapter also evaluates the suitability of
semi-dry EFB and biogas as the feedstock for electricity generation as well as applicability of
on-site, off-site and co-operative installation of the POMR-RES. The latter part of this chapter
evaluates the technical and techno-economic feasibility of the dried-EFB based POMR-RES
and recommends the economically feasible size of POMR-RES.

5.1 Introduction

The potential of generating electricity from POMR has been widely discussed in various energy
statistic reports and academic publications. The EFB and biogas of POME were proven as the
residues that could have attained the highest electricity generation potential of 521 MW and
320 MW respectively (Ali, Daut and Taib, 2012). However, there are only limited examples of
actual installation of POMR-RES. This implies that the POMR-RES had failed to progress into
implementation phase thus was unable to meet the initial electricity generation expectations
(Ali, Daut and Taib, 2012; Khor and Lalchand, 2014). According to Yatim et al. (2017),
POMR-RES has not yet achieved commercial maturity and remains at the proof of concept
stage (TRL 3). The system still suffers from competitive pressure from conventional electricit y
generation sectors whereby it remains as an unprofitable system due to the high cost-intensive
nature of the technology. It is known as the highly fragmented market, and only a few examples
of successful business cases exist (Petinrin and Shaaban, 2015; Shirley and Kammen, 2015;
Yatim et al., 2017).
One of the reasons for failure in implementation of the POMR-RES is the absence of analytical
justification for its technical and techno-economic feasibility. Importance of justifying the
technical and techno-economic feasibilities of RE system were highlighted by Yilmaz and
Selim (2013). It was also recognized in the systematic conclusion done by Tomberlin and
Mosey (2013). Besides that, the classical studies conducted by Bridgwater (1995) and Fiala,
Pellizi and Riva (1997) stated the needs for a feasibility study before the installation of RE
system. These studies highlighted the three benefits of conducting the feasibility study as
follows: it provides a reliable basis to justify the incompatibility and impracticality of RE

86
system installations; it assesses the effectiveness of RE incentives being offered; and it serves
as a guide for the relevant parties to carefully plan an appropriate layout to achieve the RE goal.

Currently, there is no standard feasibility assessment method that can guide the installation of
a POMR-RES. In fact, selection of the feasibility indicators used in the assessment is still
puzzling (Petinrin and Shaaban, 2015; Alam et al., 2016). Although there are some research
studies on the feasibility of the POMR-RES, the data collected are found to be rather descriptive
and contradicting (Chin et al., 2013; Abdul Malek et al., 2017). This chapter seeks to overcome
this problem by offering a comprehensive framework to assess the technical and techno-
economic feasibility of EFB- and biogas- based POMR-RES. The framework will also focus
on suggesting the suitable residues to be used as the feedstock to generate electricity. It will
also recommend the appropriate installation scheme for different situations. This will give an
initial glimpse of the POMR-RES performance and determine the likelihood of success.

Effects due to the absence of a standard method for a feasibility assessment of POMR-RES
were gathered from the literature and field observations. Due to this lack of standard methods,
most of the POMR-RES installations to date were made based on a rough estimation derived
from general assumptions that lack comprehensive technical, techno-economic and
environmental feasibility justification (Umar, Jennings and Urmee, 2012; Loh and Abdul
Majid, 2016). As a result, four out of six existing POMR-RES installations were found to be
overly designed with poor techno-economic performance (Nasrin et al., 2011). This
unappealing technical and techno-economic performance had negatively affected the industry
(Verbruggen et al., 2010; Ahmad, Ab Kadir and Shafie, 2011; Maulud and Saidi, 2012; Umar,
Jennings and Urmee, 2014b). However, Umar, Jennings and Urmee (2014b) had revealed that
75% of the 85 respondents from the palm oil millers in Malaysia had expressed their interest
in collaborating with the government to contribute towards achieving the national renewable
energy targets. They were assuming that the technical and techno-economic performance of
the POMR-RES would be reasonable. Similar responses were received by the researcher while
interviewing the selected palm oil millers who also pointed out the needs for a technology
adoption guideline to patrol the actual implementation process. These findings confirm the
interests of the palm oil millers towards the utilization of POMR for electricity generation and
in the importance of feasibility justification. The findings also indicate the need for a custom
technology adoption guideline for POMR-RES.
In a recent study, Wan et al. (2016) had developed a comprehensive fuzzy-based optimization
framework to evaluate the feasibility of biomass-based heat and power system. Till date, this

87
framework has only been applied using the residues from the sago industries. Hence, there is a
need for a standard framework that can be used to assess the feasibility of POMR-RES.

5.2 Feasibility Study for POMR-RES

This section provides an overview of published work on the feasibility of POMR-RES in


Malaysia. It is a complete explanation from the early stages to the most recent feasibility study
conducted. The review includes area covered by the assessment, the method used to conduct
the assessment and the feasibility indicators selected, as summarised in Table 5.1. From the
summary, it is clear that the first systematic feasibility assessment was conducted after three
years of integrating RE into the national energy portfolio in 2000. From the tabulated summary,
it shows that most of the feasibility studies were conducted using the aggregated large-scale
data with very few attempts made to justify the feasibility of limited site-specific data. The
summary also revealed that the feasibility assessment conducted is still evolving from the
technical and resource perspectives. From thirty-five studies, there are only eight studies that
genuinely assessed the techno-economic feasibility of the installation using a few different
indicators. The environmental aspect of the electricity generation was evaluated in ten different
studies. These studies mainly provided the indications on global warming risk potential, and
the carbon emission resulted from this electricity generation. Recently, researchers are starting
to show their interest in finding more information related to the effectiveness of the relevant
policies.

Although a few extensive studies had been carried out on assessing the feasibility of electricity
generation from the mixture of palm oil residues, no single study had explicitly covered EFB
and biogas. The separate studies done by Yeoh (2004), Michael Griffin et al. (2014), Loh et al.
(2017b) and Abdul Malek et al. (2017) laid an important foundation to structure a standard
method to assess the feasibility of electricity generation from various perspectives. However,
there has been little attempt to combine the advantages of these studies and produce a more
systematic feasibility assessment method for the bioenergy system. Therefore, this study offers
a comprehensive framework to assess the technical and techno-economic feasibility of the
POMR-RES in Peninsular Malaysia. The framework was developed based on the combination
of traditional as well as sophisticated waste-to-energy feasibility assessment methods
deliberated from the previous studies (Sadhukhan et al., 2009; Wang et al., 2009; Patel et al.,
2011; Wan et al., 2016; Wu et al., 2017).

88
Table 5.1: Feasibility Assessment of POMR-RES in Malaysia
Author (s) Assessment Coverage Assessment Method Feasibility Indicator
Residues Technical/ Economic Environment Policy National-Level
Resources (NL)/
Site-Specific
(SS)
Koh and Hoi Mixed √ NL Resource to Energy Resource Availability
(2003) Conversion (REC) (RA), Energy
Generation Potential
(EGP)
Yeoh (2004) POME √ √ NL REC, Cost Benefit Analysis EGP, Return on
(CBA) Investment (ROI) ,
payback period (PP)
Yusoff (2006) Mixed √ √ NL REC, Policy Matching (PM) RA, EGP, Policy
Effectiveness (PE)
Sumathi, Chai and Mixed √ NL REC RA, EGP
Mohamed (2008)
Shuit et al. (2009) Mixed √ NL REC RA, EGP
Poh and Chong POME √ √ REC, Life Cycle Assessment EGP, Global Warming
(2009) (LCA) Potential (GWP)
Nasrin et al. (2011) Mixed √ NL
Chiew et al. (2011) Mixed √ SS REC EGP
Stichnothe and EFB,POME √ √ NL REC, LCA EGP, GWP
Schuchardt (2011)
Ng et al. (2011) EFB, POMR √ NL REC RA,EGP
Ong, Mahlia and Mixed √ NL REC RA,EGP
Masjuki (2011)
Hashim and Ho Mixed √ √ NL REC, PM RA, EGP, PE
(2011)

89
Author (s) Assessment Coverage Assessment Method Feasibility Indicator
Residues Technical/ Economic Environment Policy National-Level
Resources (NL)/
Site-Specific
(SS)
Yoshizaki et al. POME √ √ NL,SS REC, Techno-economic RA, EGP, ROI
(2012) Assessment (TEA)
Mahlia et al. Mixed √ √ NL REC, Techno-economic RA, EGP, Levelized
(2012) Assessment (TEA) Cost of Electricity
(LCOE)
Shafie et al. (2012) Mixed √ √ NL REC, PM RA, EGP,PE
Chin et al. (2013) POME √ √ REC, Techno-economic RA, EGP, Fuel Cost
Assessment (TEA) Saving (FCS)
Madaki and Seng POME √ √ NL REC, Life Cycle Assessment EGP, Global Warming
(2013) (LCA) Potential (GWP)
Lam et al. (2013) Mixed √ √ NL REC, PM RA, EGP, PE
Chiew and Mixed √ √ NL REC, LCA EGP, Global Warming
Shimada (2013) Potential (GWP)
Hasanuzzaman et EFB √ NL REC RA,EGP
al.(2014)
Abdul Rahim Mixed √ NL REC, LCA GWP, Carbon
(2014) Emission Saving
(CES)
Michael Griffin et Mixed √ √ √ NL REC, LCA, Linear Cost RA,EGP,FCS,CES,
al. (2014) Optimization Model Generation Cost
(LCOM) Minimization
Khor and Lalchand √ NL PM PE, Future Pathway
(2014)
Bazmi et al. (2015) Mixed √ SS REC RA,EGP
Petinrin and Mixed √ NL PM PE
Shaaban (2015)

90
Author (s) Assessment Coverage Assessment Method Feasibility Indicator
Residues Technical/ Economic Environment Policy National-Level
Resources (NL)/
Site-Specific
(SS)
Basri, Ramli and EFB, POME √ √ √ SS REC, CBA, Social RA, EGP, Social
Aliyu (2015) Perception Analysis Perception Indicators
Md Jaye, EFB, POME √ √ SS REC, LCA RA, EGP, GWP
Sadhukhan,
Murphy (2016)
Aghamohammadi Mixed √ NL,SS REC, Spatial Ananlysis RA, EGP,
et al. (2016) Infrastructures
Availability
Yatim et al. (2016) Mixed √ NL PM PE, Future Pathway
Loh (2017) Mixed √ NL REC RA, EGP
Loh et al. (2017b) POME √ √ √ √ NL REC, TEA, LCA, PM RA, EGP, GWP,
CES,PE
Abdul Malek et al. Mixed √ √ √ SS REC, TEA, LCA RA, EGP, GWP, CES
(2017)
Wu et al. (2017) Mixed √ SS REC RA, EGP
Oh et al. (2017) Mixed √ NL PM PE, Future Pathway
Yatim et al. (2017) Mixed √ NL Risk Assessment Risk Management
Indicator

91
5.3 POMR-RES Feasibility Assessment Framework

The proposed feasibility assessment framework for EFB based- and biogas based- POMR-RES
is given in Figure 5.1. The technical and techno-economic performance of POMR-RES was
assessed based on the electricity and heat generation potential that was established in Chapter
4. The detailed descriptions of each assessment are presented in subsequent sections. Three
case studies, namely OnS case study (on-site installation), OffS case study (off-site installation)
and Co-op case study (cooperative installation) were used to illustrate the outcome of the
framework. The technical and techno-economic justification of each case study is provided,
the suitable residue for electricity generation is suggested, and the appropriate installation
scheme for the systems is recommended. Finally, the economically feasible size of POMR-
RES based on ROI and net present value (NPV) analysis is presented in this chapter.

Palm Oil Mill Residues

Chapter
EFB Biogas 4

POMR-RES Power Plant POMR-RES CHP Plant

OnS OffS Co-op

Technical Techno-Economic

Sensitivity Analysis

Economically Feasible Size POMR-RES

Figure 5.1: POMR-RES Feasibility Assessment Framework

92
5.3.1 Technical Feasibility Assessment

The technical feasibility of the EFB- and biogas-based POMR-RES was measured differently
from most of the previous studies (Chau et al., 2009; Tan et al., 2014; Eleftheriadis and
Anagnostopoulou, 2015). In this thesis, both of the systems were considered as technically
feasible when the electricity generated is sufficient to accommodate the parasitic load of the
electricity generation process itself, fulfil the electricity or/and heat demand from the mills with
the surplus electricity to be fed to the grid. Mathematically, the POMR-RES was considered as
technically feasible when the Eq. (13) produces a null answer.

𝐸𝐺𝐸𝑁 − 𝐿𝑜𝑎𝑑𝑃 − 𝑀𝑖𝑙𝑙𝐷 − 𝐸𝑔𝑟𝑖𝑑 = 0 (13)

Here, 𝐸𝐺𝐸𝑁 is the amount of electricity generated from modelling of the system, 𝐿𝑜𝑎𝑑𝑃 is
parasitic load of the system, 𝑀𝑖𝑙𝑙𝐷 is electricity demand of the mill, and 𝐸𝑔𝑟𝑖𝑑 is surplus
electricity to be connected to the grid. Searcy et al. (2007) stated that the parasitic load of
biomass-based electricity generation is at a fraction of 8.5% of the total electricity generated
while the Electric Power Research Institute quoted 11% of the electricity generated was used
to facilitate the electricity generation, transmission, and distribution. The higher fraction value
was used in this chapter to represent the parasitic load of POMR-RES. Data from several
sources identified the electricity and heat demand to operate the mill and to support the wet-
milling process of CPO. The electricity needed to process one tonne of fresh fruit bunch (FFB)
is 24 kWh – 27 kWh (Chiew, Iwata and Shimada, 2011; Nasution, Herawan and Rivani, 2014).
It stated that 0.65 tonnes to one tonne of steam is required to cook the similar amount of FFB
(Chin et al., 2013; Tenaga Sulpom pers. comm., 27/09/2017). A mathematical spread sheet was
used to assess the technical feasibility of POMR-RES with mill’s electricity and heat demand
kept at 27 kWh, and 0.65 tonnes for each tonne of FFB processed.

5.3.2 Techno-Economic Feasibility Assessment

The techno-economic feasibility assessment is a vital tool to justify the economic viability of
POMR-RES. This assessment allows the detailed descriptions and estimation of potential costs,
anticipated revenues and relative profitability of the system. The assessment was performed
originally from methodologies presented by Yeoh et al. (2004) and Sadhukhan, Ng and
Hernandez (2014). According to these methods, the techno-economic assessment will be
conducted in three phases: first, the cost estimation; second, the revenue interpretation; third,
the cash flow and profitability evaluation. The European Commission (2014) suggested an

93
economic life of 15 years of a feasible bioenergy plant. Similarly, Abdul Malek et al. (2017)
used 15 years plant lifetime to assess the techno-economic performance of biomass-based
power plant in Malaysia. Furthermore, under renewable energy power purchasing agreement
(REPPA), the long term expectation of commitment to supply RE to the grid is fixed at 15-16
years depending on the type of resources used to generate electricity. Therefore, a techno-
economic spreadsheet was developed to assess the economic feasibility of the system
corresponding to 15 years of economic life. Additionally, the techno-economic assessment was
also conducted for 20 years of power plant economic life.

5.3.2.1 Cost Estimation

The cost to develop an energy system comprises capital cost and operation cost (European
Commission, 2014). Sadhukhan, Ng and Hernandez (2014) defined capital cost as the cost of
building the energy system that includes direct and indirect capital costs and working capital.
The operation cost is equally defined as the cost to run the energy system, which is divided into
fixed and variable operating costs. Two cost estimation methods were used in this chapter,
namely process engineering design cost estimation and direct cost to capacity correlation
estimation.
The cost of equipment to build the system uses the base cost and size data as listed in Table
5.2. Studies by Wan et al. (2016), Lam et al. (2013) and Patel et al. (2011) were used as
references for equipment size and cost. The list was verified by the mill owners when the
researcher visited the mills. A critical problem faced by the researcher during cost estimation
stage was limited access to quotes from vendors, hence there could be variations between cost
estimation accuracies.
Table 5.2: POMR-RES Equipment List
No. Items Reference Size Reference Cost
(million US$)
1 Conveyers 0.87 t/hr (wet basis) 0.0019
2 Feedstock Feeding System 0.87 t/hr (wet basis) 0.0011
3 Boiler 0.62 kg/s 0.4323
4 Back Pressure Steam Turbine 231 kW 0.2371

The estimated cost of equipment is assumed to be driven by economies of scale, therefore, the
power law in Eq. (14) is applied to calculate the cost of equipment at a desired size (Wright
and Brown, 2007).

94
𝐶𝑜𝑠𝑡2 𝑆𝑖𝑧𝑒2 𝑆
= ( ) (14)
𝐶𝑜𝑠𝑡1 𝑆𝑖𝑧𝑒1

The power law reckons that the equipment used can be scaled up or down exponentially with
the scaling factor (S) where 𝐶𝑜𝑠𝑡1 and 𝑆𝑖𝑧𝑒1 are the reference cost and base size of the
equipment while 𝐶𝑜𝑠𝑡2 and 𝑆𝑖𝑧𝑒2 are the desired cost and size of the equipment, respectively.
As suggested by Ngyuen and Prince (1996) and Kumar, Cameron and Flynn (2003), scaling
factor of 0.8 was used in this study to represent the effect of economies of scale. The reference
size and cost of the equipment in Table 5.2 were used.
In due diligence to the time value of money, the cost of equipment was updated to the cost at
the current year (𝐶𝑜𝑠𝑡2𝐶 ) using the cost index method as in Eq. (15) (Sadhukhan, Ng and
Hernandez, 2014).

𝐼𝐶
𝐶𝑜𝑠𝑡2𝐶 = ( ) 𝑥 𝐶𝑜𝑠𝑡2𝑅 (15)
𝐼𝑅

Here, 𝐼𝐶 and 𝐼𝑅 are the current cost index and reference cost index value respectively, and
𝐶𝑜𝑠𝑡2𝑅 is the reference for cost. The Chemical Engineering Plant Cost Index (CEPCI) is used
as the cost index value (Ng and Sadhukhan, 2011; Patel et al., 2011; Sadhukhan, Ng and
Hernandez, 2014). In this recent assessment, CEPCI for the year 2014 (574.4) was used as the
reference cost index value and CEPCI for August 2017 (571.9) as the current cost index value.
The cost of equipment is assumed to be free on board which means that the delivery cost of
equipment has not been included (von Sivers and Zacchi, 1995; van Kasteren and Nisworo,
2007; Sadhukhan, Ng and Hernandez, 2014). Furthermore, the total cost of equipment was
calculated by summing up the cost of individual equipment in Table 5.2.

The Lang’s method is used to estimate capital cost required to build the system (Amigun and
Von Blottnitz, 2009; Lemmens, 2015). This method allows direct estimation of capital cost by
multiplying the Lang Factor (LF) value with the total cost of equipment (Wain, 2014). The LF
used varies depending on the types of generation plant as tabulated in Table 5.3 (von Sivers
and Zacchi, 1995; Haas et al., 2011). This approach is recommended by the American
Association of Cost Engineer for any installation of new chemical and processing plants with
the ability to provide an estimation accuracy of 20%-30% (Sadhukhan, Ng and Hernandez,
2014). The LF for solid processing was used to estimate the capital cost of EFB-based
generation plant while the LF for fluid processing was used for biogas-based generation plant.

95
Table 5.3: Lang Factor for Capital Cost Estimation
Plant Types
Cost Parameters Solid Processing Fluid Processing
Total Direct Capital Cost 2.69 3.60
Total Indirect Capital Cost 1.28 1.44
Working Capital 0.70 0.89
Total Capital Cost 4.67 5.93
(TDC + TIC +WC)

Two other potential costs that are the administrative cost and new grid installation cost were
also included in cost estimation. This inclusion was made to reflect the current practice in
Malaysia’s renewable energy landscape, where the RE developers practically bear these costs.
The administrative cost was charged differently depending on the scale of the POMR-RES.
This cost includes the FiAH application and processing fees, provisional and permanent public
generation licensing fees, power system study fees and acceptance and reliability test run fees
(SEDA, 2018). The new grid installation cost was quoted at US$ 227,300/km (Tenaga Sulpom
pers. comm., 27/09/2017; Havys Oil Mill Sdn. Bhd pers. comm., 06/06/2016). To date, there
is little agreement on the reasonable distance of the new grid installation for RE generation in
Malaysia (Akikur et al., 2013). Akikur et al. (2013) and Borhanazad et al. (2013) used 5 km
and 10 km as a reasonable distance to install new grid for the solar hybrid power system in
Malaysia. A sustainability study of power generation from POMR by Aghamohammadi et al.
(2016) indicated that additional new grid installation cost for more than 10 km has a substantial
impact on the techno-economic performance of the power plant. For this reason, in this thesis,
the cost of a new grid installation was estimated for a distance of 5 kilometers as suggested by
Akikur et al., (2013). Similar suggestions are also gathered from the millers’ and authorities’
personnel. A sensitivity analysis of the selection was conducted using 10 km and 20 km,
respectively.

On the other hand, the fixed and variable operating costs of the POMR-RES were estimated on
an annual basis. Lemmens (2015) stated the yearly operating cost as the summation of the fixed
and variable operating costs. The suggested operating cost specification from Sadhukhan, Ng
and Hernandez (2014) was adopted in this chapter. The fixed operating cost, that includes
maintenance, labours, plant overheads, capital charges, insurance, and taxes, can be determined
in relation to indirect capital cost and labour cost. Two, four and six workers assumed to be
hired for small-, medium- and large-scale POMR-RES. It was stated that the workers would

96
work in two shifts, each shift lasting for 10 hours. It was further assumed that the workers
would receive a payment of US$10 per hour (Wan et al., 2016).
On the contrary, the variable operation cost, that comprises the feedstock cost, feedstock
transportation cost and utility cost, depends on the electricity generation plant’s scales. For the
on-site POMR-RES, the feedstock was obtained at no cost as it was previously being treated
as a waste from the mills. Therefore, no weightage was given to the cost of purchasing
feedstock. The feedstock cost will become a liability when:
1. Off-site POMR-RES were built by investors other than mill owners. The EFB was
bought from the millers to sustain the operation of the system.
2. EFB supplies are required for the cooperative POMR-RES.
The estimated cost of wet EFB was about US$ 4.50 – US$ 13.80 per tonne (Kazi et al., 2010;
Abdul Malek et al., 2017). Ng et al. (2014) segregated the cost for wet and dry EFB by
providing a price variation of US$10 – US$ 50 per tonne of wet EFB and US$50 – US$ 110
per tonne of dry EFB, respectively. Here, the price of semi-dried EFB was fixed at US$
5.00/tonne as recommended by the millers; this fact was told during the interview sessions with
the millers’.
Truck was selected as the primary feedstock transportation mode for two reasons: first, most
of the mills are accessible on a paved road; second, trucks have a negligible economy of scale
and have more stable distance variable component (Calvete and Gal, 2010; Leboreiro and
Hilaly, 2011). Various methods have been used in literature to estimate the feedstock
transportation distance and cost (Lauven, Liu and Geldermann, 2013; Reeb et al., 2014). Nagel
(2000) ascertained that the feedstock transportation distances have a significant influence on
the economic feasibility of biomass-based energy system. Foo et al. (2013) in their study used
10 km to 180 km as the feedstock transportation distance to assess the palm oil-based bioenergy
supply chain. Hoeltinger, Schmidt and Schmid (2012) recommended 10 to 40 km as the
feedstock transportation for the biomass-based power plant. According to the supply chain
analysis and LCA of EFB for bioenergy production in Malaysia by Reeb et al. (2014), the
optimal feedstock transportation distance is estimated at 10 km (i.e., 20 km for two-way). The
feedstock transportation distance in this thesis is set based on Reeb et al. (2014) with sensitivity
analysis conducted for 20 km and 30 km of feedstock transportation distance.

The feedstock transportation cost was computed analogously using the Peninsular Malaysia’s
feedstock transportation cost linear equation as shown in Table 5.4 (Ong et al., 2016). Ten-
tonne trucks were used to transport the EFB in this study as the larger trucks are commonly

97
used to transport dry or finished product while smaller truck possesses subtle economic
disadvantages (Ong et al., 2016).

Table 5.4: Feedstock Transportation Linear Equations


Truck Size Transportation Cost Linear equations
(tonne)
1 MYR/tonne = distance (km) × 1.89 + 132.00
3 MYR/tonne = distance (km) × 0.67 + 69.10
10 MYR/tonne = distance (km) × 0.26 + 49.30
26 MYR/tonne = distance (km) × 0.19 + 39.50

The transportation cost of a tonne of EFB was specified at US$ 12.55, US$ 13.15 and US$
13.75 for the distance of 20 km, 30 km and 40 km, respectively. This cost was found to be
moderately higher than the cost interpreted by Lam et al. (2013). The transportation of biogas
using gas pipelines or trucks is relatively unusual in Peninsular Malaysia even though there is
an extensive natural gas pipelines network (Wahid and Ujir, 2012; Hosseini and Abdul Wahid,
2014a, 2014b). Therefore, biogas was only used for on-site electricity generation to minimize
the transportation uncertainty. Other costs such as sales expenses, general overheads and
research and development were also included, bringing an additional 20% (0.2 × (fixed
operating cost + variable operating cost)) to the annual operating costs.

For the second cost estimation method, the capital cost of the POMR-RES was estimated based
on installation cost of one-megawatt electricity generation capacity. For this method, detailed
information on the equipment used was not provided. A reasonable capital cost of US$1.85 to
2.30 million for one-megawatt installation was advised by SEDA (SEDA email comm.,
26/03/2018). The maximum cost was used as the base cost in this part of the assessment. Eq.14
and Eq.15 were applied to introduce economies of scale function and to concede the time value
of money in this cost estimation.

The operating cost for the second cost estimation method, however, has not been articulated
clearly. Kazi et al. (2010) proportionate the annual operating cost of the system as 10% of the
capital cost. Chin et al. (2013) exercised 6% as the correlation factor to represent the annual
operating cost of biomass-based energy generation plant and found to be in a good agreement
with the estimation made by the International Renewable Energy Agency (IRENA, 2012).
However, no clear separation between fixed and variable operating cost has been made in these
two cost correlations. In this assessment, the fixed operating cost for POMR-RES was set at

98
10% of the capital cost estimated earlier while the variable operating cost was varied depending
on the scale and the POMR-RES installation requirements.

5.3.2.2 Revenue Interpretation

The fixed revenues for the POMR-RES were made up of the income from FiT payment and
cost saving made by the millers from the steam and electricity self-sufficiency. The FiT
payment could be received based on the amount of electricity connected to the grid. The basic
payment and bonus rate of per kWh electricity used in this revenue interpretation following the
rate established by SEDA for biomass- and biogas-based electricity generation plant are
tabulated in Table 5.5. It was assumed that for every kWh electricity connected to the grid from
EFB-based POMR-RES, it is entitled to receive the basic FiT rate and bonus rates (Item ii-iv).
Similar entitlement was assumed for biogas-based POMR-RES excluding the bonus rate for
using gas engine technology.

Table 5.5: FiT Rate Entitlement for POMR-RES

Resources Description Rate


(US$/per kWh)
Biomass
Basic FiT rates having installed capacity of:
i. up to and including 10 MW 0.0711
ii. above 10 MW up to including 20 MW 0.0665
iii. above 20 MW up to including 30 MW 0.0619

Bonus FiT rates having the following criteria


(one or more):
i. use of gasification technology +0.0046
ii. use of steam-based electricity generation +0.0023
system with overall efficiency on above 20%
iii. use of locally manufactured or assembled +0.0115
boiler or gasifier
iv. use of solid waste as fuel source +0.0226

Biogas
Basic FiT rates having installed capacity of:
i. up to and including 4 MW 0.0733
ii. above 4 MW up to including 10 MW 0.0688
iii. above 10 MW up to including 30 MW 0.0642

Bonus FiT rates having the following criteria


(one or more):
i. use of gas engine technology with +0.0046
electrical efficiency of above 40%
ii. use of locally manufactured or assembled +0.0115
gas engine technology
+0.0181

99
iii. use of landfill, sewage gas or agricultural
waste including animal waste as fuel source

The cost that could be avoided from having to generate steam using other fuel and buying the
electricity from the grid was treated as indirect revenue to the millers. The price of producing
one-tonne steam was US$ 26, and the electricity tariff from the grid was US$ 0.101/kWh under
the special agricultural tariffs (Andiappan, Ng and Bandyopadhyay, 2014; Malaysia Energy
Commission, 2018).
5.3.2.3 Cash Flow and Profitability Evaluations

Cash flow and profitability evaluations were used to forecast the economic viability and
profitability of POMR-RES. Cash flow statement was developed to provide detailed
information of cash transactions and to evaluate the profitability of the system over a period of
15 years as suggested by Sadhukhan, Ng and Hernandez (2014). The profitability evaluation
was conducted for two project funding mechanisms. Firstly, the private-funded mechanism
assumes that the plant received full financial assistance from the financial institution with a
fixed interest rate of 12% per annum (Havys Oil Mill Sdn. Bhd 2016 pers. comm.,06/06/2016).
Secondly, the self-funded mechanism assumes that the plant owner invested the capital from
internal resources, thus incurring no interest cost. Two profitability indicators were used in this
assessment (i.e., ROI and NPV) together with PP and break-even point (BEP). ROI is used to
assess the efficiency of investment made over a short-term time frame while NPV is commonly
used to calculate the rate of return from the investment taking into account the time value of
money (Sadhukhan, Ng and Hernandez, 2014). The system with positive ROI and NPV, and
PP and BEP no later than the economic lifetime was considered as economically feasible
(Clean Development Management, 2006; Seabra and Macedo, 2011). A discount rate of 12%
is used to calculate NPV of the private-funded POMR-RES, “reflecting the hurdle rate they
might use when exploring an investment (Leach M., email comm., 10/02/2019). The interest
rate is according to the financial institution is fixed at 12%. Differently, a discount rate of 5%
is used for the self-funded POMR-RES as suggested by Tabe-Ojong and Habiyaremye (2017)
for social bioenergy projects.
5.3.2.4 Levelized Cost of Electricity

According to Shum (2017), the levelized cost of electricity (LCOE) is the cost of using the
energy conversion technology to generate per unit electricity. IRENA, in their RE cost analysis
paper, describes LCOE as the price of the electricity generated in a power plant (IRENA, 2012).

100
The estimated LCOE for every kilo-watt hour biomass-based electricity ranges between
US$0.06 and US$0.29 (IRENA, 2012). Eq. (16) was used to project the LCOE generated in
POMR-RES using the annual cost of the system to the amount of the electricity generated.

𝐴𝑛𝑛𝑢𝑎𝑙 𝐶𝑎𝑝𝑖𝑡𝑎𝑙 𝐶𝑜𝑠𝑡 + 𝐴𝑛𝑛𝑢𝑎𝑙 𝑂𝑝𝑒𝑟𝑎𝑡𝑖𝑜𝑛 𝐶𝑜𝑠𝑡


𝐿𝐶𝑂𝐸1𝑘𝑊ℎ = (16)
𝐴𝑛𝑛𝑢𝑎𝑙 𝐴𝑚𝑜𝑢𝑛𝑡 𝑜𝑓 𝐸𝑙𝑒𝑐𝑡𝑟𝑖𝑐𝑖𝑡𝑦 𝐺𝑒𝑛𝑒𝑟𝑎𝑡𝑒𝑑

5.4 Sensitivity Analysis

The robustness of the POMR-RES performance was investigated using one-at-a-time


sensitivity analysis method. The sensitivity analysis focuses more on the techno-economic
performance of the POMR-RES. The variations in the profitability indicators were observed
when one of the sensitive variables was varied at a time (Pianosi et al., 2016). Two sets of
sensitive variables were used. First, from the project costing perspective, e.g., interest rate, grid
installation cost, feedstock, and feedstock transportation cost. Second, from project revenue
perspective, e.g., electricity and steam cost avoidance and FiT rate.

5.5 Outcomes Illustration

Three different case studies were used to demonstrate the effectiveness of the proposed
framework. These case studies allow the assessment of the technical and techno-economic
performance of the on-site, off-site and cooperative POMR-RES installation. The technical
assessment spreadsheet in Appendix C-1 was used to assess the technical feasibility of the
system according to Eq. (13). This equation explained that the electricity generated must be
sufficient to accommodate the parasitic load from the electricity generation plant, electricity
demand from the mills with surplus electricity to be connected to the grid. For the CHP plant,
the amount of heat supplied to the mills was also used to justify the feasibility of the system.
The electricity and heat generation potential from both residues established in Chapter 4 was
applied in this assessment. On the other hand, the system was considered to be techno-
economically feasible when it produces a positive ROI and an ability to pay back the
investment made within the economic lifetime of the plant. The techno-economic spreadsheets
in Appendix C-2 were used to assess the techno-economic performance of POMR-RES. Based
on the technical and techno-economic performance indicators; the feasibility of each case study
was compared and ranked. From the ranking, suggestion for suitable residues for electricity
generation and recommendation for the appropriate installation scheme for POMR-RES were
made.

101
5.5.1 OnS case study: On-site POMR-RES installation

This section of the thesis presents the technical and techno-economic performance of the on-
site POMR-RES installation. The performances of the EFB based POMR-RES is presented
first and followed by the biogas.

5.5.1.1 EFB-based POMR-RES


5.5.1.1.1 EFB-based Power Plant

The technical performance of the EFB-based power plant is presented in Table 5.6. The results
confirmed the technical feasibility of EFB-based power plant that, on average, 59% of the
electricity generated was available for grid connection. This agrees with the previous findings
that appraised the technical feasibility of generating electricity from EFB (Chiew, Iwata and
Shimada, 2011; Ng et al., 2011; Mahlia et al., 2012; Ansori, Herawan and Rivani, 2015).

Table 5.6: Technical Performance of EFB-based Power Plant


Plant Total Electricity Parasitic Load Mill Electricity Net Electricity
Size Generated Demand Available for Grid
(MW) (MWh/yr) (MWh/yr) (MWh/yr) (MWh/yr)
0.9 7122 2160 783 4179
1.8 14246 4320 1567 8359
2.7 21368 6480 2351 12538
3.6 28491 8640 3134 16717
4.4 35614 10800 3918 20896
5.3 42737 12960 4701 25076
6.2 49859 15120 5485 29255
7.1 56982 17280 6268 33434
8.0 64105 19440 7052 37613
8.9 71228 21600 7835 41793
9.8 78350 23760 8619 45972
10.7 85473 25920 9402 50151

Next, the techno-economic feasibility of the 0.9 MW to 10.7 MW private-funded and self-
funded power plant was assessed using the two cost estimation methods (process engineering
design and direct cost to capacity) and a 12% annual interest rate. Firstly, Table 5.7 provides
summary of the estimated capital and operating cost for the power plant. It can be observed
that capital cost of power plants was estimated at US$ 6.60 million – US$ 40.70 million using
the first method and US$ 3.30 million – US$ 16.50 million using the second method. The
operation cost was estimated at US$ 0.10 million – US$ 0.70 million and US$ 0.30 million –

102
US$ 1.70 million, respectively. There are remarkable cost differences between the two methods
adopted in this chapter. The capital cost estimated according to the first cost estimation method
was found to be about double than the cost estimated using the second method. For instance,
the capital cost for the 0.90 MW power plants was determined to be US$ 6.60 million using
the first method while about US$ 3.30 million was required as a capital according to the second
method. This cost differences are due to the adaptation of Lang’s Factor in the first cost
estimation method. On the contrary, the operating cost estimated using the first method was
found to be lower than the second method. The variation in the estimated costs has resulted in
the annual cost estimation inconsistencies.

Table 5.7: Cost Estimation of On-Site EFB-based Power Plant


Plant Size Capital Cost (million US$) Operating Cost (million US$/yr)
(MW) Method 1 Method 2 Method 1 Method 2
0.9 6.60 3.30 0.10 0.30
1.8 10.70 4.90 0.10 0.50
2.7 14.30 6.30 0.20 0.60
3.6 17.70 7.60 0.30 0.80
4.4 20.60 8.70 0.40 0.90
5.3 23.70 9.90 0.40 1.00
6.2 26.70 11.10 0.40 1.10
7.1 29.70 12.20 0.60 1.20
8.0 32.50 13.30 0.60 1.30
8.9 35.30 14.40 0.70 1.40
9.8 38.00 15.50 0.70 1.60
10.7 40.70 16.50 0.70 1.70

103
Based on the cost estimated above, the profitability analysis of the EFB-based POMR-RES is
presented using ROI first followed by NPV. The outcome of the cash flow and ROI analysis
of the power plant is summarized in Table 5.8. The table shows that the higher capital cost
estimated using the first cost estimation method has reduced the probability of the power plant
to produce a positive ROI in comparison to the second cost estimation method. The table also
shows that using the second cost estimation method, the self-funded power plant starts to
provide a positive ROI when the power plant reached the capacity of 2.7 MW.

From a series of interviews and consultations with millers and respective authorities, in regard
to the noticeable techno-economic performance differences and unfamiliarity of the Lang’s
Factor concept, the first cost estimation method was found to be impractical for the under-
developed POMR-RES landscape in Peninsular Malaysia (SEDA pers. comm., 20/09/2017;
Tenaga Sulpom pers. comm., 27/09/2017; Havys Oil Mill Sdn. Bhd 2016 pers. comm.,
06/06/2016). Hence, the second cost estimation method that is the direct cost to capacity
correlation was used for assessing techno-economic feasibility of the POMR-RES in this study.

104
Table 5.8: Techno-Economic Performance of EFB-based Power Plant based on ROI
Plant Size (MW) 0.9 1.8 2.7 3.6 4.4 5.3
Cost Estimation Method 1 2 1 2 1 2 1 2 1 2 1 2
Private-funded
ROI (%) -516 -473 -488 -401 -471 -357 -459 -324 -450 -293 -442 -272
PP (years) >15 >15 >15 >15 >15 >15 >15 >15 >15 >15 >15 >15
LCOE (US$/kWh) 0.32 0.18 0.31 0.15 0.28 0.13 0.26 0.10 0.25 0.10 0.24 0.09
Self-funded
ROI (%) -167 -141 -127 -47 -104 11 -87 53 -77 94 -65 120
PP (years) >15 >15 >15 10.00 >15 7.60 >15 6.40 >15 5.60 >15 5.20
LCOE (US$/kWh) 0.07 0.06 0.07 0.04 0.06 0.04 0.06 0.03 0.05 0.03 0.05 0.03

Plant Size (MW) 6.2 7.1 8.0 8.9 9.8 10.7


Cost Estimation Method 1 2 1 2 1 2 1 2 1 2 1 2
Private-funded
ROI (%) -435 -254 -433 -239 -428 -225 -422 -212 -418 -202 -413 -204
PP (years) >15 >15 >15 >15 >15 >15 >15 >15 >15 >15 >15 >15
LCOE (US$/kWh) 0.23 0.09 0.22 0.08 0.22 0.08 0.21 0.08 0.21 0.07 0.20 0.07
Self-funded
ROI (%) -57 143 -56 163 -48 181 -41 197 -34 210 -28 208
PP (years) 11.20 5.00 11.20 4.70 10.60 4.50 10.00 4.40 10.00 4.30 9.50 4.30
LCOE (US$/kWh) 0.05 0.03 0.05 0.03 0.05 0.03 0.05 0.03 0.05 0.03 0.04 0.03

105
Besides that, Table 5.8(a) also shows the significant techno-economic performance differences
between the self-funded and private-funded power plants. These differences are mainly due to
high annual rates imposed by financial institution. In this case study, the techno-economic
performance of the private-funded power plants was severely affected by the 12% interest rate
that has been levied on the loan amount. For instance, US$ 3.30 million is required as an initial
capital to develop a 0.90 MW power plant. With a fixed annual interest rate, the loan repayment
amount is estimated at US$ 9.30 million where 64% of this amount is used to repay the interest
charged. Several studies, for example, Verbrugen at al. (2010), Yusoff and Kardooni (2012)
and Yoshizaki et al. (2013) highlighted the direct impact of interest rate towards the techno-
economic performance of the bioenergy plant. Deng and Parajuli (2016), ACE (2016) and Lam
et al. (2017) emphasized that the economic survival of the bioenergy plant was dependent on
the ability to be developed in a ‘conducive environment’. Such environment includes the
provision of a reasonable annual interest rate and straightforward loan application process. Due
to this fact, it was concluded that higher interest rate and complicated loan application process
are among the barriers to the integration of RE-based power plant into the existing electricity
generation system.

In addition to the high-interest rate, an extra cost for new grid installation was also identified
as a contributing factor to the poor techno-economic performance of the private-funded project.
The new grid installation cost contributes a significant share to the total project cost depending
on the scale of the power plant. For instance, the cost to install 5 km new grid added 35% and
7% to the total project cost of a small-scale plant and large-scale plant, respectively. From the
sensitivity analysis conducted, the cost to install 10 km and 20 km new grid connection have a
higher impact on the total project cost of the small-scale power plant than the large-scale power
plant. This cost, however, can be considered as an avoidable cost if there is a standard
coordination between the electricity generators and grid operators. The grid operators can either
allow electricity to be connected to the current grid automatically or are willing to share the
responsibility by providing new grid infrastructure when it is required (IRENA, 2018).
However, there is an on-going debate until today to find the best strategy to balance this cost
and to name the responsible actor for this cost obligation. Swider et al. (2008) named grid
operator as a sole responsible party to install, maintain and operate the grid. This has been
opposed by Hiroux and Saguan (2010) whereby it allows the cost recovery to be made by the
grid operator through electricity tariff increment. In order to address the issue of new grid
installation, Hiroux and Saguan (2010) recommended a reasonable responsibility sharing

106
between the electricity generators and grid operators as the appropriate mechanism to smoothen
the process of integrating RE in the existing grid system (Hiroux and Saguan, 2010).

These two contributing factors have been found to further influence the LCOE for the system.
The LCOE of the private-funded power plant was found to be a lot higher than the FiT rate
offered. The FiT rate for every kWh of electricity connected to the grid is US$ 0.11 while the
LCOE for 0.90 MW power plants was estimated at US$ 0.18. This result shows that the current
FiT rate offered to RE from biomass is incapable of neither increasing the investment certainty
nor providing reasonable profits to the investors (Couture, Cory and Williams, 2010; Ahmad,
Ab Kadir and Shafie, 2011; Polzin et al., 2015). This LCOE estimation is also found to be 41%
higher than the LCOE in selected ASEAN countries (ACE, 2016). Inversely, the LCOE of the
self-funded power plant was found to be competitive with the current FiT rate and falls within
the LCOE from biomass range established by IRENA (2012) and REN 21 (2013). The price of
generating electricity in the self-funded plant was also found to be competitive with the average
electricity generation cost in India and China (IRENA, 2018). However, to date, the results on
the LCOE of the POMR-RES have not been described or published in previous studies.

From the cash flow analysis conducted, it is found that prolonging the economic life to 20 years
and 30 years can increase the amount of interest repayment to the financial institution. For
instance, the interest repayment for 0.9 MW increases from US$ 6 million for 15 years to US$
8 million and US$ 12 million for 20 years and 30 years economic life, respectively. The
increment in the interest rate payment can further reduce the ROI value of the power plant.

Next, the discounted cash flow (DCF) analysis to estimate the NPV at the 15th year of the EFB-
based power plant is presented in Table 5.8(b). The NPV is determined based on cost estimated
using the second cost estimation method to analyse the profitability of the investment. The
result in Table 5.8(b) indicates that the power plant under private funding mechanism has
negative NPV which represents that the sum of the cash flows is less than the initial investment
made. However, the self-funded power plant starts to provide a positive NPV when the power
plant reached the capacity of 1.8 MW, with BEP of this power plant is found to be in the twelfth
year. Based on the DCF analysis in Appendix C-2, factors such as interest rate imposed by the
financial institution, discount rate used in the DCF analysis and grid installation cost influence
the net present value (NPV) of the power plant.

107
Table 5.8(b): Techno-Economic Performance of EFB-based Power Plant based on NPV
Plant Size (MW) 0.9 1.8 2.7 3.6 4.4 5.3 6.2 7.1 8.0 8.9 9.8 10.7
Private-funded
NPV (mill US$) -8.50 -8.60 -9.20 -9.40 -8.90 -8.80 -8.40 -8.00 -7.50 -6.80 -6.10 -6.90
BEP (years) >15 >15 >15 >15 >15 >15 >15 >15 >15 >15 >15 >15

Self-funded
NPV (mill US$) -2.50 0.60 3.70 7.00 10.90 14.6 18.3 22.1 26 30 34.1 35.8
BEP (years) >15 12 7 6.5 5.3 5 4.4 4.2 4 3.7 3.6 3.6

108
5.5.1.1.2 EFB-based CHP Plant

The technical feasibility of CHP plant was assessed based on the amount of electricity and heat
generated. The assessment for the electricity generation was conducted similar to the technical
assessment spreadsheet in Appendix C-1. The feasibility of heat generation was analysed based
on the amount of heat that was to be supplied to the mill. Subsequently, the techno-economic
feasibility of the plant was assessed according to the techno-economic spreadsheet in Appendix
C-2. The performances of the CHP plant are summarised in Table 5.9, which shows that the
electricity generation capacity of the CHP plant decreases with the decreasing plant size. On
average, the electricity generation capacity in the CHP plant is 75% lower than the power plant
installed in the similar mill. The lower electricity generation capacity in CHP plant is due to
variation on the magnitude of work produced in BPST resulting from fixing higher pressure
exit steam for the turbine (2.5 bar, 145°C). Although the electricity generated is able to satisfy
the parasitic load and the mill’s electricity demand, less amount of surplus electricity is
available to be connected to the grid. This reduced the contribution of CHP plants toward the
national target of having 800 MW palm oil biomass-based grid-connected electricity by 2020.

The significant electricity generation capacity reduction indicates the low applicability of the
CHP plant to maximize the electricity generation potential from EFB. Raj, Iniyan and Goic
(2011) in his review of RE-based cogeneration technologies concluded that although the
cogeneration technology provides a higher overall system efficiency, it is not the best
installation option when maximizing the electricity generation becomes the main priority. This
view is also supported by Ali, Tesfamichael and Hassan (2014) who observed the CHP plant
performance in regard to the variation of the inlet and outlet steam pressure. In the same case
study, Raj, Iniyan and Goic (2011) had proposed a practical solution to increase the electricity
generation capacity of the CHP plant. This was done by replacing the BPST with CEST. A
simulation study by Alves et al. (2015) for the CHP plant with BPST and CEST had accurately
justified Raj’s proposal by acknowledging the capability of CEST to increase the CHP plant’s
electricity generation capacity. It stated that the CEST is able to improve the electricity
generation’s capacity of the CHP plants by 23%; however, these may lead to an increment in
the capital cost. The CEST is known to be more expensive than BPST. The concern on the
performance and cost variations between these two types of turbine has been further
emphasized by IRENA (2012). In this report, IRENA (2012) claimed that BPST is cheaper

109
than CEST; however, they are less suitable for stream extraction purposes. Nevertheless, there
is not much cost and benefit analysis conducted to compare these two types of turbines.

Next, the technical performance of the CHP plant was also measured by the amount of heat
that can be supplied to the mills. Based on the simulation results of the EFB-based CHP plant,
about 14,100 to 169,400 tonnes of medium pressure steam (~ 2.5 bar, 145°C) were extracted
from the BPST. This estimation accounted for approximately 27% of the total heat demand
from the mills. This finding was conflicting with the information previously gathered from
interviews with the millers. However, this estimation holds some similarities to the conclusions
made by Çakir, Çomakli and Yüksel (2012). They described that the heat extracted in the CHP
plants is able to fulfil roughly up to 58% of the heat demand from the designated industrial
process where an additional boiler is required to meet the remaining heat demand.

Likewise, in the previous section, the economic feasibility of the CHP plant was assessed using
the similar approach presented in Appendix C-2. The simultaneous generation of electricity
and heat has been profitable when most plants earn positive ROI except for the small-scale
private-funded plant. This result can be explained by the fact that the CHP plant benefitted
from the heat self-sufficiency cost saving. This cost saving contributes to 60% of the total
revenue of the system. This result also reveals that a cost saving can outweigh the impact of
the high interest rate on the techno-economic performance of the plant. Even though the CHP
plant shows an outstanding techno-economic performance, the LCOE from the system is 33%
- 42% higher than the LCOE of the power plant described in Section 5.5.1.1.1.

110
Table 5.9: Technical and Techno-Economic Performance of On-Site EFB-based CHP Plant
Plant Size (MW) 0.30 0.70 1.00 1.20 1.70 2.10 2.40 2.70 3.00 3.50 3.70 4.10
Technical
Electricity Generated (GWh/yr) 2.40 5.60 8.00 9.60 13.60 16.80 19.20 21.60 24.00 28.00 29.60 32.80
Parasitic Load (GWh/yr) 0.30 0.60 0.90 1.00 1.50 1.90 2.10 2.30 2.60 3.10 3.20 3.60
Mill Electricity Demand (GWh/yr) 2.10 4.30 6.50 8.60 10.80 13.00 15.10 17.30 19.40 21.60 23.80 25.90
Electricity to Grid (GWh/yr) 0.00 0.70 0.60 0.00 1.30 1.90 2.00 2.00 2.00 3.30 2.60 3.30
Steam feed to Mill (k t/yr) 14.10 28.30 41.90 56.10 70.20 85.60 96.70 112.70 126.30 141.70 154.00 169.40
Techno-Economic
Private-Funded
ROI (%) -393 -278 -158 -73 -70 22 53 99 132 156 157 210
PP (years) >15 >15 >15 11.70 11.70 6.30 5.50 4.70 4.30 4.10 4.10 3.60
LCOE (US$/kWh) 0.33 0.18 0.17 0.14 0.13 0.13 0.12 0.11 0.11 0.10 0.12 0.12
Self-Funded
ROI (%) -36 113 265 266 375 499 541 608 649 674 676 746
PP (years) 9.60 5.30 3.80 3.80 3.30 2.90 2.80 2.60 2.50 2.50 2.50 2.40
LCOE (US$/kWh) 0.11 0.06 0.06 0.05 0.05 0.04 0.04 0.04 0.04 0.03 0.03 0.03

111
Moreover, the DCF analysis of the private funded EFB-based CHP plant has shown that the
plant with 1.20 MW capacity starts to produce positive NPV and the biggest plant with 4.1
MW capacity has NPV value of US$ 2.7 million. The break-even point of these plants varies
from 14 years to 2 years, respectively. Subsequently, all the self-funded EFB-based CHP plants
have positive NPV (US$ 3.3 million to US$ 8.9 million) with BEP between 10.5 years to 1
year.

5.5.1.2 Biogas-based POMR-RES

This section discusses the technical and techno-economic performances of biogas-based


POMR-RES. Performances of biogas-based power plants are first presented followed by the
CHP plant. Generally, a lower electricity generation capacity was predicted from biogas. From
Table 5.10, it can be observed that although the biogas-based POMR-RES power plant is
considered as technically feasible, the contribution to the grid is minimal. Only about 0.5%-
2% of the total electricity generated is available for the grid connection. The techno-economic
assessment, on the other hand, revealed that the private-funded power plants are exceptionally
uncompetitive with all the plants projecting negative ROIs (-324% to -557%) with the PP
projected longer than 15 years. Conversely, for the self-funded power plant, the plant was
forecasted to produce a positive ROI when the capacity exceeds 2.70MW.

However, different technical and techno-economic performances were observed for the biogas-
based CHP plant. Table 5.11 shows that the amount of electricity generated in CHP plants is
insufficient to meet the demand from the mills. Therefore, no surplus electricity was available
for grid connection. With the inability to fulfil the electricity demand from the mills, the biogas-
based CHP plant was classified as technically infeasible. Furthermore, only 4% of the steam
demand from the mills was supplied by the medium-pressure steam extracted from the steam
turbine. From the techno-economic assessment conducted, it appeared that the cost saving from
the heat and electricity self-sufficiency, and FiT payment was inadequate to recover the plant
cost. The negative ROI and PP beyond 15 years signify the unfeasibility of private-funded CHP
plant. The factors behind the poor techno-economic performance of biogas CHP plant are
closely related to the effects of the economies of scale, high-interest rate together with
additional cost for grid installation, discussed earlier.

The results provided in Table 5.11 were contradicting the previous results reported by Yeoh
(2004) and Chin et al. (2013). Both of the studies had positively appraised the technical and
techno-economic performance of the biogas-based CHP plant. This unsatisfactory performance

112
is also inconsistent with the previous results reported in the literature (Seabra and Macedo,
2011; Hunpinyo et al., 2013; Lauven, Liu and Geldermann, 2013). The inconsistencies of the
results obtained in this study with the previous studies were entirely understood. Most of the
previous studies were conducted for a larger-scale biogas CHP plant that is profoundly known
to have higher economic advantages (Walla and Schneeberger, 2008; White, Kirk and
Graydon, 2011). From this technical and techno-economic elaboration, one question needs to
be asked, whether the biogas-based POMR-RES installation is essential.

113
Table 5.10: Feasibility Assessment of On-Site Biogas-based Power Plant
Plant Size (MW) 0.30 0.60 0.90 1.20 1.50 1.80 2.10 2.40 2.70 3.00 3.30 3.60
Technical
Electricity Generated (GWh/yr) 2.40 4.90 7.30 9.70 12.20 14.60 17.00 19.50 21.90 24.30 26.80 29.20
Parasitic Load (GWh/yr) 0.30 0.50 0.80 1.00 1.30 1.60 1.90 2.10 2.40 2.70 2.90 3.20
Mill Electricity Demand (GWh/yr) 2.10 4.30 6.50 8.60 10.80 13.00 15.10 17.30 19.40 21.60 23.80 25.90
Electricity to Grid (GWh/yr) 0.00 0.01 0.02 0.03 0.04 0.04 0.05 0.06 0.07 0.07 0.08 0.09
Techno-Economic
Private-Funded
ROI (%) -557 -508 -465 -443 -416 -393 -382 -364 -349 -342 -328 -324
PP (years) >15 >15 >15 >15 >15 >15 >15 >15 >15 >15 >15 10
LCOE (US$/kWh) 0.39 0.28 0.22 0.20 0.19 0.16 0.16 0.15 0.14 0.14 0.13 0.13
Self-Funded
ROI (%) -249 -185 -130 -101 -67 -37 -22 -0.3 20 29 47 52
PP (years) >15 >15 >15 >15 12.00 9.70 8.90 8.00 7.30 7.00 6.50 6.40
LCOE (US$/kWh) 0.13 0.09 0.07 0.07 0.06 0.05 0.05 0.03 0.03 0.04 0.04 0.04

114
Table 5.11: Feasibility Assessment of On-Site Biogas-based CHP Plant

Plant Size (MW) 0.04 0.08 0.14 0.18 0.23 0.28 0.33 0.37 0.41 0.46 0.50 0.55
Technical
Electricity Generated (GWh/yr) 0.32 0.64 1.10 1.40 1.84 2.24 2.64 2.96 3.30 3.60 4.00 4.40
Parasitic Load (GWh/yr) 0.04 0.07 0.12 0.15 0.20 0.25 0.29 0.33 0.36 0.40 0.44 0.48
Mill Electricity Demand (GWh/yr) 2.20 4.30 6.50 8.60 10.80 13.00 15.10 17.30 19.40 21.60 23.80 25.90
Electricity to Grid (GWh/yr) 0 0 0 0 0 0 0 0 0 0 0 0
Steam fed to the Mills (k t/yr) 1.92 4.00 5.80 7.60 9.60 11.40 13.60 15.20 17.20 19.20 20.80 22.80
Techno-Economic
Private-Funded
ROI (%) -625 -567 -524 -491 -457 -430 -400 -379 -356 -334 -318 -300
PP (years) >15 >15 >15 >15 >15 >15 >15 >15 >15 >15 >15 10
LCOE (US$/kWh) 0.260 0.142 0.106 0.084 0.072 0.064 0.058 0.053 0.049 0.046 0.043 0.041
Self-Funded
ROI (%) -337 -249 -206 -164 -120 -85 -45 -19 -10 24 60 85
PP (years) >15 >15 >15 >15 >15 13.80 10.20 8.80 7.60 7.00 6.30 5.80
LCOE (US$/kWh) 0.085 0.042 0.035 0.028 0.024 0.021 0.018 0.017 0.016 0.015 0.014 0.013

115
To the best of the author’s knowledge, no other study has presented and compared the technical
and techno-economic performance of the EFB and biogas for electricity and heat generation in
Malaysia. The results presented led to the remarkable performance comparisons of both
residues in Table 5.12. The tabulated comparison revealed explicit evidence to distinguish the
suitable residues for electricity and heat generation. The EFB is more suitable for electricity
generation with higher technical and techno-economic feasibility. This confirms our earlier
findings that reckon higher electricity generation capacity from EFB (Md Jaye, Sadhukhan and
Murphy, 2018). Table 5.12 also provides a performance comparison between the power plant
and the CHP plant. This comparison allows the recommendation for appropriate installation
scheme for POMR-RES to be made.

These comparisons, however, have not confirmed the previous research on the feasibility of
generating electricity from palm oil mill residues. Therefore, it is served as a foundation to
strategize appropriate courses of action to translate the electricity generation potential from the
suitable residues to its actual implementation. More broadly, this will eventually become a
useful aid that leads to a better uptake of the technology in the future and for a more sustainable
utilization of POMR.

Table 5.12: Summary of POMR-RES Performance

EFB Biogas
Assessment Power Plant CHP Power Plant CHP
Technical √√ √√ √ ≠
Techno-Economics
Private-Funded ≠ √√ ≠ √
Self-Funded √√ √√ √ √

Note:
√√ - highly feasible
√ - feasible
≠ - not feasible

5.5.1.3 Sensitivity Analysis

A sensitivity analysis was conducted for the EFB-based power plant by focusing on its techno-
economic performance. It was widely claimed that the techno-economic performance of power
plant is generally sensitive towards the capital cost and the interest rate imposed by the financial
institution (ACE, 2016). Arena, Gregorio and Santonastasi (2010) had acknowledged that the
techno-economic performance of bioenergy system is affected by these two cost parameters.
116
In this sensitivity analysis, the changes in the techno-economic performance of the power plant
was observed by, first, modulating the interest rate to 8% and 4% as suggested by Yeoh (2004)
and Clean Development Mechanism (2006); second, excluding the new grid installation cost
in the capital cost; third, fixing the capital cost for one-megawatt installation at USD 1.85
million; fourth, increasing the FiT rate by 20% as suggested by Maulud and Saidi (2012).

The sensitivity analysis of the ROI value was presented first followed by the NPV value. The
variation in the ROI value was observed by alternately changing the interest rate from 12% to
8% and 4% in the techno-economic spreadsheet in Appendix C-2. The variation in the ROI
value for the three different interest rates is presented in Figure 5.2. The graph shows that there
is a steady increase in the ROI value, and the power plant is more likely to become
economically feasible when the interest rate was lowered to 4%. At this interest rate setting
(4%), the power plant starts to produce a positive ROI when the capacity reaches 4.80 MW.
For the power plant to operate at this capacity, it has to be installed at medium scale mills (i.e.,
mills processing capacity of 50 t/hr and above). Note that the larger scale systems have higher
likeliness in techno-economic performance due to higher energy density and the effect of the
economies of scale (Searcy and Flynn, 2009; IRENA, 2012; McIlveen-Wright et al., 2013;
Codina Gironès et al., 2017).

12% 8% 4%

200

100

0
0.9 1.8 2.7 3.6 4.4 5.3 6.2 7.1 8 8.9 9.8 10.7
-100
ROI (%)

-200

-300

-400

-500

-600
PLANT SIZE (MW)

Figure 5.2: Impact of Interest Rate Variation on ROI of on-site POMR-RES intallation
Next, it was observed that the ROI shifted to slightly higher values when the cost of the new
grid installation was excluded in the cost estimation for the power plant. A significant shift was
observed for small- and medium-scale mills before plateaued at large-scale mills as shown in
Figure 5.3.

117
With Grid Cost Without Grid Cost

0
-50 0.9 1.8 2.7 3.6 4.4 5.3 6.2 7.1 8 8.9 9.8 10.7

-100
-150
-200
ROI (%)

-250
-300
-350
-400
-450
-500
PLANT SIZE (MW)

Figure 5.3: Impact of Exclusion of New Grid Installation Cost on ROI of on-site POMR-RES
intallation
Subsequently, the graph in Figure 5.4 illustrates the sensitivity analysis results when the capital
cost per MW installation was reduced to US$ 1.85 million. An improvement in the ROI value
for all generation plants was observed; however, it still failed to surpass a 0% ROI threshold
point. From the graph, the large-scale plants show greater ROI improvement due to economies
of scale.

US$ 2.3 mil US$ 1.85 mil

0
-50 0.9 1.8 2.7 3.6 4.4 5.3 6.2 7.1 8 8.9 9.8 10.7

-100
-150
-200
ROI (%)

-250
-300
-350
-400
-450
-500
PLANT SIZE (MW)

Figure 5.4: Impact of Cost Variation on ROI of on-site POMR-RES intallation


Lastly, the sensitivity analysis was conducted by increasing the FiT rate by 20% as suggested
by Maulud and Saidi (2012). Increasing the FiT rate has altered the ROI values with an average
improvement of 27%.

118
From results of sensitivity analysis, there was a noticeable improvement in the ROI value with
any changes in the sensitivity parameters used in this subsection. The results show that
modulating the interest rate is the most influential sensitivity parameter to improve the techno-
economic performance of the POMR-RES power plant, followed by fixing a lower capital cost
required to install the power plant. It has been proved that the adjustment of the FiT rate and
exclusion of new grid installation cost resulted in minor techno-economic performance
improvement. From this sensitivity analysis, it can be concluded that only medium- to large-
scale power plants with an electricity generation capacity above 4.80 MW can produce a
positive ROI when the interest rate is adjusted to 4%.

The sensitivity analysis also found that the choice either to connect all of the electricity
generated to the grid or only the surplus electricity after fulfilling the mill demands does not
demonstrate a significant difference in the ROI value of the power (Figure 5.5). This
insignificant ROI value difference was due to marginal price difference between the FiT rate
offered to the electricity generated from biomass (US$ 0.11/kWh) and the grid electricity price
(US$ 0.10/kWh).

12% Surplus 12% All 4% Surplus 4% All

200
100
0
0.9 1.8 2.7 3.6 4.4 5.3 6.2 7.1 8 8.9 9.8 10.7
-100
ROI (%)

-200
-300
-400
-500
-600
PLANT SIZE (MW)

Figure 5.5: ROI Value Comparison between Connecting All Electricity and Surplus
Electricity to Grid

Altogether, the results from sensitivity analysis shown that there is a slight decrement on the
ROI value for 9.8 MW power plant to 10.7 MW power plant (i.e. Figure 5.2 to Figure 5.5).
This condition has resulted from the reduction in the FiT rate offered to the power plant which
size above 10 MW (i.e., from US$ 0.1045/kWh to US$ 0.1029/kWh in Table 5.5). It is also
observed that the graphs in Figure 5.2 to Figure 5.5 are not linear showing that the techno-
economic performance of the power plant is sensitive to the changes in the plant size.

119
Differently, for DCF, the sensitivity analysis indicated that the medium-scale power plant
produces positive NPV when the discount rate is reduced to 8% as shown in Figure 5.6. The
non-linear graph in Figure 5.6 illustrates that the techno-economic performance of the power
plant is sensitive to the time value of money.

12% 8% 4%

40
35
30
25
NPV (MILL US$)

20
15
10
5
0
-5 0.9 1.8 2.7 3.6 4.4 5.3 6.2 7.1 8 8.9 9.8 10.7
-10
-15
POWER PLANT SIZE (MW)

Figure 5.6: Impact of Interest Rate Variation on NPV of on-site POMR-RES installation

Furthermore, the power plant provides higher NPV value when the cost to install the new grid
connection was excluded from the overall cost of the power plant. The exclusion of the new
grid connection has a substantial impact on the NPV value for small- and medium scale power
plant. Additionally, the power plant starts to produce positive NPV at the scale of 8.0 MW
when the CAPEX is reduced to US$ 1.85 million/MW installation. On the other hand, the
power plant starts to produce positive NPV at the scale of 9.8 MW as the FiT rate was increased
by 20% following the suggestion from Maulud & Saidi (2012).

5.5.2 OffS case study: Off-site POMR-RES installation

This case study assumed that the POMR-RES was installed at a location away from the mills.
For this type of installation, only the EFB-based power plant with a similar capacity as the on-
site power plant was considered with an assumption that there is no heat and electricity demand
from the mills. Therefore, the electricity generated is dedicated to supplying to the grid. The
results show that the off-site power plant is technically feasible with the amount of available
electricity for grid connection increased by 25-30%.

120
The techno-economic performance of the power plant was assessed following the techno-
economic spreadsheet in Appendix C-2. The cost to transport EFB to the proposed plant
location was included as the variable operating cost. It was assumed that semi-dry EFB was
transported to the proposed plant location using ten-tonne trucks after being mechanically dried
in the mills. Techno-economic assessment disclosed that presence of transportation cost further
lowered the plant’s techno-economic performance. The ROI has been estimated between -
492% and -269% for the private-funded power plant and -174% and -98% for the self-funded
power plant. The PP for the private-funded power plants exceeds 15 years while a self-funded
power plant has a PP from 5 to 11 years, depending on the power plant scale.

From these techno-economic assessment results, although there is additional revenue from
selling extra electricity to the grid, it was insufficient to offset the feedstock and transportation
costs. This loss has impacted the techno-economic performance of the power plant. This
situation reflects the arguments by Searcy et al. (2007), Leboreiro and Hilaly (2011) and Lewis
(2018) on the effect of the feedstock transportation cost towards the economic feasibility of the
bioenergy systems. The adverse effect from the feedstock costs on the profitability of the power
plant found in this study is in favour of the verdict by Cameron, Kumar and Flynn (2007).
Cameron, Kumar and Flynn (2007) claimed that, the feedstock cost, which is a combination of
cost to buy and transport feedstock is the bottleneck to the development of bioenergy system.
Similarly, Wan et al. (2016) showed that the variation in the feedstock cost had prolonged the
PP of sago-based energy generation. The techno-economic assessment results obtained in this
study confirm that the techno-economic performance of the power plant is sensitive to the
feedstock transportation cost. This finding is consistent with the findings reported by Deng and
Parajuli (2016), Alex Marvin et al. (2012) and Kazi et al. (2010). Additionally, the excess
feedstock cost has subsequently increased the LCOE of power plant, making it non-competitive
compared to the FiT rates offered. This finding underlines the needs to minimize the feedstock
cost to enhance the techno-economic performance of the power plant.

From a simulation study of the effect of feedstock cost on the techno-economic performance
of bioenergy plant conducted by Rodriguez et al. (2011), it is apparent that the renewable
energy certificates (RECs) were capable of reconciling the impact of the feedstock cost on the
techno-economic performance of the bioenergy generation system. This recommendation
sounds sensible since the revenue from the RECs were able to counterbalance the plant cost.
However, the author offers no guarantee of the RECs trading and would appear to be over-
ambitious in their recommendation.

121
The feedstock transportation cost (USD 0.14 million to USD 1.66 million) also has an impact
on the DCF of the off-site power plant. The NPV with and without feedstock transportation
cost has been estimated at -US$ 15.3 million and -US$ 8.0 million for the private-funded power
plant and -US$ 4.1 million and US$ 28.2 million for the self-funded power plant, respectively.
The BEP for the private-funded power plants exceeds 15 years and BEP for self-funded power
plant varies from 14.5 years and 4.5 years, respectively.

5.5.2.1 Sensitivity Analysis

Sensitivity analysis was conducted on the techno-economic performance of the off-site


installation. Similar sensitive variables were used to examine the changes in the techno-
economic performance as in subsection 5.5.1.3. Firstly, by modulating the interest rate, the
graph in Figure 5.7 shows gradual ROI value increment. The positive ROI was produced when
interest rate was reduced to 4% which only a large-scale power plant with a capacity above
8.10 MW was considered as techno-economically feasible. A minor reduction in ROI was
observed for a plant with 10.70 MW capacity following the reduction in FiT rate offered as the
plant size exceeded 10.00 MW.

12% 8% 4%

100

0
0.9 1.8 2.7 3.6 4.4 5.3 6.2 7.1 8 8.9 9.8 10.7
-100

-200
ROI (%)

-300

-400

-500

-600
PLANT SIZE (MW)

Figure 5.7: Impact of Interest Rate Variation on ROI of Off-site POMR-RES intallation

Next, the sensitivity analysis result from excluding the grid installation cost and capping the
capital cost to US$ 1.85 million/MW yielded moderate changes in the ROI (range 8%-10%)
for small-scale plant and an exceptional ROI improvement between 15%-30% for medium to
large-scale plant. Next, the sensitivity analysis result from introducing higher FiT rate as
recommended by Maulud and Saidi (2012) revealed a noticeable improvement on the ROI

122
value primarily for medium- and large-scale plants. Despite the increases in the ROI value,
adjusting the FiT rates may result in restructuring the current electricity tariff which requires
the users to pay higher electricity price (Umar, Jennings and Urmee, 2014a; Shahmohammadi
et al., 2015; Theo et al., 2017).

Furthermore, the positive NPV value is obtained as the discounted rate is reduced to 4% for
4.4 MW power plant. The exclusion of grid installation cost, lowering the capital cost to US$
1.85 million/MW from 2.30 million/MW and increasing the current FiT rate by 20% have
provided positive NPV value mainly for medium- and large-scale power plants.

5.5.3 Co-op case study: Co-operative POMR-RES Installation

The idea behind the co-operative POMR-RES installation is to reckon the concept of co-
location of power generators and promote the resource and technology sharing concept
between mills in proximity to each other (Osmani et al., 2013). The co-operative installation
allows the high electricity generation cost to be shared between the mills and reduces the risk
from an inconsistent residue supply (Cattani and Schmidt, 2005; Tan et al., 2016). In this case
study, two scenarios using EFB were assessed to justify the technical and techno-economic
feasibility of the co-operative installation. The two scenarios reflect the previous two case
studies where the co-operative EFB-based power plant was installed, first, in the compound of
one of the mills that agreed to participate in the co-operative generation (on-site co-operative
installation); and second, at a location outside of the mills (off-site co-operative installation).
The interest rate is set at 4% guided by the techno-economic assessments in the previous two
case studies (OnS case study and OffS case study).

5.5.3.1 On-site Co-operative POMR-RES Installation

The on-site co-operative installation can be described as the installation of the power plant in
one of the mills that agreed to participate in the co-operative electricity generation. This mill
acts as the technology hub while the other mills in the association act as the EFB supplier to
support a continuous electricity generation process. Two samples of on-site co-operative
installation were assessed to represent its technical and techno-economic performance. First,
the co-operative association comprises three small-scale mills (one mill with 30 t/hr and two
mills with 10t/hr FFB processing capacities) and power plant is present in the bigger mill.
Secondly, the power plant installation is considered in the medium-scale mill and has the EFB
supplies from one small-scale mill (one mill with 50 t/hr and one mill with 30t/hr FFB

123
processing capacities). The transportation distance is kept at 10 km, and the EFB can be
transported using ten-tonne trucks.

The technical and techno-economic performances of the first sample of co-operative electricity
generation between three small-scale mills (30-10-10) are presented in Table 5.13.

Table 5.13: Technical and Techno-Economic Performance of On-Site Co-Operative


Installation (Co-op 1)
Technical
Plant Size (MW) 4.40
Electricity Generated (GWh/yr) 35.60
Parasitic Load (GWh/yr) 3.90
Mill Electricity Demand (GWh) 6.50
Electricity to Grid (GWh/yr) 25.20
Techno-Economic
Private-Funded
ROI (%) -38
PP (years) 9.40
LCOE (US$/kWh) 0.07

Self-Funded 63
ROI (%) 6.10
PP (years) 0.05
LCOE (US$/kWh)

Results in Table 5.13 show that with additional EFB from two 10 t/hr mills, the electricity
generation capacity of the 30 t/hr mill increases from 2.70 MW to 4.40 MW. The increment in
electricity generation capacity indicates that the co-operative installation generates more
electricity with extra surplus electricity available for the grid. With an additional transportation
cost of US$ 0.21 million for the semi-dry EFB, the ROI value for the private-funded installation
remains below the zero percent ROI threshold line (-38%). Interestingly, this ROI value is
higher compared to the individual on-site power plant installation, indicating that the co-
operative installation has a better techno-economic performance than the individual
installations.

These convincing technical and techno-economic performances of the Co-op 1 installation can
be further explained using results shown in Table 5.14. The technical performance comparison
between the individual installations of 30 t/hr, 10 t/hr and 50 t/hr with Co-op 1 highlights that
the co-operative installation has extra surplus electricity to be connected to the grid. For
instance, although having the same 4.40 MW electricity generation capacity, Co-op 1 power

124
plant can supply 17% more electricity to the grid as compared to the plant installed at 50 t/hr
mills. This is due to the higher electricity demand from the 50 t/hr mills which then limits the
amount of electricity to be connected to the grid.

Table 5.14: Technical and Techno-economic Performance Comparison between Individual


Installations and Co-op 1 Installation
Mill Processing Capacity (t/hr) 30 10 10 50 Co-op 1
Technical
Plant Size (MW) 2.70 0.90 0.90 4.40 4.40
Electricity Generated (GWh/yr) 21.40 7.10 7.10 35.60 35.60
Parasitic Load (GWh/yr) 2.40 0.80 0.80 3.90 3.90
Mill Electricity Demand (GWh/y) 6.50 2.10 2.10 10.80 6.50
Electricity to Grid (GWh/yr) 12.50 0.80 0.80 20.90 25.20
Techno-Economic
Private-Funded
ROI (%) -89 -231 -231 -11 -38
PP (years) >15 >15 >15 8.20 9.40
LCOE (US$/kWh) 0.07 0.09 0.09 0.05 0.07
Self-Funded
ROI (%) 10 -140 -140 93 63
PP (years) 7.50 >15 >15 5.10 6.10
LCOE (US$/kWh) 0.04 0.06 0.06 0.03 0.05

Besides the effects of economies of scale, the significant techno-economic performance


improvement presented in Table 5.14 is closely linked to several contributing factors. These
include the reduction in the interest rate payments, avoidance of grid installation and
administrative costs, and reduction in the annual operating cost as summarized in Table 5.15.

Table 5.15: Costing Comparison of the Individual Installations with Co-op Installation
Plant Size (t/hr) 30 10 10 30+10+10 50 Co-op 1
Total Capital Cost 6.30 3.30 3.30 12.90 8.70 8.70
(million US$)
Total Operating Cost 0.60 0.30 0.30 1.10 0.90 1.10
(million US$)
Interest Rate Payment 3.80 2.00 2.00 7.80 5.20 5.20
(million US$)
Annual Cost 1.50 0.80 0.80 3.10 2.10 2.30
(million US$)

From the summary table, there is a reduction by 33% in the total capital cost of the Co-op 1
installation as compared to the cost of the three individual installations. Similarly, the amount

125
of interest rate repayment is reduced by 32%. Approximately 26% reduction in the annual
operating cost can be observed for the Co-op 1 installation even though an additional
transportation cost has been considered. Likewise, supplying 17% more electricity to the grid
as compared to the plant installed at 50 t/hr mills can contribute to the positive shift in the
investment returns. With this defence, it recognizes the accountability of the co-operative
installation concept to be considered as an alternative to the individual power plant
installations, especially for small-scale mills.

Subsequently, Table 5.16 presents the feasibility assessment outcomes for the installation of
the electricity generation plant in the 50 t/hr mill receiving additional 27,000 tonnes of EFB
supply (Co-op 2). As expected, the results confirm that the technical feasibility of the Co-op 2
installation with 56.40 GWh electricity generated annually. The Co-op 2 installation produces
an ROI of 24% and will break even at approximately the seven-year mark. This indicates that
sharing the resources and technology between a medium- and a small-scale mill results in the
improvement of the power plant technical and techno-economic performances.

Table 5.16: Technical and Techno-Economic Performance of On-Site Co-Operative


Installation (Co-op 2)
Technical
Plant Size (MW) 7.0
Electricity Generated (GWh/yr) 56.40
Parasitic Load (GWh/yr) 6.20
Mill Electricity Demand (GWh/yr) 10.80
Electricity to Grid (GWh/yr) 39.40
Techno-Economic
Private-Funded
ROI (%) 24
PP (years) 6.80
LCOE (US$/kWh) 0.052
Self-Funded
ROI (%) 129
PP (years) 5.00
LCOE (US$/kWh) 0.041

The technical and techno-economic performance comparison between the individual


installations and the Co-op 2 installation is further synthesized and presented in Table 5.17.
The table precisely appraises the technical feasibility of the Co-op 2 installation which has
more surplus electricity for the grid connection in comparison to the individual installations.

126
Besides the technical performance, the information in the table instils an important argument
on the techno-economic performance of the cooperative installation. A 24% ROI is projected
from the Co-op 2 installation while individual power plant installation in 50 t/hr and 30 t/hr
have provided negative ROIs previously.

This techno-economic performance improvement indicates that the co-operative generation


could be the realistic option to promote and encourage the participation of small-scale mills in
generating electricity from their available EFB. However, the ROI value for the Co-op 2
installation is lower than the individual on-site installation of 7.10 MW, reflecting the influence
of the feedstock transportation cost on the profitability of the power plant.

Table 5.17: Technical and Techno-economic Performance Comparison between Individual


Installations and Co-op 2 Installation
Mill Processing Capacity (t/hr) 50 30 80 Co-op 2
Technical
Plant Size (MW) 4.40 2.70 7.10 7.0
Electricity Generated (GWh/yr) 35.60 21.40 57.00 56.40
Parasitic Load (GWh/yr) 3.90 2.40 6.30 6.20
5Mill Electricity Demand 10.80 6.50 17.30 10.80
(GWh/y) 20.90 12.50 33.40 39.40
Electricity to Grid (GWh/yr)
Techno-Economic
Private-Funded
ROI (%) -11 -89 53 24
PP (years) 8.20 14.30 6.20 6.80
LCOE (US$/kWh) 0.046 0.055 0.040 0.052
Self-Funded
ROI (%) 93 10 163 129
PP (years) 5.10 7.50 4.70 5.00
LCOE (US$/kWh) 0.031 0.038 0.033 0.041

Detailed DCF analysis for the cooperative installation has also provided a valuable result that
acknowledged the techno-economic benefits of such installation. The NPV of the cooperative
installation is projected to be higher than the individual installations.

Despite these encouraging results, cooperative installation is still fairly undefined. Currently,
there is an inadequate discussion made for the co-operative POMR-RES installation, and the
actual feasibility has not yet been established. Nevertheless, the features of the results achieved
in this chapter corroborates well with the resource pooling principle reported by Cattani and
Schmit (2005). This finding, nonetheless, sheds new lights on the installation of the POMR-

127
RES, especially for small-scale mills which previously did not confirm the techno-economic
feasibility requirement in terms of a positive ROI.

5.5.3.2 Off-site Co-operative POMR-RES Installation

This subsection presents the feasibility assessment outcomes for the off-site co-operative
POMR-RES installation. This kind of installation required the EFB to be transported to the
proposed power plant location. The transportation distance is kept at 10 km to simplify the cost
estimation process. The electricity generation potential of the off-site installation is identical to
the on-site installation with no electricity demand from the mills. It is ironic that the cost to
transport the EFB has adversely affected the techno-economic performance of the power plant
(Hamelinck, Suurs and Faaij, 2005; Silva Herran and Nakata, 2012; Reeb et al., 2014). The
ROI of the 4.40 MW and 7.00 MW was estimated at -78% and -23% while the NPV for the
similar power plant was estimated at -US$ 6.5 million and -US$ 15 million, respectively. The
negative ROI and NPV indicated that this type of installation is techno-economically
unfeasible. These results accord well with our earlier findings for the off-site POMR-RES
installation in subsection 5.5.2 in which the power plant experienced a techno-economic
performance declination due to the additional liability for transportation cost (Jenkins, 1997;
Sultana, Kumar and Harfield, 2010; Miao et al., 2012). The transportation costs can also lower
the ROI and NPV of the self-funded power plants.

5.5.3.3 Sensitivity Analysis

The sensitivity analysis results indicated that the on-site installation becomes economically
viable when the new grid’s installation cost is removed from the investor’s onus, the capital
cost is reduced to US$ 1.85 million/MW, and the FiT rate is increased to US$ 0.13/kWh. The
PP of the power plant is expected between the fourth and seventh year. This PP has been found
to agree with the findings in the literature (Bakos, Tsioliaridou and Potolias, 2008; Wood and
Rowley, 2011; Van Dael et al., 2013) and with our initial findings on the on-site installation.
As expected, the techno-economic performance of the larger scale installation (7.00 MW) is
more intensified with the flexibility in reducing the costs and increasing the revenue. Similarly,
the techno-economic performance of the self-funded project continues to surge up with the ROI
values over 100% for 4.40 MW and 7.00 MW, respectively. The PP for both installations is
between four to five years.

128
For the off-site co-operative installation, variations in the sensitivity analysis parameters have
shown improvement trends in the techno-economic performance. The private-funded 4.40 MW
installation is considered techno-economically feasible when the FiT rate is adjusted to US$
0.13/kWh while the other two options have failed to provide positive ROIs. An improvement
in the techno-economic performance is observed for the self-funded 4.40 MW installation with
a variation in any sensitivity parameters. Subsequently, the sensitivity analysis results for the
7.00 MW installation reveal that the exception of the new grid installation cost, reduction of
the capital cost for each MW installation to US$ 1.85 million and increment of the FiT rate by
20% are able to offset the EFB transportation cost, thereby providing an encouraging ROI and
and a shorter PP.

The findings on the co-operative installation, while preliminary, have proven the technical and
techno-economic feasibility of generating electricity by combining the available EFB from few
small-scale mills. It can, therefore, be assumed that the on-site co-operative installation is
preferable to minimize the effect of transportation cost. Nonetheless, this kind of installation
may require further intervention and support from the government since it demands more than
lowering the interest rate to 4% from 12%. A particular business partnership model is essential
before the association of a co-operative power generation is established (Umar, Urmee and
Jennings, 2017). Further studies with more focus on this kind of installation are, therefore,
recommended. Amongst the suggested areas that require further studies include the
development of cost and benefits allocation model to allow a fair cost and benefit distribution
amongst the participants, and the establishment of an appropriate mechanism to determine the
ownership of the plant. Holistic studies to assess the advantages and disadvantages of the co-
operative installation are also essential. The combination of findings from this study with the
proposed studies provide substantial support to develop the conceptual model for the co-
operative POMR-RES installation as prompted by Umar, Jennings and Urmee. (2014b) and
further emphasized in Umar, Urmee and Jennings (2017). In recent years, there has been a
growing interest among the researchers in exploring the actual feasibility of co-operative
electricity generation and formulating the optimum strategy to establish the co-operative
association for electricity generation (Maali, 2009; Cheng, Chang and Jiang, 2014; Tan et al.,
2016; Song et al., 2017).

129
5.6 Technical and Techno-Economic Assessment of Dried EFB-based Power Plant

The technical and techno-economic performances of the power plant using the dried EFB are
assessed in this section. Both of the assessments were conducted following the methodology
discussed in subsections 5.3.1 and 5.3.2. The technical performance of the dried EFB-based
power plant was assessed corresponding to the technical assessment spreadsheet using the
estimated amount of electricity generated that was identified in Chapter 4 (Tables 4.15 and
4.16). This assessment confirmed the technical feasibility of the power plant. Next, the cost of
drying the EFB was included in the techno-economic performance assessment of the power
plant. The drying cost comprises its capital and operating costs. The cost of the heater and other
accessory equipment is neglected at this point of estimation. According to Pang and Mujumdar
(2010) and Luk et al. (2013), the dryer cost is estimated based on the heat demand of the drying
process or the physical size of the dryer. The size of the dryer is measured depending on the
volume of the EFB inside the drier (Gebreegziabher, Oyedun and Hui, 2013). In this chapter,
the dryer cost (𝐶𝑜𝑠𝑡𝐷 ) is estimated based on the physical size of the dryer using Eq. (17) and
Eq. (18) as follows:
𝐸𝐹𝐵𝐷
𝑉𝐸𝐹𝐵 = [( × 𝑂𝑝𝑒𝑟𝑎𝑡𝑖𝑜𝑛 𝑇𝑖𝑚𝑒) /3600] × 0.15 (17)
𝜌𝐸𝐹𝐵
𝐶𝑜𝑠𝑡𝐷 = 𝐸𝐹𝐵𝐷 + (𝐸𝐹𝐵𝐷 × 𝑉𝐸𝐹𝐵 0.6 ) (18)
where 𝑉𝐸𝐹𝐵 is the volume of the EFB in the dryer in m3, 𝐸𝐹𝐵𝐷 is the amount of the EFB
required drying in t/hr, 𝜌𝐸𝐹𝐵 is the density of the EFB quoted as 70 kg/m3 (Tenorio and Moya,
2012) and the 𝑂𝑝𝑒𝑟𝑎𝑡𝑖𝑜𝑛 𝑇𝑖𝑚𝑒 is fixed at 8000 hours annually. The volume of the EFB in the
dryer is assumed to represent approximately 15% from the overall dryer volume while the
remaining volume is represented by the free space for the hot air to flow during the drying
process (Gebreegziabher, Oyedun and Hui, 2013). The annual operating cost of the drying
process (𝐶𝑜𝑠𝑡𝑂𝐷 ) is computed using Eq. (19) :
(𝐶𝑜𝑠𝑡𝑆 × 𝑄𝑑𝑟𝑦𝑒𝑟 × 𝑂𝑝𝑒𝑟𝑎𝑡𝑖𝑜𝑛 𝑇𝑖𝑚𝑒)
𝐶𝑜𝑠𝑡𝑂𝐷 = (19)
3600
where 𝑄𝑑𝑟𝑦𝑒𝑟 is the heat requirement for the drying process in MW and 𝐶𝑜𝑠𝑡𝑆 is the steam
cost quoted at US$ 100 for every MW of heat demand (Gebreegziabher, Oyedun and Hui,
2013). Here, the heat demand for the drying process obtained in Section 4.8.1 is used to
estimate the operating cost.

130
5.6.1 Cost for Drying the EFB

Table 5.18 presents the estimated dryer cost and operating cost for drying the EFB.

Table 5.18: Estimated Cost for Drying the EFB

𝑄𝑑𝑟𝑦𝑒𝑟 (MW) 𝐶𝑜𝑠𝑡𝐷 (mill US$) 𝐶𝑜𝑠𝑡𝑂𝐷 ( US$/yr)


0.60 0.05 133.30
1.30 0.14 244.40
1.80 0.26 355.60
2.30 0.41 466.70
2.80 0.58 555.60
3.20 0.78 644.40
3.70 1.00 733.30
4.00 1.24 822.20
4.40 1.49 888.90
4.70 1.76 955.60
5.00 2.05 1022.20
5.30 2.36 1088.90

The estimated dryer cost increases from US$ 0.05 million to 2.36 million while the annual
operating cost increases from US$ 133.30 to US$ 1088.90 as the heat demand for drying the
EFB increases. This estimation is consistent with the estimates made by Gebreegziabher,
Oyedun and Hui (2013). This estimation is found to be much higher than the cost reported by
Wan et al. (2016) and far lower than the cost claimed by Haque and Somerville (2013). Note
that the estimated cost has been made based on the theoretical calculations and has not been
verified with the retail dryer cost.
The additional cost of drying raises the electricity generation cost and has negatively impacted
the ROI of the power plant. The ROI of the private-funded power plant varies between -464%
and -244% and between -129% and 155% for the self-funded power plant, respectively. The
PP of private-funded power plant exceeds 15 years while the self-funded power plant has a PP
between 5 and 10 years. These techno-economic performance downturns indicates that the
proposal of installing the biomass dryer requires further consideration as it increases the capital
cost burden to the investors (Ståhl et al., 2004; Rajendran, 2017). However, the installation of
the biomass dryer is sometimes technically unavoidable. Therefore, it must be assessed beyond
the techno-economic concern, more from the overall system operational and environmental
concerns (Brammer and Bridgwater, 2002; Fagernäs et al., 2010). The similar argument has
been discussed earlier in Section 4.8.1 of this thesis.

131
5.7 Summary of Findings

The technical and techno-economic feasibility assessments of the new installation of EFB- and
biogas-based power plant and CHP plant were presented in this chapter. Three case studies,
namely the on-site, off-site and co-operative plant installations were used to illustrate the
feasibility assessment results. Through the technical feasibility assessment that was carried out,
firstly, EFB was found as having a better potential to be used for electricity generation and
secondly, extracting the exiting steam from the BPST in the CHP plant has significantly
reduced the electricity generation capacity. The steam extracted from the BPST supply about
5%-35% of the heat demand from the mills. It can be concluded that the utilization of EFB is
specified for electricity generation while the practicality of CHP plant installation should be
re-evaluated since most of the mills are considered as heat and electricity self-sufficient by
combusting the MF and PKS as discussed in Section 3.1.2. To maximize the electricity
generation and, at the same time, to enhance the utilization of the biogas, an alternative POMR-
RES energy system with a heat-only biomass boiler is proposed as shown in Figure 5.7. In this
alternative energy system, EFB and biogas are combusted in two separate boilers. EFB is
combusted and the heat is recovered into high pressure superheated steam which is then
expended to generate electricity. Biogas can generate medium pressure steam to cook the FFB.
The proposed energy system will maximize the utilization of EFB and biogas and release the
MF and PKS for other value-added applications such as light-weight concrete, fillers, and
others since both of these residues have a higher existing and future market demand with a
better market price. However, further technical design validation is required to accurately
confirm the technical details of the proposed energy plant. Technically, approximately 0.75 -
8.95 MW of heat is extractable from combusting the biogas. This estimated extractable heat is
sufficient to dry the EFB using the rotary dryer as outlined in Table 4.14.

132
Figure 5.7: POMR-RES Energy System

Next, the techno-economic feasibility assessment shows that the techno-economic performance
of the private-funded projects is very vulnerable towards the interest rate charged by the
financial institution. The sensitivity analyses in Section 5.5.1.3, 5.5.2.1 and 5.5.3.2 show that
the current interest rate of 12% is too high to empower the electricity generation from EFB and
biogas. These sensitivity analyses recommend a 4% interest rate as the ceiling rate to spearhead
the development of POMR-RES in the mainstream. In comparison, the self-funded projects are
more economically competitive as they are free from the interest rate liabilities.

With the 4% interest rate assumption, positive return on investment is in favour of on-site
POMR-RES installation which is currently limited to medium- and large-scale plants. The off-
site POMR-RES installation, on the other hand, is only applicable to the large-scale plants
when the cost to transport the EFB was included in the costing estimation. It became infeasible
when the EFB cost was added on top of the transportation cost. The on-site co-operative
installation exhibits an exceptional techno-economic performance; however, further studies
related to the installation ownership and cost and revenue allocation among the participants are
highly demanded.

The sensitivity analysis also suggested that capping the installation cost of one-megawatt
electricity generation capacity to US$ 1.85 million, in essence, will improve the techno-
economic performance of POMR-RES. However, it is subjected to technology market price
volatility. Presently, Malaysia is still relying on the foreign technology with an incarnation of

133
few local boiler and turbine manufacturers and assemblers. The other two sensitivity
parameters, which are eliminating the cost of the new grid connection installation and revising
the existing FiT rate, has less impact towards the techno-economic performance of POMR-
RES.

Taken together, the results from this chapter provide insightful evidences concerning the
technical and techno-economic feasibility of POMR-RES. However, Freppaz et al. (2004)
argued the biomass exploitation for bioenergy should go beyond assessing the bioenergy
potential to a greater extent, the formation of the more optimal bioenergy models. Despite its
name, finding the optimal bioenergy plant model does not necessarily provide the optimum
solution in terms of the overall complex structure of the bioenergy supply chain (Ba, Prins and
Prodhon, 2016; Ghaderi, Pishvaee and Moini, 2016). Finding the optimal bioenergy plant size,
instead, provides feasible solutions for translating the bioenergy generation potential into its
actual implementation (Yue, You and Snyder, 2014). The main innovations offered by the
optimal bioenergy models include better planning of the resources, coordinating the resource
supply chain, improving the economic performance of bioenergy, and emphasizing the
environmental advantages as well as providing the ability to verify the overall performance of
the bioenergy system (Freppaz et al., 2004; Mafakheri and Nasiri, 2014).

Most literature acknowledged the needs for establishing the economically feasible size of the
bioenergy plant along with the detailed technical, economic and environmental feasibility
assessments (Hoeltinger, Schmidt and Schmid, 2012; Steubing et al., 2014; Zema, 2017).
Various approaches were introduced and discussed in the literature including mathematical
calculations, linear and non-linear programming (Baños et al., 2011), and heuristic techniques
(Kim, Sen and Maravelias, 2013). Currently, limited information is available in the literature
that discusses a method to identify the economically feasible size for POMR-RES. Therefore,
in the next sub-section, an attempt has been made to provide comprehensive technical and
techno-economic arguments for generating electricity from POMR by proposing the
economically feasible size of the EFB-based power plant. By operating at this size, the plant is
expected to generate profitable electricity that manifests a favourable ROI.

5.8 Economically Feasible Size of POMR-RES

Generating renewable electricity is still observed from commercial rather than an


environmental protection and social responsibility perspective. Finding the economically

134
feasible size for the POMR-RES in this study was driven by an aim to secure ROI values of
20% from generating electricity using EFB. The techno-economic indicator values of 20% was
used as the objective functions guided by the profitability projection made by Yoshizaki et al.
(2012) and Lam et al. (2013). Yoshizaki et al. (2012) specified a proposal for a new installation
of RE-based electricity generation plant which had projected an ROI lower than 15%, making
it a financially unattractive and not a business-worthy proposal. This statement is in line with
the benchmark set by the United Nation Framework Convention on Climate Change (UNFCC)
for Felda Serting Hilir biogas electricity plant project (Clean Development Mechnism, 2006).
For placing more emphasis, Lam et al. (2013) indicated that the notable decline in interest
among prospective investors is when the projected ROI is less than 15%. However, there has
been little discussion that used NPV as the indicator to establish the economically feasible
power plant size.
The mathematical calculation optimization method prompted by Lauven, Liu and Geldermann
(2013) was used to identify the economically feasible size of POMR-RES by adjusting the
capital cost of the plant to a point where the revenue obtained is sufficient to provide 20% ROI
and positive NPV. The ROI with a 4% annual interest rate and NPV profile with a 8% annual
interest rate of on-site EFB-based power plant as shown in Table 5.19 were used as basis to
determine the economically feasible size for the power plant. Using mathematical calculation,
the power plant with an electricity generation capacity of 5.70 MW and 5.30 MW yields an
investment return of 20% and a positive NPV, respectively. A period of five to seven years is
required to recover the cost of investment. Using the 5.70 MW as the economically feasible
size, the power plant is capable of generating approximately 45.60 GWh of electricity annually
using 59,100 tonnes semi-dried EFB (~76,800 tonnes of raw EFB). Based on the annual
electricity consumption of 4.6 MWh per capita in Malaysia (World Bank, n.d), the amount of
electricity generated can supply electricity to 10,800 consumers.

135
Table 5.19: POMR-RES Costing Profile

Mill Processing POMR-RES Size ROI PP NPV BEP


Capacity (MW) (%) (yr) (mill US$) (yr)
(t/hr)

10 0.90 -231 >15 -6.20 >15

20 1.80 -143 >15 -5.10 >15

30 2.70 -90 14.30 -4.10 10.00

40 3.60 -49 10.20 -2.90 8.00

50 4.40 -12 8.20 -1.00 6.30

60 5.30 13 7.20 0.60 5.50

70 6.20 35 6.60 2.40 5.10

80 7.10 54 6.10 4.30 4.70

90 8.00 71 5.80 6.30 4.40

100 8.90 85 5.50 8.30 4.20

110 9.80 98 5.30 10.50 4.00

120 10.70 96 5.40 10.70 4.10

The economically feasible sizes identified in this subsection provide answers to the third
research objectives that is to propose the economically feasible size for POMR-RES to generate
profitable electricity. From this economically feasible size, the amount of EFB required to
generate profitable electricity is estimated. With this estimation, the millers will be able to self-
assess whether the available EFB in their mill is adequate for generating profitable electricity.
On a broader context, identification of the economically feasible size of POMR-RES in
Peninsular Malaysia gives an indicator for the number of the installable plants and the
attainable electricity generation potential from the EFB (Fiala, Pellizzi and Riva, 1997;
Hoefnagels, Junginger and Faaij, 2012). The economically feasible size may limit the
possibility of overly-excited investment among the prospective investors, and it will allow a
fair RE quota allocation to eligible investors, thus reinforcing the effectiveness and
sustainability of the RE Fund. With these remarkable credits from confirming the economically
feasible size power plant, an explicit alliance involving all the stakeholders is highly demanded.
This alliance is expected to create a business-enabling environment for POMR-RES and attract

136
more future investments in the renewable electricity generation sector. This primarily will lead
to a more sustainable RE market, encourage broader deployment of POMR-RES, and
contribute towards achieving the national RE target.
Since most of the mills (59%) are classified as small-scale, only 35 mills are identified to have
a sufficient EFB supply to generate the profitable electricity with the total accumulated
generation capacity of 200 MW. This accumulated capacity accounts for 25% of the 2020’s
target for palm oil biomass under the National Renewable Energy Policy and Action Plan
(Shamsuddin, 2012; Umar, Jennings and Urmee, 2014a). Note that, ten out of these 35 mills
are capable of generating electricity at a capacity of 8.40 MW.

5.9 Conclusions

The economically feasible size of the EFB-based power plant that is capable of generating a
20% ROI to become economically feasible is identified. The amount of EFB needed to generate
electricity at the economically feasible size is also calculated. This research can be considered
as the first attempt to determine the economically feasible size of the EFB-based power plant
in Peninsular Malaysia. The cost profile of the on-site EFB-based power plant with the annual
interest rate of 4% was used as the basis for identifying the economically feasible size. The
5.70 MW electricity generation capacity is deduced as the economically feasible power plant
size as these power plants provide an ROI of 20%. Approximately 59,100 tonnes and 86,900
tonnes of semi-dried EFB are required annually to generate electricity at the economically
feasible size. There are 35 mills capable of generating electricity at the capacity of 5.70 MW,
bringing the accumulated generation capacity to 200 MW. However, this accumulated capacity
is still far from the target set for palm oil mill biomass. To accelerate the technology
development and foster the policy achievement in a profitable manner, the installation of power
plant must go beyond individual installation. Co-operative installation is seen as one of the
practical solutions to enhance the abilities to generate electricity using EFB among small-scale
mills. The technical and techno-economic feasibility assessment conducted appraised the
higher likelihood of the co-operative installation to generate profitable electricity from EFB.
Nevertheless, a completely new optimization process is required for the co-operative
installation due to the difference in the cost profile between these two types of installation.

137
CHAPTER SIX

ENVIRONMENTAL PROFILE OF POMR-RES

This chapter focuses on the use of Life Cycle Assessment (LCA) to identify the GHG and other
impacts per kilowatt-hour of electricity provided to the national grid from palm oil mill
residues. It provides a detailed description of the goal and scope of the study, data inventory,
environmental impact assessment and interpretation of results.

6.1 Introduction

The performance of renewable electricity generation systems in comparison with conventional


electricity generation using fossil fuels has been assessed in many studies using various
sustainability indicators (AGECC, 2010; IEA, 2015). Common indicators include the
availability of the renewable resources, the efficiency of the electricity generation process,
capital cost, return on investment, greenhouse gas emission, and local social-economic
development (Evans, Strezov and Evans, 2009). Since RE generation has been widely claimed
as not economically competitive, it is necessary to highlight the environmental advantages of
this alternative electricity generation (Verbruggen et al., 2010; IRENA, 2018). Many studies
e.g. Martinez-Hernandez, Campbell and Sadhukhan (2013), Bilgili et al. (2017) etc. have
shown that RE generation is more environmentally friendly than conventional electricity
generation.

Based on the interest in using biomass for electricity generation, the ability of such resources
to reduce GHG emissions has to be thoroughly assessed. According to past works, the
generation of RE from wind, solar, hydro and geothermal does not directly emit any GHGs,
whereas, RE generation from biomass is regarded as ‘carbon-neutral’ (C-neutral) (Perry,
Klemeš and Bulatov, 2008; Zhu, Long and Ort, 2010). It is assumed that the carbon neutrality
of biomass is due to its implicit or explicit sequestration credit (Johnson, 2009); the implicit
sequestration credit considers biomass as C-neutral because of its inherent carbon sinking while
the explicit sequestration deems that the creation of new biomass removes as much GHG as
emitted during its combustion (Rabl et al., 2007). However, recent literature has shown
contradictory claims about the GHG emission from RE generation. Johnson (2009) concluded
that the biomass carbon credits should not be used solely to define the carbon neutrality of RE
from biomass but rather as means to lower net GHG emissions. This conclusion accords with
the findings reported by Rabl et al. (2007), Choo et al. (2011) and Turconi, Boldrin and Astrup

138
(2013): these studies emphasise that every life cycle of the GHG emission of electricity
generation process from biomass starts at the resources acquisition phase and continues through
an explicit accounting of GHG emission at each subsequent stage of the electricity generation
process. This information can be further used to direct future life cycle improvements and as
support for sustainable electricity generation.

As reported by the National Academy of Sciences (2010), electricity generation from


renewable resources still emits GHG, and some of these emissions are significant. The
Environment Agency of UK indicated that the GHG emissions from electricity generated using
biomass are generally, but not always, lower than fossil fuels (Bates, Edberg and Nuttall, 2009).
For instance, electricity generated from short rotation coppice chips produces GHG emissions
which are 35% to 85% less than the combined cycle gas turbine. However, electricity generated
from wheat straw emitted 35% more GHG emissions. A comparison between the GHG
emissions by electricity generation from various renewable resources is presented in Figure 6.1
(a) and (b). The data indicate that while electricity generation from these resources has emitted
a certain amount of GHG, these are substantially lower compared with the combustion of coal
and natural gas. It is worth mentioning that the amount of CO 2eq emissions presented excludes
emission from any land use change.

80
60
40
g CO2eq/kWh

20
0
Hydro Solar Wind Geothermal Biomass
-20
-40
-60
Types of Resources

Min Average Max

Figure 6.1 (a): GHG Emission from Renewable Electricity (National Academy of Sciences,
2010)

139
1200

1000

g CO2eq/kWh
800

600

400

200

0
Coal Natural Gas
Types of Resources

Min Average Max

Figure 6.1 (b): GHG Emission from Fossil-Based Electricity (National Academy of Sciences,
2010)
Lower GHG emission from biomass combustion is also recognized by Balch et al. (2016).
Based on annual global data on emission from fossil fuel and biomass combustion from 1997
to 2010, Balch et al. (2016) reported the average GHG emission from biomass combustions is
about one-third of the GHG emitted by fossil fuel combustion. However, this study reported
burning some types of biomass (e.g. evergreen broadleaf tree, deciduous needle leaf and
savannas) emitting higher GHG than fossil fuel. Additionally, the GHG emissions from
combusting biomass are also varied according to the energy conversion technology used, and
the assumption made when estimating the emissions (Balch et al., 2016). The variation in GHG
emissions estimations is further influenced by several other factors, including the
diversification of carbon accounting methods, the geographical dependence of the resources
and variation of technological options. The GHG emission can also be presented as an emission
range depending on the goal and scope of the study and its coverage of such variations above
(Varun, Bhat and Prakash, 2009). More recent literature also recognizes that increasing
complexity of electricity generation systems means that transparency of the data inventory and
openness to reveal the actual outcomes from the carbon accounting processes are important to
support the credibility of the emission values presented (Buytaert et al., 2011; Turconi, Boldrin
and Astrup, 2013; Strantzali and Aravossis, 2016).

6.2 Life Cycle Assessment for Biomass-Based Electricity Generation

LCA is a comprehensive tool to analyse the environmental impact throughout the life cycle of
a product, process or service, ideally from the cradle-to-grave (WBCSD, 2000; British

140
Standards Institution, 2011). ‘Carbon footprints’ based on GHG emissions are of major interest
from LCA studies but, typically, LCAs also output results for several other environmental
impact indicators such as acidification, resource depletion, ecotoxicities etc. LCA has been
used widely in academic and non-academic work to estimates the environmental impact in
various electricity and heat generation plants (e.g. Henrique et al., 2014; Asdrubali et al., 2015;
Treyer and Bauer, 2016; Beaudry et al., 2017). The environmental impact is estimated by
dividing the electricity generation process into the sub-processes to allow a thorough
examination of the input and output flows of each of these (Varun, Bhat and Prakash, 2009).
Furthermore, the division also enables the identification of ‘hotspots’, where in the electricity
generation system the most significant environmental impact occur (Piringer et al., 2016).
Martinez-Hernandez et al. (2013) used LCA to assess the environmental impact of biomass-
based CHP plant in the UK while Steubing et al. (2014) have identified the environmentally
optimal electricity generation plant using wood-based synthetic natural gas. Recently, Beaudry
et al. (2017) used an LCA approach to investigate the CO2eq emission of various utilization
routes for palm oil mill residues in Thailand. These studies illustrate the attributes of LCA and
show its application in this area for various project goals and scopes, the diversity of input and
output inventory data that can be used and the range of environmental impact indicators
presented. The range of possible configurations of even quite similar LCA studies illustrate the
challenges inherent in setting consistent and specific methodological choices and harmonizing
the results of assessments (Archer, Murphy and Steinberger-Wilckens, 2018).

Various studies have acknowledged the flexibility and universality of LCA and, in general,
suggest that it should be carried out on a case-by-case basis (Choo et al., 2011; Steubing et al.,
2014). Similarly, some authors have raised concerns over by how much and by how far the
results from LCA can be aggregated or generalized (Russell, Ekvall and Baumann, 2005;
Soimakallio and Koponen, 2011). Such methodological concerns over generalizing the
outcomes of LCA has led to difficulties in drawing consensus between the findings of different
studies even of rather similar systems.

Overall, although there is flexibility in conducting LCA studies for biomass-based electricity
generation, it is typically carried out in four sequential phases, which are goal and scope
definition, inventory analysis, environmental impact assessment and result interpretation.
These phases have been set out in two international standards: ISO 14040 (LCA Principles and
Framework) and ISO 14044 (Requirements and Guidelines for Carrying out LCA) (British
Standards Institution, 2006, 2018)

141
Different LCA studies have assessed the environmental impact of palm oil plantations and
crude palm oil production. Nevertheless, there are rather few studies conducted to assess the
environmental impact of the use of POMR for various value-added purposes (Beaudry et al.,
2017). In general, most oil palm focussed studies have been focused in the production of
biodiesel, briquette, pellets, fertilizers, and bio-charcoal (Sampattagul, Nutongkaew and
Kiatsiriroat, 2011; Chiew and Shimada, 2013; Soraya et al., 2014; Siregar et al., 2015;
Stichnothe and Bessou, 2017). Surprisingly, little attention has been given to estimating the
environmental impact of generating surplus, renewable, exportable electricity from mills from
PO mills from POMR, despite the recognised possibility of their use for this purpose
(KeTTHA, 2008; Economic Planning Unit, 2015, Beaudry et al., 2017). Chanlongphitak et al.
(2015) investigated the different environmental impacts of the electricity generation cycles
using EFB combustion in Thailand and determined that approx.. 150 kgCO2eq /MWh electricity
generated occur, based on an energy-basis allocation procedure for products and co-products
of the palm oil milling process. Similarly, Chiew and Shimada (2013) indicate approx.100
kgCO2eq/MWhelec were emitted from POMR-based CHP plant in Malaysia. GHG emissions
from the electricity generated from palm oil mill residues was also considered in the simulation
study by Beaudry et al. (2017). These authors estimated approx. 31 kgCO2eq/MWh emitted
from the combustion of residues from a 50 t/hr palm oil mill in Thailand. The variation of the
findings from these studies indicates that while they are conducted on similar system, their
outcomes are influenced by their goals and scopes, system boundaries, data inventories,
technology modelled and the geographical location (Chiew and Shimada, 2013; Beaudry et al.,
2017).

6.3 Life Cycle Assessment for POMR-RES

The following sections and sub-sections set out the LCA research conducted in this thesis
following the principles and general approach of BS EN ISO 14040/44. The LCA was
conducted as academic research. Therefore, it was intended that the findings are presented in
this PhD thesis and associated publications in the literature as open research enquiry available
to interested parties.

6.3.1 Goal and Scope

The purpose of this LCA is to evaluate the environmental impacts (EI) associated with
electricity generation in EFB-based power plants. The Functional Unit for this LCA is defined
as:
142
“the production of 1 kWh of electricity generated from the combustion of EFB at the oil palm
mill and exported to the national grid in Peninsular Malaysia”
This LCA is configured as a cradle-to-grave study to include the oil palm plantations, crude
palm oil (CPO) milling process, combusting the empty fruit bunches (EFB) and on-site
electricity generation incorporating grid connection. The LCA was modelled using the SimaPro
v 8.4.0.0 software. The present data inventory of the Malaysian palm oil industry and electricity
generation sectors were obtained from the Eco-invent v-3 database (Wernet et al., 2016). The
EI from the production of 1 kWh of electricity in the EFB-based power plant at economically
feasible size (5.70 MW generating 45.60 GWh of electricity annually, see Chapter 5) was used
in a benchmarking comparison with electricity from the national grid to evaluate its potential
environmental benefits as well as the possible drawbacks.
Figure 6.2 presents the system boundary for the modelling of the on-site EFB-based power
plant. This is broadly similar to that of Chanlongpitak et al. (2015), Shafie, Masjuki and Mahlia
(2014) and Fran et al. (2013). The LCA models the environmental impact of 1 kWh of
electricity generated in the EFB-based power plant for each sub-process, starting from the
plantation production of FFB, production of the CPO with the EFB (and other co-products) co-
produced alongside CPO, combustion of EFB, and finally the on-site generation of the
exportable electricity. The power plant under study was aimed to maximize the electricity
generation potential from EFB. In this regard, the mill’s operational heat and electricity demand
were fulfilled by using the existing internal energy generation system using MF and PKS.
Information on this arrangement was gathered from interviews conducted with millers and
consultations with SEDA.

143
Fuel FFB
Oil Palm Plantation (3.9 tonne)
Pesticides
Transportation
Fertilizers

CPO
Fuel Palm Oil Milling Process (1 tonnes)
- Sterilization US$ 541 / tonne
Water - Stripping
- Oil Extraction PKO
- Oil Clarification (0.09 tonnes)
USD 851 / tonne

MF PKS EFB POME


(0.5 tonnes) (0.12 tonnes) (0.9 tonnes) (2.54 tonnes) Biogas
US$ 9 / tonne US$ 30 /tonne US$ 5 / tonne
tonne
Electricity
(450 kWh)
Existing In-House Power
Heat Generation Combustion Electricity Bottom and Fly
Generator Ash
Electricity
Exhaust Heat

Co-generation Process from POMR-RES


SimaPro
Figure 6.2: EFB-Based Power Plant

144
Only limited site-specific Life Cycle Inventory (LCI) data were available to conduct the LCA
for the EFB-based power plant as there are currently only inventories for the plantation of FFB
and the milling of CPO available in the literature (Table 6.1). Although these two processes
have been subject to previous LCA study, Table 6.1 shows that the inventories from such
studies are incomplete and inconsistent. Even though the inventory in Choo et al. (2011) is
considered as the most comprehensive, it still omits several parameters. Choo et al (2011) in
their study assumed that the mill’s electricity and heat demand are supplied by the internal
boiler with PKS and MF as the feedstock.

Data availability constraints can be attributed to various factors such as the traditional method
of cultivating oil palms (i.e. following the long-established method with little attention given
on documentations) as well as unique palm oil milling processes. Based on the interviews with
two oil palm plantation owners, it can be concluded there is still no standard approach used to
measure the amount of fertilisers and pesticides being used during the plantation cycle. They
pointed out that the amounts of fertilizers and pesticides used differ based on the types of soil,
the type and price of the fertilizers and pesticides, as well as the weather conditions. Therefore,
in most cases, the amount of fertilizer is self-estimated depending on the suitability and current
needs of the plantation (pers. comms., 13/07/2015 – 30/06/2016). From the previous visit to
the mills, it could be summarized that, most of the millers are somewhat unwilling to share
information regarding the types and amounts of chemical substances and lubricants used in
their milling process.

Due to such constraints on data availability, this study has used generic data for oil palm
plantation and crude palm oil milling processes in Malaysia, based on and adapted from the
Eco-invent database with cross-checks with literature, agencies such as MPOB and personal
communications with millers. Similarly, the industrial boiler model from the USA for
combusting a wood waste with 50% moisture content and the 100 MW gas power plant module
(rest of the world) were used to represent the overall emissions from electricity generation
process from EFB. It is assumed that the information from these databases is a reasonable
representation for Malaysia biomass combustion in the modern system, and thus for the
environmental impact from the electricity generation. The specific LCI data used in this LCA
are found in Appendix D.

145
Table 6.1: Life Cycle Inventory for Oil Palm Plantation and Milling Process – Data Comparison

Country Malaysia Malaysia Indonesia Thailand Malaysia Thailand


Authors As modelled in Choo et al. (2011) Soraya et al. Kaewmai et al. Yusoff & Hansen Sampattagul et al.
this study (2014) (2012) (2007) (2011)
Input
FFB (tonnes) 3.9 5.2 6.28 6.07 5 5.8
Land Use (ha) - - 0.134 - 0.25
Fertiliser (tonnes)
N-fertiliser (as N-) 0.01 0.214 0.0002 0.024 0.220
• Ammonium Sulphate 0.006 0.002 - - - 0.154
• Ammonium Nitrate - 0.005 - - - -
• Ammonium Chloride - 0.003 - 0.0002 - -
• Urea - 0.002 0.104 - - -
• NPK Mixed - 0.036 - - - -
• Muriate of Potash - 0.005 - - - -
0.001 0.031 0.110 - - 0.066
• Phosphate
- 0.07 - - - -
Kieserite
Borate 0.003 0.004 0.01 - - 0.033
Dolomite - - 0.098 - - -
- - 0.232 - - -
Bunch Ash
Phosphorus 0.009 - - - 0.007 -
- - - - 0.043 0.22
Potassium
Magnesium - - - - 0.012 -
Herbicide (tonnes)
Paraquat - - 0.002 - - -
Glyphosate - - 0.003 - - 0.0018
Koalin - - - 0.012 - -
Sodium Chloride - - - 0.0004 - -
Hydrochloric Acid - - - 0.0005 - -
Sodium Hydroxide - - - 0.0005 - -

146
Pesticide (tonnes) 1.142E-7 - - - - 0.0014
Electricity from Grid 135 95.81 10.63 373 313
(kWh) 135 - - - -
• Nursery -
• Plantation - - 95.81 10.63
• Mill (Grid) -
Diesel (litre) 26.81 4.90 3.28 34 60
• Nursery 12.03 - - - 56
• Plantation 8.50 - - 4 4
• Transportation (field 0.026 6.29 4.90 - 30 -
to mill)
• Milling - - 3.28 - -
Water (litre) 7912 5750 5310 - 4095
• Nursery - 759 - - - -
• Plantation - - - - - -
• Mill (steam and - 7153 5750 5310 - -
cooking fruitlets)
Output
CPO (tonne) 1 1 1 1 1 1
PKO (tonnes) 0.09 0.09 - 0.05 - -
EFB (tonnes) 0.90 1.20 1.47 1.21 1.17 1.30
POME (m3) 3.54 3.83 3.86 2.50 3.88
Shell (tonnes) 0.07 0.17 0.22 0.37 0.38 0.35
Kernel (tonnes) 0.05 0.27 0.35 0.20 0.33 0.40
Fibres (tonnes) 0.50 0.67 0.83 0.58 0.78 0.78

147
The Life Cycle Impact Assessment (LCIA) method used was ReCiPe 2016 MidPoint (H)
v1.00. This method can represent 18 midpoint impact categories such as global warming
potential, acidification, eutrophication, human toxicity, mineral source depletion and fossil fuel
depletion. It has been used in several LCA studies investigating the environmental
consequences of the bioenergy systems (Vázquez-Rowe et al., 2014; Prado, Wender and
Seager, 2017).
An economic allocation method was adopted in this study to apportion the environmental
burdens from the milling process between the primary products and the co-products. In this
regard, the CPO and PKO are the primary products while EFB, POME, PKS, and MF are co-
products. An economic allocation was adopted as it is considered the most rational way to
allocate a fair share of the environmental burdens for the systems which produce large
quantities of co-products with low economics values (Ardente and Cellura, 2012). It has been
used widely by LCA practitioners to study the electricity generation from biomass and biofuels.
This is because mass- and energy-based allocation are considered inappropriate to describe the
actual technical value of the biomass to be used in electricity generation (Frischknecht, 2000;
Guinée, Heijungs and Huppes, 2004; Stougie et al., 2018).
Table 6.2 summarizes basis for the economic allocation share of the environmental burden
from the milling process based on milling 3.9 tonnes of FFB per tonne of CPO produced as
specified in Eco-invent database (see Table 6.1 and Table D-1 and D-2). This amount found to
be compatible with the actual amount of FFB required to produce one tonne of CPO in the
mills in Peninsular Malaysia (approx. 4 to 4.5 tonnes FFB) (Tenaga Sulpom, pers. comm.,
27/09/2017; Havys Oil Mill Sdn. Bhd, pers. Comm., 06/06/2016). The 2018 average price of
the CPO and PKO for Peninsular Malaysia was obtained and adapted from the Economics and
Industry Development Division, MPOB. The price of the co-products was obtained from
personal communication with the millers and from the literatures as described in Chapter 5
(Kazi et al., 2010; Abdul Malek et al., 2017).

148
Table 6.2: Economic Allocation for Products and Co-products of CPO Milling Process
Products/Co-Products Weight (t) Price (US$/t) Share (%)
CPO 1.00 541 86%
PKO 0.09 851 12%
EFB 0.60 5 0.5%
MF 0.50 9 0.7%
PKS 0.12 30 0.6%
POME * 2.54 0 0%
* As discussed in Chapter 3 and Chapter 5, there is no external market for POME and
it is assigned as a 0 economic value. It is perhaps best regarded as a by-product of negligible
that has no economic value.

6.3.2 LCIA Results and LCA Interpretations and Discussion

In this section, the selected environmental impacts (EIs) for the FU of 1 kWh electricity
generated in the EFB-based power plant are discussed. To generate 1 kWh of export electricity,
1.44 kg EFB (from milling 9.6 kg FFB) are required to produce the 10 MJ heat energy to drive
the back pressure steam turbine and the electricity generators. The specific EIs used were global
warming potential (GWP), terrestrial acidification potential (TAP), freshwater eutrophication
potential (FEP), human carcinogenic toxicity potentials (HTPsc), human non-carcinogenic
toxicity potentials (HTPsnc) and fossil fuel scarcity (FFS) as discussed below.

Although many LCA and related studies often highlights GWP, it is also important to assess
the acidification, eutrophication and toxicity potential contributed from e.g. the application of
inorganic fertilizers (e.g. ammonium sulphate, ammonium nitrate and phosphate) during the
plantation of the oil palms (Kimming et al., 2011; Havukainen et al., 2018). The TAP expressed
as SO2eq measures the acid formation potential in soils arising from the atmospheric
decomposition of inorganic fertilizers commonly from nitrogen, sulphur, hydrogen chloride
and fluoride. Furthermore, the SO2 and NOx emission from combustion process also increases
the TAP in biomass-based electricity generation. Acid formation to some extend exceed the
natural acidity level in soils change the acidity level in the soil, reduce the soil pH level which
causing a species loss (Huijbregts et al., 2016). The FEP is mainly caused by increase in the
nutrient concentration from phosphorus discharge into freshwater. This leads to additional
nutrient uptake by autotropic organisms such as algae and heterotrophic species such as fish
and invertebrates that can cause loss of organisms and species (Borrion, McManus and

149
Hammond, 2012). The HTPc and HTPnc measure the degree to which the chemical substance
can increase the risk of cancer and non-cancer disease in humans (McKone and Hertwich,
2001). Lastly, FFS measures the amount of fossil fuel used across the lifecycle of electricity
generation process. The choice of these EI categories is based on the fact that they are common
in the presentation of the environmental performance of bioenergy and its comparison with
fossil fuels (Cespi et al., 2014; Patel, Zhang and Kumar, 2016).

6.3.2.1 Land Use Change

It is commonly known that the effect from land use change (LUC) is an important consideration
when assessing the EI from biomass energy system. There are two aspects of LUC – direct
LUC (dLUC) and indirect LUC (iLUC). In this regard, dLUC occurs when there is a change in
the way land is used, for example, conversion from rain forest, peat forest or rubber plantation
to oil palm plantation. dLUC may result in land quality reduction, biodiversity loss and carbon
stock imbalance (Wu, 2008). On the other hand, iLUC is described as an indirect effect of a
change in use of a given piece of land, for example, switching the use from soy bean production
into corn production leading to ‘new’ geographically-disconnected land being required
elsewhere to maintain the production/meet market demand for soybean (Ahlgren and Di Lucia,
2014).
In Peninsular Malaysia, the dLUC for palm oil is very limited and governed by the Land
Conversion Act 1960. In East Malaysia, especially in Sarawak, the dLUC is governed under
different types of act because of the degree of autonomy (Gunarso et al., 2013). Vijay et al.
(2016) emphasized that until 2010, the rapid dLUC from conversion of forest to oil palm
plantation occurred in East Malaysia. Further regarding dLUC, Wicke et al. (2011) develop a
projection for oil palm in Peninsular Malaysia for 15 years and states that the CPO demand up
to 2025 can be met without further forest cover loss. This study projected that there are
adequate areas of deforested, agricultural and degraded land for replanting the oil palm to full
fill the CPO demand in the future. Likewise, based on the data gathered from the Economics
and Industry Development Division of the Malaysia Palm Oil Board has shown in Figure 6.3,
that there has been insignificant change in the oil palm plantation area in Peninsular Malaysia
from 2013 to 2017 (EIDD, 2018).

150
14
12
10
8
6
4
2
0
2013 2014 2015 2016 2017
Year

Area (ha) CPO Production (tonnes)

Figure 6.3: Peninsular Malaysia Oil Palm Plantation and CPO Production Trend (EIDD,
2018)

Various studies have identified iLUC as an important factor in the GHG profile and
sustainability of a number of bioenergy systems (e.g. Plevin et al., 2010; Van Stappen, Brose
and Schenkel, 2011; Yui and Yeh, 2013). However, in the case of Peninsular Malaysia, the
effect of iLUC in the oil palm sector is considered negligible because i) most palm oil
plantations were established on forest land or declining rubber plantations and ii) due to the
‘grandfathering’ effect of much of those oil palm plantations being established many years ago.
(Böhringer and Lange, 2005; Martinez and Neuhoff, 2005; Zulkifli et al., 2010). For instance,
the palm oil mill in Melaka (Mill B) receives FFB from three different oil palm plantations
which have been established for more than 30 years (Sri Lingga pers. comm., 20/07/2015).
Therefore, the EI associated with electricity generated in the EFB-based power plant was
assessed with the base-case assumption that there is no effect of dLUC and iLUC from the oil
palm plantation. However, some publications have reported the occurrences of LUC associated
with oil palm plantation (e.g. Padfield et al., 2011; Petrenko, Paltseva and Searle, 2016; Vijay
et al., 2016). Therefore, the EI was analysed based on two scenarios, first, the base-case without
LUC effects and second, using the aggregated land use factor for ‘general’ oil palm plantations
in Malaysia in the Eco-invent database. The database suggested that 2.258 × 10−5 ha of land
is required to produce one kg of FFB. Additionally, Page et al. (2011) posited that the annual
GHG emission factor for LUC for oil palm plantations in Malaysia as 10,600 tonnes CO2eq/ha.
Considering a production of about 20 tonnes FFB/ha, the annual GHG emission due to LUC
for this second scenario based on this figure is estimated as 0.541 kg CO2eq/kg FFB.

151
6.3.2.2 Oil Palm Plantation

The EI from the production of the 9.6 kg FFB for the with/without LUC are presented in Table
6.3 and Figure 6.4. It is assumed that the oil palms are planted on mineral soils at a planting
density of 142 palms per hectare (Zulkifli et al., 2010).

Table 6.3: Environmental Impacts from the Production of 9.6 kg of FFB for the FU at the
plantation
Impact Unit Without LUC With LUC
GWP kg CO2eq 0.867 8.97
TAP kg SO2eq 0.0134 0.0148
FEP kg Peq 0.000242 0.000243
HTPc kg 1,4-DCB 0.0167 0.0225
HTPnc kg 1,4-DCB -29 -28.9
FFS kg Oileq 0.0993 0.107

100

50

0
%

GWP TAP FEP HTPsc HTPsnc FFS

-50

-100
Environmental Impact

FFB w/o LUC FFB w LUC

Figure 6.4: Environmental Impact from the Production of 9.6 kg FFB (for the FU) at the
Plantation
The inclusion of LUC resulted in large increase in the GWP for the FFB required for the FU
compared with the without LUC base case. The other EIs shown minimal changes. The GWP
from the production of 9.6 kg FFB with consideration of the LUC effect is ten times higher
than GWP without LUC. Previous studies have acknowledged the GHG released from LUC is
a significant contributor to the carbon profile of biofuel and bioenergy (Kim, Kim and Dale,

152
2009; Stichnothe and Bessou, 2017). Correspondingly, Kim and Dale (2011) emphasized that
the overall GHG reduction benefits from biofuel and bioenergy are sensitive to the both dLUC
and iLUC factors. Stichnothe and Bessou (2017) perceived a strong interlink between
maximizing bioenergy production, reducing environmental impacts and ensuring sustainable
land use change in the palm oil industry in Indonesia. However, Searchinger et al. (2008) and
Perlack et al. (2005) asserted that, due to LUC, the GHG footprint from the production of
bioenergy and biofuel could only be minimized when bioenergy and biofuel were derived from
waste such as municipal solid waste and crop waste.

Based on the data in Table 6.3, it was found that the production of 9.6 kg FFB emits
approximately 0.867 kg CO2eq without LUC and 8.97 kg CO2eq with LUC. The base case GWP
measured was found to be lower than the 3.3 kg CO2eq made by Soraya et al. (2014), Choo et
al. (2011) and Zuklifli et al. (2010). The EI measured from the FFB plantation are contributed
from the utilization of ammonium sulphate (N-type fertilizer), herbicide, transportation of
planting material from the nursery to the plantation and transporting FFB from plantation to
the mill, as well as the use of pesticides as shown in Figure 6.5. The ‘negative’ impact in HTPnc
is due to mechanistic assumptions in the LCA methodology whereby mineral and toxic
materials in the soil are absorbed by the oil palm plantation in the plantation. As some of this
plantation biomass is removed by harvesting, there is a removal of these materials form the
plantation hence reducing the local HTPnc score.

153
100

50

0
GWP TAP FEP HTPc HTPnc FFS

-50
%

-100

-150

-200
Environmental Impact

FFB Ammonium Sulphate Herbicide Transportation Pesticide

Figure 6.5: Environmental Impact Contributors for FFB Plantation

6.3.2.3 EFB Production Life Cycle Phase (Cumulative, after Allocation)

Table 6.4 presents the EI for producing the 1.44 kg of EFB needed for the FU via the milling
process of the FFBs. This production emits 0.0046 kg CO2eq without LUC and 0.042 kg CO2eq
with LUC respectively. The EI presented here is a cumulative value after allocation, i.e. it
includes both the impacts from the oil palm plantation and the milling process of CPO; the
values represent the proportion of the upstream burdens that have been allocated (economic
basis) to the EFB.

154
Table 6.4: Environmental Impacts from the quantity of EFB Production for the FU
(Cumulative, Allocated)
Impact Unit Without LUC With LUC
GWP kg CO2eq 0.0046 0.042
TAP kg SO2eq 7.03E-5 7.66E-5
FEP kg Peq 1.41E-6 1.42E-6
HTPc kg 1,4-DCB 0.00015 0.00017
HTPnc kg 1,4-DCB -0.0033 -0.00284
FFS kg Oileq 0.0006 0.0006

There is still a substantial effect of including the LUC factor on the GWP score at this stage of
the system, it being nine times higher with LUC than without. The plantation stage is the major
contributor to the EIs in Table 6.4.

The EI from EFB, as an individual co-product from CPO milling process, is still not widely
reported. A few studies, for example, Beaudry et al. (2017), Chanlongpitak et al. (2015) and
Reeb et al. (2014) have attempted to report the GHG emission from EFB (without LUC).
Beaudry et al. (2017) determined the emission from EFB produced in a 50 t/hr FFB processing
mill as 0.015 kg CO2eq/kg EFB. Chanlongpitak et al. (2015) used a mass allocation approach
to estimate the GHG emission from EFB in Thailand while Reeb at al. (2014) used an economic
allocation approach to estimate the GHG emission from EFB in Malaysia with the results being
0.004 kg CO2eq /kg EFB in Thailand and 0.0009 kg CO2eq/kg EFB in Malaysia. The emission
obtained in the present research reasonably agrees with Chanlonngpitak et al. (2015), although
both studies used different environmental burden allocation approaches. In this light, the
present study has made a clear and transparent presentation of an economically allocated
individual EFB emission evaluation. These results provide evidence of an EI profile for EFB
that can be useful in wider studies on EFB potential for electricity generation and/or for other
value-added products.

6.3.2.4 EFB Combustion Life Cycle Phase (Cumulative, after allocation)

The EIs from combusting 1.4 kg of EFB using the industrial boiler model for wood waste from
the USA are presented in Table 6.5. This model assumed that the wood residues has the
moisture contents typically near 50% with CO2, CH4 and N2O are produced during combustion
process. Almost all the carbon content (99%) in the residues is converted to CO 2 during the
155
combustion process and as this CO2 is considered as part of short-term CO2 cycle of the
biosphere and therefore not counted as a GHG emission (EPA, 2003). Furthermore, the
formation of N2O is minimized by keeping the combustion temperature above 800°C and small
amount of CH4 is emitted during incomplete combustion and during the period of low
temperature (i.e. start-up or shut-down cycles for boilers).

Table 6.5: Environmental Impacts from EFB Combustion (Cumulative, Allocated)


Impact Unit Without LUC With LUC

GWP kg CO2eq 0.0105 0.103


TAP kg SO2eq 0.00135 0.00158
FEP kg Peq 0.000 3.12E-06
HTPc kg 1,4-DCB 2.000 2.000
HTPnc kg 1,4-DCB 0.544 0.538
FFS Kg Oileq 0.000 0.000

As expected, there were only substantial differences in the impact values in the GWP category
when LUC factor had been taken into account. Combusting 1.44 kg EFB produced 10 MJ heat
energy with the cumulative GWP emission of between 0.0105 to 0.103 kg CO2eq. While there
is still no specific data on the GHG emission from combusting EFB in the literature so far, past
studies have extensively assessed the environmental impacts of combusting other types of
biomass e.g. work by Caserini et al. (2010) shows about 0.08 – 1.08 kg CO2eq emitted from
combusting one kg of dry biomass in Italy. Chungsangunsit, Gheewala and Patumsawad (2005)
estimated the combustion of rice husk has nearly zero GHG emission, taking advantage of the
natural carbon cycle. Kimming et al. (2011) and Havukainen et al. (2018) acknowledged
biomass combustion has lower GHG emission than fossil fuel; however, highlighting that
combusting biomass emitted a higher amount of sulphur dioxide (SO2) and nitrogen oxide
(NOx) resulting in higher acidification and eutrophication potential.

The GHG emission value from the combustion of 1.44 kg EFB without LUC obtained in this
analysis is minimal compare to the emission from natural biodegradation of 1.44 kg EFB at the
50% anaerobic condition. The mid-point estimation made by Hansen, Olsen and Ujang (2012)
and Stichnothe and Schuchardt (2011) for, natural biodegradation of 1.44 kg EFB emits approx.
0.94 kg CO2eq. The lower GHG emission from EFB combustion than the natural biodegradation

156
further supports the findings from Beaudry et al. (2017) and Chiew and Shimada (2013) and
reinforces the potential usefulness of EFB as a feedstock for power generation.

6.3.2.5 EFB-based Power Plant

The EIs for the whole 5.7 MW generating plant used to model the 1 kWh FU of electricity are
tabulated in Table 6.6. These EIs were derived from Ecoinvent dataset for Gas power plant,
100 MW electrical {RoW}|construction|Alloc Def, U accounting for materials and energy used
to construct at 100 MW gas power plant with 180 000 hours operation time. The power law as
in Eq. (14) with scaling factor of 0.8 was used to scale down the power plant to 5.7 MW. The
accumulated emission from the construction of 5.7 MW power plant is then divided equally
across the 1.02 TWh of total output of electricity to generate a contribution for the generating
plant ‘infrastructure’ for the FU.

Table 6.6: Environmental Impacts from the assigned proportion of the infrastructure 5.7 MW
POMR-RES

Impact Unit 1 kWh of


Electricity
GWP kg CO2eq 0.657E-6
TAP kg SO2eq 5.224E-9
FEP kg Peq 0.122E-6
HTPc kg 1,4-DCB 92.592E-9
HTPnc kg 1,4-DCB 0.149E-3
FFS Kg Oileq 0.167E-6

6.3.2.6 Overall Impacts of Electricity Generated in POMR-RES

Based on the cumulative, after allocation EIs from EFB combustion and the EIs for the 5.7
MW EFB-based power plant, the GHG emission and other EIs associated with 1 kWh
generated from EFB are given in Table 6.7.

157
Table 6.7: Environmental Impacts for the FU of 1 kWh electricity from EFB after inclusion
of the power plant (Cumulative, Allocated)

Impact Unit Without LUC With LUC


GWP kg CO2eq 0.0105 0.1032
TAP kg SO2eq 1.5E-3 1.58E-3
FEP kg Peq 1.22E-7 3.24E-6
HTPc kg 1,4-DCB 2.000 2.000
HTPnc kg 1,4-DCB 0.544 0.538
FFS Kg Oileq 1.67E-7 1.67E-7

The approx. 0.01 kg CO2eq GHG emission without LUC obtained in this analysis is lower than
the 0.147 and 0.232 kg CO2eq/kWh electricity generated in Chanlongpitak et al. (2015) and
Chiew and Shimada (2013) respectively. This may due to how Chanlongpitak et al. (2015)
have accounted for their GHGs in their modelling and possibly their mass allocations to their
co-products although further understanding is limited by their somewhat untransparent LCI
dataset. Chiew and Shimada (2013) modelled a combustion of a mixture of EFB, MF and PKS
for their electricity generation. The approx. 0.010 kg CO2eq/kWh GHG emission obtained here
without LUC is lower than the 0.845 kg CO2eq/kWh electricity generated estimation of Shafie,
Masjuki and Mahlia (2014) for rice straw. The substantial difference in the GHG emission
between the EFB-based electricity and rice-straw based electricity is believed to have been
contributed from the rice cultivation phase and its allocation due to high utilization of
fertilizers, the soil chemical and physical properties and irrigation water (Singh and Singh,
2017). The GHG emission from EFB without LUC in the present study is lower than that
reported for wood residues (0.265 kg CO2eq/kWh), poplar (0.159 kg CO2eq/kWh) and willow
(0.122 kg CO2eq/kWh) (González-García and Bacenetti, 2018). Furthermore, the estimated
emission from 1 kWh electricity generated in this analysis is also lower than the estimation of
approx. 0.22 to 0.43 kg CO2eq/kWh electricity made by Shafie (2015) in his preliminary study
using rice husk in Malaysia. It is observed that the GHG emission from 1 kWh of EFB-based
electricity obtained in this analysis are also towards the lower end of the GHG emission range
for various biomass-based electricity generation systems provided by Kadiyala, Kommalapati
and Huque (2016) (0.007 kg CO2eq/kWh – 2.54 kg CO2eq/kWh). Based on this range, it can be

158
concluded that the GHG emission from electricity generated from EFB is lower than several
types of biomass (e.g. municipal solid waste, wheat straw, poplar) but is comparable with forest
residues, miscanthus and short rotation chips.

6.3.3 LCA Interpretation

In this section, the main findings and interpretation of the LCA are summarized, primarily via
a hotspot analysis. The limitations of the LCA modelling are also discussed alongside
suggestions for further refinement.

6.3.3.1 Hotspot Analysis of Electricity Generation via POMR-RES

A hotspot analysis was conducted to identify where, or in which, sub process of the life cycle
(i.e. plantation, EFB production, EFB combustion, on-site electricity generation) the main
drivers occur for the overall EIs (see Figure 6.6).

100

50

0
GWP TAP FEP HTPc HTPnc FFS
%

-50

-100

-150
Environmental Impact

Oil Palm Plantation EFB EFB Combustion Power Plant

Figure 6.6: Hotspot Analysis of Electricity Generation Process


The hotspot analysis of the base case (without the LUC) revealed the oil palm plantation phase
to be the main contributor to all EIs. The plantation contributes approx. 98% to GWP, 90% to
TAP, 99% FEP, 99% to HTPnc and to 99% to FFS, respectively. A similar scenario is also
159
reported by Chary et al. (2018) where LUC and fertilizer during energy cane cultivation (no
allocation) contributed 45% to the overall GWP of the electricity generation process. Shafie
(2015) reported that 97% of GWP is contributed by N2O emission during paddy plantation
coupled with diesel consumption for field operations (for ploughing, harvesting). Adams
(2017) emphasized that cultivation and harvesting stages are usually accountable for the largest
source of emissions in the crop-based electricity generation lifecycle where the emission from
the production process of N-fertiliser and diesel consumption as well as N2O emissions from
the soil were identified as the main contributors. Other studies, e.g. Schmidt (2010) and Zulkifli
et al. (2010) have also arrived at the similar conclusions. The EFB combustion process, on the
other hand, contributes approx. 10% to TAP and 98% to HTP C which due to the releasing of
SO2, CH4, N2O and the sum of hydrocarbons during combustion process (Kuprianov,
Kaewklum and Chakritthakul, 2011). The hotspot analysis was also conducted for the with
LUC case. Similarly, the oil palm plantation phase remained as the main contributor of GWP,
TAP, FEP, HTPnc and FFS while combustion process holds the highest contribution to HTP c.

In this regard, considering that electricity via POMR-RES has potential to increase the energy
security and to contribute towards the national renewable energy target, further efforts to
reduce the EIs especially from oil palm plantation is required. Adams (2017) highlighted that
the GWP from biomass-based RE is lower if it is sustainably sourced and harvested. Afriyanti,
Kroeze and Saad (2016) and Saswattecha et al. (2017) highlighted that emissions produced at
the oil palm plantation stage can be controlled and mitigated by integrating more sustainable
oil palm plantation and harvesting practices. Continuous effort has been made by the
Roundtable on Sustainable Palm Oil (RSPO) and MPOB to facilitate more sustainable oil palm
plantation in Malaysia (Abdullah and Sulaiman, 2013). Furthermore, the Malaysian
Sustainable Palm Oil Standard (MS 2530-4:2013) has been introduced to mandate a sustainable
palm oil industries in Malaysia (Department of Standard Malaysia, 2012). From such on-going
efforts to ensure more sustainable oil palm plantation, the GHG emission reduction from oil
palm plantation can be considered as realistic: however, achieving this will require stringent
compliance controls, wider participation and engagement from various actors, and enhanced
collaboration with relevant agencies and stakeholders both local and international.

160
6.3.3.2 Limitations of the LCA

Based on the LCA results interpretation so far, it is believed that these LCA results were conducted in
line with the state-of-the-art literature in this area. With limited ability to do further detailed work within
the research scope and from a conservative perspective, the LCA analyses presented here is best
regarded as the LCA scoping assessment based on the best available data to give insight on the amount
of GHG saving that is potentially available from POMR-RES. Although attempts were made to gain
engagement from the data providers, this was not always possible. Further work to obtain more detailed
and granular data by working with individual mills is highly desirable in order to build more
comprehensive and site-specific LCI datasets that can provide precise and transparent LCA results and
at the same time minimize uncertainty. Cooperation between Malaysian and international LCA
practitioners is likely to be very helpful to build the engagement and trust with data providers.

6.3.4 Comparison of Electricity Generation via POMR-RES with Peninsular Malaysia


Grid Average Generation

Comparing the environmental performance of RE generated via POMR-RES with the


Peninsular Malaysia grid average is critical in order to provide a sense of the potential
environmental benefits from such generation. This comparison quantifies the potential
reduction of EIs from this POMR RE generation vs grid average and how much it can
contribute towards decarbonizing the electricity supply system in Peninsular Malaysia. Table
6.8 and Figure 6.7 provide the comparisons between the EI of 1 kWh of electricity generated
by POMR-RES with and without LUC and electricity from the national grid (Electricity,
medium voltage {MY} market for | Alloc Def, U).

Table 6.8: Environmental Impacts Comparison of 1 kWh Electricity


Impact Unit POMR-RES National
Without With Grid
LUC LUC
GWP kg CO2eq 0.010 0.103 0.835
TAP kg SO2eq 1.5E-3 1.58E-3 0.00318
FEP kg Peq 1.22E-7 3.24E-6 0.000316
HTPc kg 1,4-DCB 2.000 2.000 0.000195
HTPnc kg 1,4-DCB 0.544 0.538 10.8
FFS kg oileq 1.67E-7 1.67E-7 0.222

161
100

80

60

40

20

0
%

GWP TAP FEP HTPc HTPnc FFS


-20

-40

-60

-80

-100
Environmental Impact

Without LUC With LUC National Grid

Figure 6.7: Environmental Impacts Comparison of 1 kWh Electricity from POMR-RES with
the National Grid

It can be concluded that every kWh of electricity generated in POMR-RES without and with
LUC emitted 98% and 86% less GHG emission compared with its grid average counterpart.
The GHG reduction from electricity generation from POMR-RES exceeds the sustainability
criteria of GHG saving set in the EU Renewable Energy Directive (50% saving). To date,
Malaysia has yet to impose any sustainability criteria for RE. Figure 6.7 reveals that generating
electricity from EFB does have impact on TAP which is attributed to the utilization of inorganic
fertilizers during the plantation stage. Butnar et al. (2010) and González-García and Bacenetti
(2018) reported that generation bio-electricity may have higher TAP than fossil counterparts
(especially for natural gas). These studies used switchgrass, willow and poplar as the feedstocks
for electricity generation and described forestry activities and diffuse emissions derived from
manure and mineral fertilizer application as the main contributors to this impact category.
Reddy and Venkataraman (2002) described the higher TAP from biomass-based electricity
generation as being due to the high sulphur content in biomass and the effect of incomplete

162
combustion. They reported that 60% of the sulphur content in biomass were converted to SO 2
during the combustion process.
Figure 6.7 also indicates that generating electricity from EFB showed a lower FEP lower than
the grid average which contradicts some previous literature which highlight that biomass-based
electricity generation has more eutrophication potential than electricity generation using fossil-
fuel (Rocha et al., 2014; Morales et al., 2015; González-García and Bacenetti, 2018).
According to González-García and Bacenetti (2018), for the case of poplar and willow, the
high FEP is contributed by the feedstock production related activities involving fertilization
and feedstock transportation. It also noted that, generating electricity from EFB has higher
HTPc than the fossil-based electricity. Wielgosinski, Lechtanska and Namiecinska (2017) and
Ren et al. (2017) indicate that this can be driven by significant amount of micropollutants,
among which are more toxic compounds such as polycyclic aromatic hydrocarbons and
dioxins.
The contradiction relatively low levels of FEP from POMR-RES in contrasts with a number of
other studies of biomass to electricity systems in the literature is due to those studies modelling
dedicated bioenergy crops, such as willow and poplar (Keoleian and Volk, 2005; Djomo,
Kasmioui and Ceulemans, 2011; Turconi, Boldrin and Astrup, 2013). Based on the economic
allocation approach in the present study, EFB was treated as a co-product that carries a very
small (0.5%) proportion of the plantation and milling environmental burdens of the cultivation
of the oil palm – most of which (98%) are allocated to CPO and PKO.
Higher TAP, FEP and HTPc have been described as the major drawbacks of generating
electricity from biomass (Borrion, McManus and Hammond, 2012; Rocha et al., 2014;
Havukainen et al., 2018). In this analysis, generating electricity from EFB has higher HTPc
than the national grid average but not for FEP and TAP (in fact the FEP of EFB is negligible).
On balance, even though POMR-RES under-performed the National grid average in terms of
HTPc, the benefits in all other categories, especially GWP, FEP and FFS as well as the other
human toxicity category, suggest an overall environmental benefit from generating electricity
from EFB.

163
6.3.5 Carbon Emission Saving from EFB-based Power Plant

This section presents two perspective on the annual carbon emission saving from the
economically size EFB based power plant.

1. Carbon emission saving from electricity generated in the EFB-based power plant when
compared with the emission from the same amount of electricity from the Peninsular
Malaysia grid average.
2. Emissions from combusting the EFB for electricity generation – in comparison with the
potential emissions if the EFB left for natural biodegradation.

At its economically feasible size (5.70 MW) without LUC, a single power plant is capable of
generating approx 45.60 GWh of electricity annually using 76,800 tonnes of EFB. Moreover,
based on the base line estimation by Hansen, Olsen and Ujang (2012) and Stichnothe and
Schuchardt (2011), about 650 kg CO2eq will be emitted from each tonne of EFB if they were
left to undergo natural biodegradation under 50 % anaerobic conditions. The annual carbon
emission savings from one EFB-based power plant and its relation with such avoided natural
biodegradation emissions are presented in Table 6.9.

Table 6.9: Carbon Emission Saving from Electricity Generated in Economically Feasible Size
Power Plant

45.60 GWh/year EFB-based Grid Electricity


76, 800 tonnes of EFB Power Plant
Carbon Emission 456 tonnes CO2eq 38, 000 tonnes
CO2eq
Carbon Emission Saving from 37,500 tonnes CO2eq
Electricity Generation
Carbon Emission Saving from 50,000 tonnes CO2eq
avoided natural biodegradation

Thus, approximately 37,500 tonnes of CO2eq could be saved from generating electricity using
EFB in a single 5.70 MW plant. Moreover, an additional 50,000 tonnes of CO2eq emission
could be avoided by managing EFB in this way rather than allowing their biodegradation. It is
interesting that avoided emission from natural biodegradation has a somewhat higher GHG
emission saving ‘value’ than the direct electricity feedstock substitution vs the national grid
average. Overall, it appears that a transition towards EFB-based electricity at this plant scale
164
could avoid a total of 87,500 tonnes CO2eq assuming that all the emission from the natural
biodegradation of EFB could be avoided during the EFB storage before it is used for electricity
generation. It is likely, however, that the natural biodegradation process cannot be entirely
avoided since EFB will need to be stored for certain period of time (perhaps several weeks or
months) to sustain the combustion process throughout the year. An appropriate management
of pre-combustion EFB is required and, hence the estimation of the actual carbon emission
saving should be refined when this detail is known.

6.4 Summary

In conclusion, this chapter has modelled the cradle-to-grave environmental impacts for a FU
of 1 kWh of electricity generated in the economically optimal sized (5.70 MW, see Chapter 5)
EFB-based power plant. The estimation was made using adaptation and extension of the
specific data inventory for Malaysia in the Eco-invent database, complemented with additional
data. The electricity generation process was divided into four sub-processes; FFB plantation,
EFB production, EFB combustion and on-site electricity generation. The environmental impact
of each sub-processes was assessed under two scenarios; with and without land use change
(LUC). It was revealed that there were substantial increases in the amount of CO2eq emission
when the effect of LUC was included. The ‘carbon footprint of electricity generated in EFB-
based power plant was modelled to be approx. 0.01 kg to 0.103 CO2eq/kWh generated without
and with LUC respectively, this being 98% to 88% lower than the emission from the current
national grid average for Malaysia. With an assumption that the minimal LUC can be seen in
Peninsular Malaysia and other factors that has been discussed in Section 6.2.3.1, therefore, the
emission from LUC is excluded from the generation electricity from EFB. However, generating
electricity using EFB leads to high TAP from the utilization of fertilizer and diesel during the
plantation of oil palm. Generating electricity using EFB also has higher HTP c than fossil fuel
due to the emission from combustion process.

At 5.70 MW EFB-based power plant was able to save in the order of 87,500 tonnes GHG
emission from the avoidance of grid electricity and of the natural biodegradation process of
EFB. The current findings substantially adds to our understanding of the potential
environmental impact of electricity generation from EFB and provides additional evidence on
the role of such biomass in decarbonizing the electricity generation system.

165
The LCA approach has been considered a useful tool to provide insights of the environmental
performance of electricity generation from EFB within the limitations that have been
highlighted. These insights and results are useful to support decision-making related to future
energy planning in Peninsular Malaysia. Furthermore, the findings can be regard as a valuable
guide for further in-depth research to assess the environmental impact associated with
electricity generation from POMR. The data-intensive nature of LCA means that, the issue of
the data gaps should be addressed to ensure a more accurate estimation of the environmental
impact from the use of PO residues for electricity generation. There is a need to establish a
comprehensive site-specific LCI database for palm oil plantations and CPO milling processes
that present detail about the actual process and their variation in Malaysia. A set of complete
and comprehensive site-specific LCIs will reduced the gaps associated with data availability,
and allows further improvement in estimation of the EIs for such ‘new’ renewable electricity
generation opportunities.

166
CHAPTER SEVEN

CONCLUSIONS AND FUTURE WORK

7.1 Conclusion

This work was devoted to assess the feasibility of generating renewable electricity from POMR
focussing on EFB and biogas available at the mills in Peninsular Malaysia. The study
demonstrates a positive case for generating renewable electricity from these residues.
Previously, limited scientific-based evidence on the feasibility of the electricity generation
process, disfunction of POMR-based electricity generation value chain and ‘non-conducive’
environment to develop POMR-RES are identified as the primary challenges that have limit
the translation of electricity generation potential into implementation. In this thesis, the
feasibility of the electricity generation system has been assessed based on the technical and
techno-economic performance as well as its environmental benefits. A comprehensive
technical and tecno-economic feasibility and environmental impact assessment framework is
used to illustrate that this approach for supplying renewable electricity to the national grid
(Peninsular Malaysia) can be considered technically feasible and consistent with also meeting
the parasitic demand of the electricity generation process and the operational demand from the
mills. Furthermore, POMR-RES is considered techno-economically feasible when it generates
a profitable electricity that provides 20% ROI. The POMR-based electricity generated is
expected to emits 95% lower GHG emission than the current fossil-based counterparts. This
thesis has also critically investigated a range of factors affecting the functionality of such value
chains and the ‘non-conduciveness’ of environment to develop POMR-RES. Practical
approaches to improve the functionality of the value chain and to create a more conducive
environment are proposed in this thesis to drive the translational of this electricity generation
potential.

This is done in the belief that, in a broader context, the work will contribute to efforts to
decarbonize the electricity supply system and subsequently contribute towards achieving
Malaysia’s national RE target and climate change commitments. The points below summarize
the key conclusions of this research.

167
i. An analytical approach to estimate accurately the available EFB and biogas at POMs
was based on research on the technical, economic and implementation considerations.
The amount of available EFB to be used for electricity generation were found to be less
than was previously reported. The amount of EFB available is only about 50% of the
theoretical amount available from PO milling while the availability of dried EFB is at
23%. In the case of biogas, only 23% of the total organic content of converted POME
is considered to become methane that can be used in the combustion process. Based on
these estimates, the risk of over estimating the residues that are available for electricity
generation from POMR can be reduced.
ii. An integrated technical and techno-economic assessment framework was applied to
evaluate systematically a range of potential POMR-RES installations. The multiple
scenarios modelled allowed a comprehensive assessment to be conducted to identify
suitable residues for electricity generation, propose a suitable hardware configuration
for the system (e.g. power plant and CHP plant) and recommend appropriate installation
schemes (e.g. on-site, off-site and co-operative) for POMR-RES. Sensitivity analysis
for the techno-economic performance of POMR-RES was also conducted using four
parameters. This modelling allows prediction of the available residues, generation
capacity, and connection to the grid more comprehensively than has been previously
reported.
iii. The EFB were found to have higher technical and techno-economic feasibility for
generating electricity than biogas. However, the estimated heat generated from biogas
is found to be sufficient to supply the operational heat demand from the mill. Co-
utilization of EFB to maximize the electricity generation potential and biogas to meet
the mill’s operational heat demand enabled release of the MF and PKS for more
competitive usage.
iv. The on-site electricity generation system provides more favourable technical and
techno-economics performance than the off-site electricity generation as a result from
including the cost to buy and to transport the feedstock from the mills to the proposed
generation location.
v. The feasibility to generate the on-site electricity is currently limited to the upper
medium- to large- scale mills with an electricity generation capacity higher than 5.7
MW. Approximately 76,800 tonnes of EFB is required to operate the power plant at
168
this scale. Additionally, the interest rate that imposed by the financial institution on the
project has to be reduced to 4% instead of the 10 to 12% currently.
vi. A co-operative generation is identified as one of the approaches to further improve the
feasibility of generating electricity among the small- and lower medium-scale mills.
This technology and resource pooling concept whereby the generation plant is shared
by some existing mills and developed in the most business-friendly location is
identified as the potential way to maximize the electricity generation potential from the
available residues in the mills and at the same time minimize the operational risk to
develop the POMR-RES.
vii. The identification of the economically feasible scale and residue types sets a benchmark
to help promote sustainable investment and residues utilization. In a broader
perspective, it indicates expected electricity outputs from POMR in relation to the
targets set in Malaysia’s policy for RE and provides insight for policymakers and
analysts to develop more reliable policies for POMR-based sustainable RE in the near
future. Reliable policies for POMR-RES should include clear policy direction and
policy impact pathways, informed in part by the present work, as a guide to achieving
the policy targets.
viii. A cradle-to-grave LCA was used to estimate the GHG emissions and other
environmental impacts per kilowatt-hour of electricity generated at the economically
feasible scale for POMR-RES. This estimation allowed comparison of the
environmental impacts of POMR-RES with the conventional electricity generation
system and existing waste management system for EFB. The incorporation of LCA in
this study provides improved and transparent confirmation of the environmental
contribution of POMR-RES as a basis for decision making and policy formulation.
ix. The cradle-to-grave LCA for the economically optimal size POMR-RES indicated the
electricity generated in EFB-based power plant is ‘greener’ than the current national
grid average for Malaysia. With an assumption that minimal LUC can be seen in
Peninsular Malaysia (i.e. the emission factor from LUC is excluded from the
environmental impact assessment), the GHG emission from every kilowatt-hour of
EFB-based electricity emits 98% less GHG than the current national grid average for
Malaysia.

169
x. The installation of the biomass dryer appears to be sensible when the efficiency of the
boiler without the dryer is lower than 70% with the provision of auxiliary heat resources
to dry the EFB. Modelling to assess the proposed biomass heating model showed that
although it can increase the electricity generation potential by approximately 10%, the
additional heat requirement for drying the biomass has increased the parasitic load of
the power plant at the same time leading to higher capital cost. Therefore, semi-dry
EFB with 45% moisture content is in a good position to be used as the feedstock for
POMR-RES.
xi. The advantages from generating electricity from POMR highlighted in this research
have not been widely appreciated previously which have been partly responsible for
low awareness and participation among stakeholders. This circumstance, together with
apparent reluctance of the off-takers (especially TNB) who tend to create hurdles to the
RE plant interconnections and approval of the RE generating capacities for various
operational and administrative reasons, has contributed to the limited number of
POMR-based power plants providing renewable electricity to the national grid to date
under the current economic environment. It is hoped that the present study will go some
way towards encouraging a wider take-up of this opportunity for renewable electricity
from POMR to be achieved.

7.2 Suggestions for Further Work

The research has identified many positive attributes of POMR-RES and indicated its potential
to make an important contribution to the generation of renewable electricity in Peninsular
Malaysia and potentially more widely. However, there remains much scope for further
development for the analyses conducted here and opportunity to extend the knowledge base on
such systems. Suggestions for useful areas for future research on POMR-RES are summarized
below:

i. Extending the analytical approach to incorporate the availability of other POM residues
beyond those investigated here would allow exploration of their potential to maximize
electricity generation or for the development of other value-added purposes. Methods
to determine with improved accuracy the ‘real’ availability of POMR would also be
beneficial for planning an integrated waste management system in the mills. This
170
improved resource assessment process will be data intensive and bespoke as, currently,
most of the data for PO and biomass residues in Peninsular Malaysia are not being
collected, maintained or archived systematically. Initiating this effort will be valuable
to minimize data uncertainties and accessibility concerning such resources in the future.
ii. While the research of this thesis has substantially increased knowledge and insight on
the techno-economic feasibility and environmental benefits of POMR-RES for
Peninsular Malaysia, implementation of such system will require additional elements
of the ‘conducive environment’ (referred to in Chapter 2) to be in place to enable full
adoption. Therefore, further work to explore the policy and investment routes that
effectively de-risk POMR-RES projects is likely to be essential to attract sufficient
investment from the prospective millers and project developers.
iii. Additional study to reduce the financial and administrative barriers and to facilitate
access to the national grid for RE is needed to raise confidence amongst the millers and
project developers to enter the POMR RE generation business.
iv. Further LCA works investigating the environmental profile of POMR-RES at different
scales is required to quantify the accumulated GHG emissions savings from using EFB
for electricity generation. For example, whilst some POMR-RESs may be economically
sub-optimal per see, their GHG emissions savings may nevertheless contribute greatly
towards decarbonizing the electricity system in Peninsular Malaysia and thus be worthy
of societal support.

171
Reference List

Abdmouleh, Z., Alammari, R. A. M. and Gastli, A. (2015) ‘Review of policies encouraging renewable
energy integration’, Renewable and Sustainable Energy Reviews. Elsevier, 45, pp. 249–262. doi:
10.1016/j.rser.2015.01.035.
Abdul Malek, A. B. . et al. (2017) ‘Techno-economic analysis and environmental impact assessment
of a 10 MW biomass-based power plant in Malaysia’, Journal of Cleaner Production. Elsevier Ltd,
141, pp. 502–513. doi: 10.1016/j.jclepro.2016.09.057.
Abdullah, N. and Sulaiman, F. (2013) ‘The oil palm wastes in Malaysia’, Biomass Now – Sustainable
Growth and Use, pp. 75–100. doi: 10.5772/55302.
Abdullah, N. and Sulaiman, F. (2013) ‘The Oil Palm Wastes in Malaysia’, Biomass Now - Sustainable
Growth and Use. doi: 10.5772/55302.
Abdullah, N., Sulaiman, F. and Gerhauser, H. (2011) ‘Characterisation of oil palm empty fruit
bunches for fuel application’, Journal of Physical Science, 22(1), pp. 1–24.
Abu Bakar, N. (2015) ‘Country Presentation on Status of Bioenergy Development In Malaysia’, p. 14.
Available at:
https://www.iea.org/media/technologyplatform/workshops/southeastasiabioenergy2014/Malaysia.pdf.
ACE (2016) Levelised Cost of Electricity of Selected Renewable Technologies in the ASEAN Member
States. Available at: http://cloud.aseanenergy.org/s/1AK7OzwGCHn5iAM#pdfviewer.
Afriyanti, D., Kroeze, C. and Saad, A. (2016) ‘Indonesia palm oil production without deforestation
and peat conversion by 2050’, Science of the Total Environment. Elsevier B.V., 557–558, pp. 562–
570. doi: 10.1016/j.scitotenv.2016.03.032.
AGECC (2010) Energy for a Sustainable Future, The Secretary-General’s Advisory Group on Energy
and Climate Change. doi: DOI: 10.1016/j.enbuild.2006.04.011.
Aghamohammadi, N. et al. (2016) ‘An Investigation of Sustainable Power Generation from Oil Palm
Biomass: A Case Study in Sarawak’, Sustainability. Multidisciplinary Digital Publishing Institute,
8(5), p. 416. doi: 10.3390/su8050416.
Ahlgren, S. and Di Lucia, L. (2014) ‘Indirect land use changes of biofuel production - A review of
modelling efforts and policy developments in the European Union’, Biotechnology for Biofuels.
Biotechnology for Biofuels, 7(1), pp. 1–10. doi: 10.1186/1754-6834-7-35.
Ahmad, S., Ab Kadir, M. Z. A. and Shafie, S. (2011) ‘Current perspective of the renewable energy
development in Malaysia’, Renewable and Sustainable Energy Reviews. Elsevier Ltd, 15(2), pp. 897–
904. doi: 10.1016/j.rser.2010.11.009.
Ahmad, S. and Tahar, R. M. (2014) ‘Selection of renewable energy sources for sustainable
development of electricity generation system using analytic hierarchy process: A case of Malaysia’,
Renewable Energy. Elsevier Ltd, 63, pp. 458–466. doi: 10.1016/j.renene.2013.10.001.
Akikur, R. K. et al. (2013) ‘Comparative study of stand-alone and hybrid solar energy systems
suitable for off-grid rural electri fi cation : A review’, Renewable and Sustainable Energy Reviews.
Elsevier, 27, pp. 738–752. doi: 10.1016/j.rser.2013.06.043.
Alam, S. S. et al. (2016) ‘A Survey on Renewable Energy Development in Malaysia: Current Status,
Problems and Prospects’, Environmental and Climate Technologies, 17(1), pp. 5–17. doi:
10.1515/rtuect-2016-0002.
172
Al-Amir, J. and Abu-Hijleh, B. (2013) ‘Strategies and policies from promoting the use of renewable
energy resource in the UAE’, Renewable and Sustainable Energy Reviews. Elsevier, 26, pp. 660–667.
doi: 10.1016/j.rser.2013.06.001.
Ali, A. A. M. et al. (2015) ‘Sustainable and integrated palm oil biorefinery concept with value-
addition of biomass and zero emission system’, Journal of Cleaner Production. Elsevier Ltd, 91, pp.
96–99. doi: 10.1016/j.jclepro.2014.12.030.
Ali, R., Daut, I. and Taib, S. (2012) ‘A review on existing and future energy sources for electrical
power generation in Malaysia’, Renewable and Sustainable Energy Reviews. Elsevier Ltd, 16(6), pp.
4047–4055. doi: 10.1016/j.rser.2012.03.003.
Ali, S. H., Tesfamichael, A. and Hassan, S. (2014) ‘Effect of Low Pressure End Conditions on Steam
Power Plant Performance’, MATEC Web of Conferences, 13(2010), pp. 1–6. doi: 10.1051/
matecconf/201 13 4 02010.
Alias, N. B., Ibrahim, N. and Hamid, M. K. A. (2014) ‘Pyrolysis of empty fruit bunch by
thermogravimetric analysis’, Energy Procedia. Elsevier B.V., 61, pp. 2532–2536. doi:
10.1016/j.egypro.2014.12.039.
Alves, M. et al. (2015) ‘Surplus electricity production in sugarcane mills using residual bagasse and
straw as fuel’, Energy. Elsevier Ltd, 91, pp. 751–757. doi: 10.1016/j.energy.2015.08.101.
Amigun, B. and Von Blottnitz, H. (2009) ‘Capital cost prediction for biogas installations in Africa:
Lang factor approach’, Environmental Progress and Sustainable Energy, pp. 134–142. doi:
10.1002/ep.10341.
Amos, W. A. (1998) Report on Biomass Drying Technology Report on Biomass Drying Technology,
National Renewable Energy Laboratory. doi: NREL/TP-570-25885.
Andiappan, V., Ng, D. K. S. and Bandyopadhyay, S. (2014) ‘Synthesis of Biomass-based
Trigeneration Systems with Uncertainties’, Industrial & Engineering Chemistry Research, 53, pp.
18016–18028. doi: 10.1016/B978-0-444-63576-1.50068-6.
Ansori, M., Herawan, T. and Rivani, M. (2014) ‘Analysis of Palm Biomass as Electricity from Palm
Oil Mills in North Sumatera’, Energy Procedia. Elsevier B.V., 47, pp. 166–172. doi:
10.1016/j.egypro.2014.01.210.
Ansori, M., Herawan, T. and Rivani, M. (2015) ‘Analysis of Palm Biomass as Electricity from Palm
Oil Mills in North Sumatera Analysis of Palm Biomass as Electricity from Palm Oil Mills in North
Sumatera’, Energy Procedia. Elsevier B.V., 47(January), pp. 166–172. doi:
10.1016/j.egypro.2014.01.210.
Archer, S. A., Murphy, R. J. and Steinberger-Wilckens, R. (2018) ‘Methodological analysis of palm
oil biodiesel life cycle studies’, Renewable and Sustainable Energy Reviews. Elsevier Ltd, 94(August
2017), pp. 694–704. doi: 10.1016/j.rser.2018.05.066.
Ardente, F. and Cellura, M. (2012) ‘Economic Allocation in Life Cycle Assessment: The State of the
Art and Discussion of Examples’, Journal of Industrial Ecology, 16(3), pp. 387–398. doi:
10.1111/j.1530-9290.2011.00434.x.
Asdrubali, F. et al. (2015) ‘Life cycle assessment of electricity production from renewable energies:
Review and results harmonization’, Renewable and Sustainable Energy Reviews. Elsevier, 42, pp.
1113–1122. doi: 10.1016/j.rser.2014.10.082.
Asia-Pacific Economic Cooperation (2014) Peer Review on Low Carbon Energy Policies in Malaysia,
Asia-Pacific Economic Cooperation.
173
Aziz, M. et al. (2017) ‘Advanced power generation using biomass wastes from palm oil mills’,
Applied Thermal Engineering. Elsevier Ltd, 114, pp. 1378–1386. doi:
10.1016/j.applthermaleng.2016.11.031.
Aziz, M., Oda, T. and Kashiwagi, T. (2015) ‘Innovative Steam Drying of Empty Fruit Bunch with
High Energy Efficiency’, Drying Technology. 2015, 33(4), pp. 395–405. doi:
10.1080/07373937.2014.970257.
Ba, B. H., Prins, C. and Prodhon, C. (2016) ‘Models for optimization and performance evaluation of
biomass supply chains: An Operations Research perspective’, Renewable Energy, 87, pp. 977–989.
doi: 10.1016/j.renene.2015.07.045.
Bakhtyar, B. et al. (2013) ‘Renewable energy in five South East Asian countries: Review on
electricity consumption and economic growth’, Renewable and Sustainable Energy Reviews. Elsevier,
26, pp. 506–514. doi: 10.1016/j.rser.2013.05.058.
Bakos, G. C., Tsioliaridou, E. and Potolias, C. (2008) ‘Technoeconomic assessment and strategic
analysis of heat and power co-generation (CHP) from biomass in Greece’, Biomass and Bioenergy,
32(6), pp. 558–567. doi: 10.1016/j.biombioe.2007.11.014.
Baños, R. et al. (2011) ‘Optimization methods applied to renewable and sustainable energy: A
review’, Renewable and Sustainable Energy Reviews, 15(4), pp. 1753–1766. doi:
10.1016/j.rser.2010.12.008.
Bartnik, R., Buryn, Z. and Hnydiuk-Stefan, A. (2016) ‘Thermodynamic analysis of annual operation
of a 370 MW power unit operating in CHP-mode’, Applied Thermal Engineering. Elsevier Ltd, 106,
pp. 42–48. doi: 10.1016/j.applthermaleng.2015.10.165.
Basiron, Y. and Weng, C. K. (2004) ‘The Oil Palm and its Sustainability’, Journal of Oil Palm
Research, 16(1), pp. 1–10.
Basri, N. A., Ramli, A. T. and Aliyu, A. S. (2015) ‘Malaysia energy strategy towards sustainability: A
panoramic overview of the benefits and challenges’, Renewable and Sustainable Energy Reviews, 42,
pp. 1094–1105. doi: 10.1016/j.rser.2014.10.056.
Bates, J., Edberg, O. and Nuttall, C. (2009) Minimising greenhouse gas emissions from biomass
energy generation. Available at:
http://www.globalbioenergy.org/uploads/media/0904_Environment_Agency_-
_Minimising_greenhouse_gas_emissions_from_biomass_energy_generation.pdf.
Batidzirai, B. et al. (2016) Biomass Supply and Trade Opportunities of Preprocessed Biomass for
Power Generation, Developing the Global Bioeconomy: Technical, Market, and Environmental
Lessons from Bioenergy. Elsevier Inc. doi: 10.1016/B978-0-12-805165-8.00005-7.
Beaudry, G. et al. (2017) ‘Greenhouse gas assessment of palm oil mill biorefinery in Thailand from a
life cycle perspective’, Biomass Conversion and Biorefinery. Biomass Conversion and Biorefinery.
doi: 10.1007/s13399-016-0233-7.
Beér, J. M. (2000) ‘Combustion technology developments in power generation in response to
environmental challenges’, Progress in Energy and Combustion Science, 26(4–6), pp. 301–327. doi:
10.1016/S0360-1285(00)00007-1.
Black, M. J. et al. (2011) ‘Life Cycle Assessment and sustainability methodologies for assessing
industrial crops, processes and end products’, Industrial Crops and Products, 34(2), pp. 1332–1339.
doi: 10.1016/j.indcrop.2010.12.002.
Böhringer, C. and Lange, A. (2005) ‘On the design of optimal grandfathering schemes for emission
174
allowances’, European Economic Review, 49(8), pp. 2041–2055. doi:
10.1016/j.euroecorev.2004.06.006.
Borgnakke, C. and Sonntag, R. E. (2013) Fundamentals of Thermodynamics. doi: 10.1016/B978-0-
12-383860-5.00001-8.
Borhanazad, H. et al. (2013) ‘Potential application of renewable energy for rural electrification in
Malaysia’, Renewable Energy. Elsevier Ltd, 59, pp. 210–219. doi: 10.1016/j.renene.2013.03.039.
Borrion, A. L., McManus, M. C. and Hammond, G. P. (2012) ‘Environmental life cycle assessment of
lignocellulosic conversion to ethanol: A review’, Renewable and Sustainable Energy Reviews.
Elsevier, 16(7), pp. 4638–4650. doi: 10.1016/j.rser.2012.04.016.
Brammer, J. G. and Bridgwater, A. V. (2002) ‘The influence of feedstock drying on the performance
and economics of a biomass gasifier - Engine CHP system’, Biomass and Bioenergy, 22(4), pp. 271–
281. doi: 10.1016/S0961-9534(02)00003-X.
Brander, M. et al. (2008) Consequential and Attributional Approches to LCA: a Guide to Policy
Makers with Specific Reference to Greenhouse Gas LCA of Biofuels. Available at:
http://www.globalbioenergy.org/uploads/media/0804_Ecometrica_-
_Consequential_and_attributional_approaches_to_LCA.pdf.
British Standards Institution (2011) PUBLICLY AVAILABLE SPECIFICATION PAS 2050: 2011
Specification for the assessment of the life cycle greenhouse gas emissions of goods and services. doi:
978 0 580 71382 8.
Buytaert, V. et al. (2011) ‘Towards integrated sustainability assessment for energetic use of biomass:
A state of the art evaluation of assessment tools’, Renewable and Sustainable Energy Reviews.
Elsevier Ltd, 15(8), pp. 3918–3933. doi: 10.1016/j.rser.2011.07.036.
Çakir, U., Çomakli, K. and Yüksel, F. (2012) ‘The role of cogeneration systems in sustainability of
energy’, Energy Conversion and Management, pp. 196–202. doi: 10.1016/j.enconman.2012.01.041.
Calvete, H. I. and Gal, C. (2010) ‘Multiple Criteria Decision Making for Sustainable Energy and
Transportation Systems’, Multiple Criteria Decision Making for Sustainable Energy and
Transportation Systems, 634, pp. 155–165. doi: 10.1007/978-3-642-04045-0.
Caputo, A. C. et al. (2005) ‘Economics of biomass energy utilization in combustion and gasification
plants : effects of logistic variables’, Biomass & Bioenergy, 28, pp. 35–51. doi:
10.1016/j.biombioe.2004.04.009.
Cattani, K. and Schmidt, G. M. (2005) ‘The Pooling Principle’, INFORMS Transactions on
Education, 5(2), pp. 17–24. doi: 10.1287/ited.5.2.17.
Cespi, D. et al. (2014) ‘Heating systems LCA: Comparison of biomass-based appliances’,
International Journal of Life Cycle Assessment, 19(1), pp. 89–99. doi: 10.1007/s11367-013-0611-3.
Chang, S. H. (2014) ‘An overview of empty fruit bunch from oil palm as feedstock for bio-oil
production’, Biomass and Bioenergy. Elsevier Ltd, 62, pp. 174–181. doi:
10.1016/j.biombioe.2014.01.002.
Chanlongphitak, S. et al. (2015) ‘Life Cycle Assessment of Palm Empty Fruit Bunch Utilization for
Power Plants in Thailand’, in International Conference on Biological, Environment and Food
Engineering (BEFE), Singapore, pp. 32–37.
Chau, J. et al. (2009) ‘Techno-economic analysis of wood biomass boilers for the greenhouse
industry’, Applied Energy. Elsevier Ltd, 86(3), pp. 364–371. doi: 10.1016/j.apenergy.2008.05.010.

175
Chen, D. et al. (2017) ‘An approach for upgrading biomass and pyrolysis product quality using a
combination of aqueous phase bio-oil washing and torrefaction pretreatment’, Bioresource
Technology, 233, pp. 150–158. Available at: http://dx.doi.org/10.1016/j.biortech.2017.02.120.
Chen, S. S. (2008) ‘International Palm Oil Sustainablility Conference 2008 Kota Kinabalu , 14 th
April 2008 Environmental Impacts Associated with A Product ’ s’, (April).
Cheng, S. L., Chang, C. T. and Jiang, D. (2014) ‘A game-theory based optimization strategy to
configure inter-plant heat integration schemes’, Chemical Engineering Science. Elsevier, 118, pp. 60–
73. doi: 10.1016/j.ces.2014.07.001.
Chiew, Y. L., Iwata, T. and Shimada, S. (2011) ‘System analysis for effective use of palm oil waste as
energy resources’, Biomass and Bioenergy. Elsevier Ltd, 35(7), pp. 2925–2935. doi:
10.1016/j.biombioe.2011.03.027.
Chiew, Y. L. and Shimada, S. (2013) ‘Current state and environmental impact assessment for utilizing
oil palm empty fruit bunches for fuel , fiber and fertilizer e A case study of Malaysia’, Biomass and
Bioenergy. Elsevier Ltd, 51, pp. 109–124. doi: 10.1016/j.biombioe.2013.01.012.
Chin, M. J. et al. (2013) ‘Biogas from palm oil mill effluent (POME): Opportunities and challenges
from Malaysia’s perspective’, Renewable and Sustainable Energy Reviews, 26, pp. 717–726. doi:
10.1016/j.rser.2013.06.008.
Choo, Y. M. et al. (2011) ‘Determination of GHG contributions by subsystems in the oil palm supply
chain using the LCA approach’, International Journal of Life Cycle Assessment, 16(7), pp. 669–681.
doi: 10.1007/s11367-011-0303-9.
Chua, S. C. and Oh, T. H. (2010) ‘Review on Malaysia’s national energy developments: Key policies,
agencies, programmes and international involvements’, Renewable and Sustainable Energy Reviews.
Elsevier Ltd, 14, pp. 2916–2925. doi: 10.1016/j.rser.2010.07.031.
Chungsangunsit, T., Gheewala, S. H. and Patumsawad, S. (2005) ‘Environmental Assessment of
Electricity Production from Rice Husk : A Case Study in Thailand’, International Energy Journal,
6(1), pp. 47–55.
Clean Development Mechnism (2006) CDM Project Design Document Form for Felda Serting Hilir.
Codina Gironès, V. et al. (2017) ‘Optimal use of biomass in large-scale energy systems: Insights for
energy policy’, Energy, pp. 1–9. doi: 10.1016/j.energy.2017.05.027.
Couture, T. D., Cory, K. and Williams, E. (2010) A Policymaker’s Guide to Feed-in Tariff Policy
Design. Available at: https://www.nrel.gov/docs/fy10osti/44849.pdf.
Van Dael, M. et al. (2013) ‘A techno-economic evaluation of a biomass energy conversion park’,
Applied Energy, 104, pp. 611–622. doi: 10.1016/j.apenergy.2012.11.071.
Dantas, G. A., Legey, L. F. L. and Mazzone, A. (2013) ‘Energy from sugarcane bagasse in Brazil: An
assessment of the productivity and cost of different technological routes’, Renewable and Sustainable
Energy Reviews. Elsevier, 21, pp. 356–364. doi: 10.1016/j.rser.2012.11.080.
Darshini, D., Dwivedi, P. and Glenk, K. (2013) ‘Capturing Stakeholders’ views in oil palm-based
biofuel and biomass utilization in Malaysia’, Energy Policy, 62, pp. 1128–1137.
Demirbas, A. (2005) ‘Potential applications of renewable energy sources, biomass combustion
problems in boiler power systems and combustion related environmental issues’, Progress in Energy
and Combustion Science, 31(2), pp. 171–192. doi: 10.1016/j.pecs.2005.02.002.
Demirbas, M. F., Balat, M. and Balat, H. (2009) ‘Potential contribution of biomass to the sustainable
176
energy development’, Energy Conversion and Management. Elsevier Ltd, 50(7), pp. 1746–1760. doi:
10.1016/j.enconman.2009.03.013.
Department of Environment (1974) Environmental Quality Act 1974. Available at:
https://www.doe.gov.my/portalv1/wp-content/uploads/2015/01/Environmental_Quality_Act_1974_-
_ACT_127.pdf.
Department of Standard Malaysia (2012) Malaysian Standard on Sustainable Palm Oil (MSPO) - MS
2530:2013. doi: MS 692:2007.
Djomo, S. N., Kasmioui, O. El and Ceulemans, R. (2011) ‘Energy and greenhouse gas balance of
bioenergy production from poplar and willow: A review’, GCB Bioenergy, 3(3), pp. 181–197. doi:
10.1111/j.1757-1707.2010.01073.x.
Economic Planning Unit (2015) Strengthening Infrastructure to Support Economic Expansion,
Rancangan Malaysia Kesebelas (Eleventh Malaysia Plan) : 2016-2020. doi:
10.1017/CBO9781107415324.004.
Eleftheriadis, I. M. and Anagnostopoulou, E. G. (2015) ‘Identifying barriers in the diffusion of
renewable energy sources’, Energy Policy. Elsevier, 80, pp. 153–164. doi:
10.1016/j.enpol.2015.01.039.
Elmer, U. and Nygaard, I. (2014) ‘Sustainable energy transitions in emerging economies : The
formation of a palm oil biomass waste-to-energy niche in Malaysia 1990 – 2011’, Energy Policy.
Elsevier, 66, pp. 666–676. doi: 10.1016/j.enpol.2013.11.028.
Energy Commision (2013) Peninsular Malaysia Electricity Supply Industry Outlook 2013. Available
at: http://www.st.gov.my/.
Energy Commission (2010) The Malaysian Grid Code.
Energy Commission (2016) Malaysia Energy Statistics Handbook 2016.
Energy Commission (2017) Peninsular Malaysia Electricity Supply Outlook 2017. Available at:
http://www.st.gov.my/index.php/en/all-publications/item/765-peninsular-malaysia-electricity-supply-
industry-outlook-2017.
EPA (2003) Wood Residue Combustion In Boilers. Available at:
http://www.epa.gov/ttn/chief/ap42/ch01/index.html.
Eurelectric (2003) Efficiency in Electricity Generation, Union of the Electricity Industry -
EURELECTRIC. doi: 10.1016/j.watres.2009.01.036.
European Commission (2014) Guide to Cost-benefit Analysis of Investment Projects: Economic
appraisal tool for Cohesion Policy 2014-2020, Publications Office of the European Union. doi:
10.2776/97516.
Evans, A., Strezov, V. and Evans, T. J. (2009) ‘Assessment of sustainability indicators for renewable
energy technologies’, Renewable and Sustainable Energy Reviews, 13(5), pp. 1082–1088. doi:
10.1016/j.rser.2008.03.008.
Faaij, A. (2006) Modern biomass conversion technologies, Mitigation and Adaptation Strategies for
Global Change. doi: 10.1007/s11027-005-9004-7.
El Fadel, M. et al. (2013) ‘Knowledge management mapping and gap analysis in renewable energy:
Towards a sustainable framework in developing countries’, Renewable and Sustainable Energy
Reviews. Elsevier, 20, pp. 576–584. doi: 10.1016/j.rser.2012.11.071.

177
Fagernäs, L. et al. (2010) ‘Drying of biomass for second generation synfuel production’, Biomass and
Bioenergy, 34(9), pp. 1267–1277. doi: 10.1016/j.biombioe.2010.04.005.
Fallis, A. . (2013) ‘Converting Waste Oil Palm Trees Into a Resource’, Journal of Chemical
Information and Modeling, 53(9), pp. 1689–1699. doi: 10.1017/CBO9781107415324.004.
Fiala, M., Pellizzi, G. and Riva, G. (1997) ‘A Model for the Optimal Dimensioning of Biomass-
fuelled Electric Power Plants’, Journal of Agricultural Engineering Research, 67(1), pp. 17–25. doi:
10.1006/jaer.1996.0142.
Foo, D. C. Y. et al. (2013) ‘Robust models for the synthesis of flexible palm oil-based regional
bioenergy supply chain’, Energy, 55, pp. 68–73. doi: 10.1016/j.energy.2013.01.045.
Freppaz, D. et al. (2004) ‘Optimizing forest biomass exploitation for energy supply at a regional
level’, Biomass and Bioenergy, 26(1), pp. 15–25. doi: 10.1016/S0961-9534(03)00079-5.
Frischknecht, R. (2000) ‘Allocation in Life Cycle Inventory Analysis for Joint Production’, The MIIM
LCA Ph.D. Club, 179(1999), pp. 85–95. doi: 10.1065/lca2000.02.013.
Gallagher, G. (2001) ‘Biomass for electricity generation’, Chemical Engineer, (725), pp. 32–33.
Ganapathy, V. (1996) ‘Heat-recovery steam generators: Understand the basics’, Chemical
Engineering Progress, 92(8), pp. 32–45.
Gebreegziabher, T. et al. (2014) ‘Design and optimization of biomass power plant’, Chemical
Engineering Research and Design. Institution of Chemical Engineers, 92(8), pp. 1412–1427. doi:
10.1016/j.cherd.2014.04.013.
Gebreegziabher, T., Olajire, A. and Yu, Z. (2014) Biomass Drying for an Integrated Power Plant :
Effective Utilization of Waste Heat, 24 European Symposium on Computer Aided Process
Engineering. Elsevier. doi: 10.1016/B978-0-444-63455-9.50094-5.
Gebreegziabher, T., Oyedun, A. O. and Hui, C. W. (2013) ‘Optimum biomass drying for combustion -
A modeling approach’, Energy. Elsevier Ltd, 53, pp. 67–73. doi: 10.1016/j.energy.2013.03.004.
Ghaderi, H., Pishvaee, M. S. and Moini, A. (2016) ‘Biomass supply chain network design: An
optimization-oriented review and analysis’, Industrial Crops and Products. Elsevier B.V., 94, pp.
972–1000. doi: 10.1016/j.indcrop.2016.09.027.
Goedkoop, M. et al. (2016) Introduction to LCA with SimaPro. Available at: https://www.pre-
sustainability.com/download/SimaPro8IntroductionToLCA.pdf.
Gómez, D. R. et al. (2006) Stationary Combustion, 2006 IPCC Guidelines for National Greenhouse
Gas Inventories. doi: 10.1016/S0166-526X(06)47021-5.
González-García, S. and Bacenetti, J. (2018) ‘Exploring the production of bio-energy from wood
biomass. Italian case study’, Science of the Total Environment. Elsevier B.V., 647, pp. 158–168. doi:
10.1016/j.scitotenv.2018.07.295.
Gruning, C. et al. (2012) National Climate Finance Institutions Support Programme Case Study : The
Thai Energy Efficiency Revolving Fund.
Guinée, J., Heijungs, R. and Huppes, G. (2004) ‘Economic allocaton: Examples and derived decission
tree’, The International Journal of Life Cycle Assessment, 9(1), pp. 23–33.
Gunarso, P. et al. (2013) Oil Palm and Land Use Change in Indonesia, Malaysia and Papua New
Guinea.
Haas, R. et al. (2011) ‘A historical review of promotion strategies for electricity from renewable
178
energy sources in EU countries’, Renewable and Sustainable Energy Reviews. Elsevier Ltd, 15(2), pp.
1003–1034. doi: 10.1016/j.rser.2010.11.015.
Hamelinck, C. N., Suurs, R. A. A. and Faaij, A. P. C. (2005) ‘International bioenergy transport costs
and energy balance’, Biomass and Bioenergy, 29(2), pp. 114–134. doi:
10.1016/j.biombioe.2005.04.002.
Hansen, S. B., Olsen, S. I. and Ujang, Z. (2012) ‘Greenhouse gas reductions through enhanced use of
residues in the life cycle of Malaysian palm oil derived biodiesel’, Bioresource Technology, 104, pp.
358–366. doi: 10.1016/j.biortech.2011.10.069.
Haque, N. and Somerville, M. (2013) ‘Techno-economic and environmental evaluation of biomass
dryer’, Procedia Engineering. Elsevier B.V., 56, pp. 650–655. doi: 10.1016/j.proeng.2013.03.173.
Hasanudin, U. et al. (2015) ‘Palm oil mill effluent treatment and utilization to ensure the
sustainability of palm oil industries’, Water Science and Technology, 72(7), pp. 1089–1095. doi:
10.2166/wst.2015.311.
Hashim, H. and Ho, W. S. (2011) ‘Renewable energy policies and initiatives for a sustainable energy
future in Malaysia’, Renewable and Sustainable Energy Reviews. Elsevier Ltd, 15(9), pp. 4780–4787.
doi: 10.1016/j.rser.2011.07.073.
Hassan, M. Y. et al. (2008) ‘Electricity market models in restructured electricity supply industry’,
2008 IEEE 2nd International Power and Energy Conference, pp. 1038–1042. doi:
10.1109/PECON.2008.4762618.
Hassan, M. Y. Bin (2009) ‘A Study of Electricity Market Models in The Restructured Electricity
Supply Industry (Kajian Terhadap Beberapa Model Pasaran Elektrik Di Dalam Penstrukturan Semula
Industri Bekalan Elektrik)’, p. 168.
Hatch, M. (2010) ‘The Role of Renewable Energy in German Climate Change Policy’, Renewable
Energy Law and Policy : RELP, 1(2), pp. 141–151. Available at:
http://ezproxy.qa.proquest.com/docview/822740009?accountid=14771\nhttp://bf4dv7zn3u.search.seri
alssolutions.com/?SS_Source=3&amp\ngenre=article&amp\nsid=ProQ:&amp\natitle=The+Role+of+
Renewable+Energy+in+German+Climate+Change+Policy&amp\ntitle=Renewable+En.
Havukainen, J. et al. (2018) ‘Life cycle assessment of small-scale combined heat and power plant:
Environmental impacts of different forest biofuels and replacing district heat produced from natural
gas’, Journal of Cleaner Production, 172, pp. 837–846. doi: 10.1016/j.jclepro.2017.10.241.
Hennig, C., Brosowski, A. and Majer, S. (2016) ‘Sustainable feedstock potential - A limitation for the
bio-based economy?’, Journal of Cleaner Production, pp. 200–202. doi:
10.1016/j.jclepro.2015.06.130.
Henrique, M. et al. (2014) ‘Life cycle assessment ( LCA ) for biofuels in Brazilian conditions : A
meta-analysis’, Renewable and Sustainable Energy Reviews. Elsevier, 37, pp. 435–459. doi:
10.1016/j.rser.2014.05.036.
Hiroux, C. and Saguan, M. (2010) ‘Large-scale wind power in European electricity markets: Time for
revisiting support schemes and market designs?’, Energy Policy, pp. 3135–3145. doi:
10.1016/j.enpol.2009.07.030.
Hoefnagels, R., Junginger, M. and Faaij, A. (2012) Capacity Study for Solid Biomass Facilities :
Scenarios for supply and demand of solid biomass for electricity and heat generation in North West
Europe.
Hoeltinger, S., Schmidt, J. and Schmid, E. (2012) ‘The green biorefinery concept: Optimal plant
179
locations and sizes for Austria ’, Journal of the Austrian Society of Agricultural Economics, 21(1), pp.
147–156. Available at: http://www.scopus.com/inward/record.url?eid=2-s2.0-
84872361126&partnerID=40&md5=0f172c0ea62306776e96be8c3db71ed7.
Holmberg, H., Ruohonen, P. and Ahtila, P. (2009) ‘Determination of the real loss of power for a
condensing and a backpressure turbine by means of second law analysis’, Entropy, 11(4), pp. 702–
712. doi: 10.3390/e11040702.
Hosseini, S. E. and Abdul Wahid, M. (2014a) ‘Development of biogas combustion in combined heat
and power generation’, Renewable and Sustainable Energy Reviews. Elsevier, 40, pp. 868–875. doi:
10.1016/j.rser.2014.07.204.
Hosseini, S. E. and Abdul Wahid, M. (2014b) ‘Utilization of palm solid residue as a source of
renewable and sustainable energy in Malaysia’, Renewable and Sustainable Energy Reviews. Elsevier,
40, pp. 621–632. doi: 10.1016/j.rser.2014.07.214.
Huijbregts, M. A. J. et al. (2016) ReCiPe 2016: A harmonized life cycle impact assessment method at
midpoint and enpoint level - Report 1 : characterization, National Institute for Public Health and the
Environment. doi: 10.1007/s11367-016-1246-y.
Hunpinyo, P. et al. (2013) ‘Evaluation of Techno-economic feasibility Biomass-to-energy by Using
ASPEN Plus®: A Case Study of Thailand’, Energy Procedia. Elsevier B.V., 42(August), pp. 640–
649. doi: 10.1016/j.egypro.2013.11.066.
Hussain, A. et al. (2006) ‘Thermochemical behaviour of empty fruit bunches and oil palm shell waste
in a circulating fluidized-bed combustor (CFBC)’, Journal of Oil Palm Research, 18, pp. 210–218.
Available at: http://palmoilis.mpob.gov.my/publications/joprv18june-hussain.pdf.
IEA (2015) Energy and Climate change, International Energy Agency. doi: 10.1038/479267b.
Iguaz, A. et al. (2003) ‘Mathematical modelling and simulation for the drying process of vegetable
wholesale by-products in a rotary dryer’, Journal of Food Engineering, 59(2–3), pp. 151–160. doi:
10.1016/S0260-8774(02)00451-X.
Institute of Strategic and International Study (2014) Reforming Peninsular Malaysia ’ s Electricity
Sector Challenges and Prospects.
International Energy Agency (2017) Key World Energy Statistics. Available at:
http://www.iea.org/statistics/.
IPCC (2006) Solid Waste Disposal, 2006 IPCC Guidelines for National Greenhouse Gas Inventories
Volume 5: Waste. Available at: http://www.ipcc-
nggip.iges.or.jp/public/2006gl/pdf/5_Volume5/V5_3_Ch3_SWDS.pdf.
IRENA (2012) Biomass for Power Generation, International Renewable Energy Agency (IRENA)
Copyright IRENA 2012.
IRENA (2018) Renewable Power Generation Costs in 2017, Springer Reference. doi:
10.1007/SpringerReference_7300.
Jaafar, M. Z., Kheng, W. H. and Kamaruddin, N. (2003) ‘Greener energy solutions for a sustainable
future: Issues and challenges for Malaysia’, Energy Policy, 31(11), pp. 1061–1072. doi:
10.1016/S0301-4215(02)00216-1.
Jenkins, B. . et al. (1998) ‘Combustion properties of biomass’, Fuel Processing Technology, 54(1–3),
pp. 17–46. doi: 10.1016/S0378-3820(97)00059-3.
Jenkins, B. M. (1997) ‘A comment on the optimal sizing of a biomass utilization facility under
180
constant and variable cost scaling’, Biomass and Bioenergy, 13(1–2), pp. 1–9. doi: 10.1016/S0961-
9534(97)00085-8.
Jirjis, R. (1995) ‘Storage and drying of wood fuel’, Biomass and Bioenergy, 9(1–5), pp. 181–190. doi:
10.1016/0961-9534(95)00090-9.
Johnson, E. (2009) ‘Goodbye to carbon neutral: Getting biomass footprints right’, Environmental
Impact Assessment Review. Elsevier Inc., 29(3), pp. 165–168. doi: 10.1016/j.eiar.2008.11.002.
Johnson, R. L. et al. (2009) ‘Pyrolysis gas chromatography mass spectrometry studies to evaluate
high-temperature aqueous pretreatment as a way to modify the composition of bio-oil from fast
pyrolysis of wheat straw’, Energy and Fuels, 23(12), pp. 6242–6252. doi: 10.1021/ef900742x.
Jorgenson, J., Gilman, P. and Dobos, A. (2011) Technical Manual for the SAM Biomass Power
Generation Model. Available at: http://purl.fdlp.gov/GPO/gpo15280.
Jue, E., Johnson, B. and Vanamali, A. (2012) Case study: Thailand’s Energy Conservation (ENCON)
Fund - How Financial Mechanisms Catalyzed Energy Efficiency and Renewable Energy Investments.
Available at: http://ccap.org/assets/Thailand-Energy-Conservation-ENCON-Fund_CCAP-Oct-
2012.pdf.
Kadiyala, A., Kommalapati, R. and Huque, Z. (2016) ‘Evaluation of the Life Cycle Greenhouse Gas
Emissions from Different Biomass Feedstock Electricity Generation Systems’, Sustainability, 8(11),
p. 1181. doi: 10.3390/su8111181.
Kampa, M. and Castanas, E. (2008) ‘Human health effects of air pollution’, Environmental Pollution,
151(2), pp. 362–367. doi: 10.1016/j.envpol.2007.06.012.
Kardooni, R., Yusoff, S. and Kari, F. (2016) ‘Renewable energy technology acceptance in Peninsular
Malaysia’, Energy Policy. Elsevier, 88, pp. 1–10. doi: 10.1016/j.enpol.2015.10.005.
van Kasteren, J. M. N. and Nisworo, A. P. (2007) ‘A process model to estimate the cost of industrial
scale biodiesel production from waste cooking oil by supercritical transesterification’, Resources,
Conservation and Recycling, 50(4), pp. 442–458. doi: 10.1016/j.resconrec.2006.07.005.
Kazi, F. K. et al. (2010) ‘Techno-economic comparison of process technologies for biochemical
ethanol production from corn stover’, Fuel. Elsevier Ltd, 89(SUPPL. 1), pp. S20–S28. doi:
10.1016/j.fuel.2010.01.001.
Keoleian, G. A. and Volk, T. A. (2005) ‘Renewable energy from willow biomass crops: Life cycle
energy, environmental and economic performance’, Critical Reviews in Plant Sciences, 24(5–6), pp.
385–406. doi: 10.1080/07352680500316334.
KeTTHA (2008) National Renewable Energy Policy and Action Plan - Malaysia, National Renewable
Energy Policy.
KeTTHA (2017) Green Technology Master Plan Malaysia 2017 - 2030.
Khan, A. A. et al. (2009) ‘Biomass combustion in fluidized bed boilers: Potential problems and
remedies’, Fuel Processing Technology. Elsevier B.V., 90(1), pp. 21–50. doi:
10.1016/j.fuproc.2008.07.012.
Khor, C. S. and Lalchand, G. (2014) ‘A review on sustainable power generation in Malaysia to 2030 :
Historical perspective , current assessment , and future strategies’, Renewable and Sustainable Energy
Reviews. Elsevier, 29, pp. 952–960. doi: 10.1016/j.rser.2013.08.010.
Kim, H. K. and Lee, H. K. (2011) ‘Use of power plant bottom ash as fine and coarse aggregates in
high-strength concrete’, Construction and Building Materials. Elsevier Ltd, 25(2), pp. 1115–1122.
181
doi: 10.1016/j.conbuildmat.2010.06.065.
Kim, H., Kim, S. and Dale, B. E. (2009) ‘Gas Emissions : Some Unexplored Variables Biofuels ,
Land Use Change , and Greenhouse Gas Emissions : Some Unexplored Variables’, Environmental
Science & Technology, pp. 961–967. doi: 10.1021/es802681k.
Kim, J., Sen, S. M. and Maravelias, C. T. (2013) ‘An optimization-based assessment framework for
biomass-to-fuel conversion strategies’, Energy & Environmental Science, 6(4), p. 1093. doi:
10.1039/c3ee24243a.
Kimming, M. et al. (2011) ‘Biomass from agriculture in small-scale combined heat and power plants -
A comparative life cycle assessment’, Biomass and Bioenergy, 35(4), pp. 1572–1581. doi:
10.1016/j.biombioe.2010.12.027.
Kuprianov, V. I., Kaewklum, R. and Chakritthakul, S. (2011) ‘Effects of operating conditions and fuel
properties on emission performance and combustion efficiency of a swirling fluidized-bed combustor
fired with a biomass fuel’, Energy. Elsevier Ltd, 36(4), pp. 2038–2048. doi:
10.1016/j.energy.2010.05.026.
Lam, H. L. et al. (2013) ‘Green strategy for sustainable waste-to-energy supply chain’, Energy.
Elsevier Ltd, 57, pp. 4–16. doi: 10.1016/j.energy.2013.01.032.
Lauven, L. P., Liu, B. and Geldermann, J. (2013) ‘Investigation of the influence of plant capacity on
the economic and ecological performance of cassava-based bioethanol’, Chemical Engineering
Transactions, 35, pp. 595–600. doi: 10.3303/CET1335099.
Leboreiro, J. and Hilaly, A. K. (2011) ‘Biomass transportation model and optimum plant size for the
production of ethanol’, Bioresource Technology. Elsevier Ltd, 102(3), pp. 2712–2723. doi:
10.1016/j.biortech.2010.10.144.
Lemmens, S. (2015) ‘A perspective on costs and cost estimation techniques for organic Rankine cycle
systems’, Proceedings of the 3rd International Seminar on ORC Power Systems, (2010), pp. 1–10.
Li, H. et al. (2012) ‘Evaluation of a biomass drying process using waste heat from process industries :
A case study’, Applied Thermal Engineering, 35, pp. 71–80. doi:
10.1016/j.applthermaleng.2011.10.009.
Lidula, N. W. A. et al. (2007) ‘ASEAN towards clean and sustainable energy: Potentials, utilization
and barriers’, Renewable Energy, 32(9), pp. 1441–1452. doi: 10.1016/j.renene.2006.07.007.
Lim, C. I. and Biswas, W. (2015) ‘An Evaluation of Holistic Sustainability Assessment Framework
for Palm Oil Production in Malaysia’, Sustainability, 7, pp. 16561–16587. doi: 10.3390/su71215833.
Loh, S. K., Abu Bakar, N., et al. (2017) ‘Biogas Capture – A Means of Reducing Greenhouse Gas
Emissions from Palm Oil Mill Effluent’, Oil Palm Bulletin, pp. 27–36. Available at:
http://palmoilis.mpob.gov.my/publications/OPB/opb75-loh.pdf.
Loh, S. K., Nasrin, A. B., et al. (2017) ‘First Report on Malaysia’s experiences and development in
biogas capture and utilization from palm oil mill effluent under the Economic Transformation
Programme: Current and future perspectives’, Renewable and Sustainable Energy Reviews. Elsevier
Ltd, 74(February), pp. 1257–1274. doi: 10.1016/j.rser.2017.02.066.
Loh, S. K. (2017) ‘The potential of the Malaysian oil palm biomass as a renewable energy source’,
Energy Conversion and Management. Elsevier Ltd, 141, pp. 285–298. doi:
10.1016/j.enconman.2016.08.081.
Loh, W. L. and Abdul Majid, A. (2016) ‘Renewable Energy Law and Policy in Malaysia’, Renewable

182
Energy Law and Policy Review (RELP), 7(1), pp. 85–88.
Loorbach, D. (2010) ‘Transition Management for Sustainable Development: A Prescriptive,
Complexity-Based Governance Framework’, Governance, An International Journal of Policy,
Administration, and Institutions., 23(1), pp. 161–183. doi: 10.1111/j.1468-0491.2009.01471.x.
Luk, H. T. et al. (2013) ‘Drying of biomass for power generation: A case study on power generation
from empty fruit bunch’, Energy. Elsevier Ltd, 63, pp. 205–215. doi: 10.1016/j.energy.2013.10.056.
Lutz, L. M. et al. (2017) ‘Driving factors for the regional implementation of renewable energy ‐ A
multiple case study on the German energy transition’, Energy Policy, pp. 136–147. doi:
10.1016/j.enpol.2017.02.019.
Lynd, L. R. (2010) ‘Bioenergy: in search of clarity’, Energy & Environmental Science, 3(9), pp.
1150–1152. doi: 10.1039/c002335n.
Maali, Y. (2009) ‘A multiobjective approach for solving cooperative n-person games’, International
Journal of Electrical Power and Energy Systems. Elsevier Ltd, 31(10), pp. 608–610. doi:
10.1016/j.ijepes.2009.06.021.
Madaki, Y. S. and Seng, L. (2013) ‘Palm Oil Mill Effluent (POME) From Malaysia Palm Oil Mills :
Waste or Resource’, International Journal of Science, Environment and Technology, 2(6), pp. 1138–
1155.
Mafakheri, F. and Nasiri, F. (2014) ‘Modeling of biomass-to-energy supply chain operations :
Applications , challenges and research directions’, Energy Policy. Elsevier, 67, pp. 116–126. doi:
10.1016/j.enpol.2013.11.071.
Magnusson, H. (2011) ‘Process Simulation in Aspen Plus of an Integrated Ethanol and CHP plant’,
Universitet UMEA, pp. 1–47.
Mahinpey, N. and Gomez, A. (2016) ‘Review of gasification fundamentals and new findings:
Reactors, feedstock, and kinetic studies’, Chemical Engineering Science, 148, pp. 14–31. doi:
10.1016/j.ces.2016.03.037.
Mahlia, T. M. I. et al. (2012) ‘Techno-economic analysis of palm oil mill wastes to generate power
for grid-connected utilization’, Energy Education Science and Technology Part A: Energy Science
and Research, 28(2), pp. 1111–1130.
MALAYSIA, M. O. I. T. A. I. (2016) ‘Miti Report 2016’.
Malaysia Energy Commission (2018) ‘Electricity Tariff Review in Peninsular Malaysia for
Regulatory Period 2 (RP2: 2018-2020) under Incentive-Based Regulation (IBR) Mechanism’, pp. 1–
42. Available at:
http://www.st.gov.my/contents/files/download/161/Electricity_Tariff_Review_in_Peninsular_Msia_fo
r_Regulatory_Period_2_under_IBR_Mechanism_30March2018V2.pdf.
Malaysia Innovation Agency (2013) National Biomass Strategy 2020: New wealth creation for
Malaysia’s palm oil industry, Agensi Inovasi Malaysia, Kuala Lumpur. doi:
10.1016/j.ijggc.2012.07.010.
Martinez, K. K. and Neuhoff, K. (2005) ‘Allocation of carbon emission certificates in the power
sector: How generators profit from grandfathered rights’, Climate Policy, 5(1), pp. 61–78. doi:
10.1080/14693062.2005.9685541.
Martinez-Hernandez, E., Campbell, G. and Sadhukhan, J. (2013) ‘Economic value and environmental
impact (EVEI) analysis of biorefinery systems’, Chemical Engineering Research and Design.

183
Institution of Chemical Engineers, 91(8), pp. 1418–1426. doi: 10.1016/j.cherd.2013.02.025.
Maulud, A. L. and Saidi, H. (2012) ‘The Malaysian Fifth Fuel Policy: Re-strategising the Malaysian
Renewable Energy Initiatives’, Energy Policy. Elsevier, 48, pp. 88–92. doi:
10.1016/j.enpol.2012.06.023.
McIlveen-Wright, D. R. et al. (2013) ‘A technical and economic analysis of three large scale biomass
combustion plants in the UK’, Applied Energy, pp. 396–404. doi: 10.1016/j.apenergy.2012.12.051.
Mckendry, P. (2002) ‘Energy production from biomass ( part 2 ): conversion technologies’, 83(July
2001), pp. 47–54. doi: 10.1016/S0960-8524(01)00119-5.
McKone, T. E. and Hertwich, E. G. (2001) ‘The human toxicity potential and a Strategy for
Evaluating Model Performance in Life Cycle Impact Assessment’, The International Journal of Life
Cycle Assessment, 6(2), pp. 106–109. doi: 10.1007/BF02977846.
Md Jaye, I. F., Sadhukhan, J. and Murphy, R. J. (2016) ‘Renewable, local electricity generation from
palm oil mills: a case study from Peninsular Malaysia’, International Journal of Smart Grid and
Clean Energy, 44, pp. 106–111. doi: 10.12720/sgce.5.2.106-111.
Md Jaye, I. F., Sadhukhan, J. and Murphy, R. J. (2018) ‘Integrated Assessment of Palm Oil Mill
Residues to Sustainable Electricity System ( POMR-SES ): A Case Study from Peninsular Malaysia’,
IOP Conference Series: Materials Sceince and Engineering, 358, pp. 1–7. doi: 10.1088/1757-
899X/358/1/012002.
Meiller, M. et al. (2017) ‘Boiler Design with Solid-Gaseous Fuel Staging to Reduce NOx Emissions
and Optimize Load Flexibility’, Chemical Engineering Technology, 40(2), pp. 289–297. doi:
10.1002/ceat.201600199.
Menon, N. R., Rahman, Z. A. and Bakar, N. A. (2003) ‘Empty Fruit Bunches Evaluation : Mulch in
Plantation vs . Fuel for Electricity Generation’, Oil PalmIndustry Economis Journal, 3 (2)(January
2003), pp. 15–20.
De Meyer, A. et al. (2014) ‘Methods to optimise the design and management of biomass-for-
bioenergy supply chains: A review’, Renewable and Sustainable Energy Reviews. Elsevier, 31, pp.
657–670. doi: 10.1016/j.rser.2013.12.036.
Miao, Z. et al. (2012) ‘Lignocellulosic biomass feedstock transportation alternatives , logistics ,
equipment confi gurations , and modeling’, Biofuels, Bioproducts and Biorefining, 6, pp. 351–362.
doi: 10.1002/bbb.1322.
Miccio, F. et al. (2014) ‘Fluidized Bed Combustion of Wet Biomass Fuel ( Olive Husks )’, Chemical
Engineering Transactions, 37(x), pp. 1–6. doi: 10.3303/CET1437001.
Ministerie van Economie Landbouw & Innovatie (2013) ‘Valorization of palm oil ( mill ) residues’,
(mill), p. 38.
MITI (2014) Ministry of International Trade and Industry Malaysia Report 2014. Available at:
http://www.miti.gov.my/miti/resources/MITI_Report_20141.pdf.
Mohamed, A. R. and Lee, K. T. (2006) ‘Energy for sustainable development in Malaysia: Energy
policy and alternative energy’, Energy Policy, 34(15), pp. 2388–2397. doi:
10.1016/j.enpol.2005.04.003.
Mohd Ghazi, T. I. and Muhammad Nasir, I. (2017) ‘Maximising Potential of Methane Production
from Biogas for Power Generation’, Pertanika Journal Science and Technology, 25(1), pp. 153–160.
Mohd Safaai, N. S. et al. (2010) ‘Projection of CO2 Emissions in Malaysia’, Environmental Progress
184
and Sustainable Energy, pp. 1–8. doi: 10.1002/ep.10512.
Morales, M. et al. (2015) ‘Life cycle assessment of lignocellulosic bioethanol: Environmental impacts
and energy balance’, Renewable and Sustainable Energy Reviews. Elsevier, 42, pp. 1349–1361. doi:
10.1016/j.rser.2014.10.097.
MPOB (2017) Oil Palm Planted Area. Available at:
http://bepi.mpob.gov.my/index.php/en/statistics/area/188-area-2017/856-oil-palm-planted-area-as-at-
dec-2017.html.
Muis, Z. a. et al. (2010) ‘Optimal planning of renewable energy-integrated electricity generation
schemes with CO2 reduction target’, Renewable Energy. Elsevier Ltd, 35(11), pp. 2562–2570. doi:
10.1016/j.renene.2010.03.032.
Mukherjee, I. and Sovacool, B. K. (2014) ‘Palm oil-based biofuels and sustainability in southeast
Asia: A review of Indonesia, Malaysia, and Thailand’, Renewable and Sustainable Energy Reviews.
Elsevier, 37, pp. 1–12. doi: 10.1016/j.rser.2014.05.001.
Mustapa, S. I., Leong, Y. P. and Hashim, A. H. (2010) ‘Issues and challenges of renewable energy
development: A Malaysian experience’, in Proceedings of the International Conference on Energy
and Sustainable Development: Issues and Strategies (ESD 2010), pp. 1–6. doi:
10.1109/ESD.2010.5598779.
Nakata, T., Silva, D. and Rodionov, M. (2011) ‘Application of energy system models for designing a
low-carbon society’, Progress in Energy and Combustion Science. Elsevier Ltd, 37(4), pp. 462–502.
doi: 10.1016/j.pecs.2010.08.001.
Nasrin, A. B. et al. (2011) ‘Assessment of the Performance and Potential Export Renewable Energy
(RE) From Typical Cogeneration Plants Used in Palm Oil Mills’, Journal of Engineering and Applied
Sciences, pp. 433–439.
Nasution, M. A., Herawan, T. and Rivani, M. (2014) ‘Analysis of palm biomass as electricity from
palm oil mills in north sumatera’, Energy Procedia. Elsevier B.V., 47, pp. 166–172. doi:
10.1016/j.egypro.2014.01.210.
Nasution, M. A., Wibawa, D. S. and Noguchi, R. (2018) ‘Comperative environmental impact
evaluation of palm oil mill effluent treatment using a life cycle assessment approach: A case study
based on composting and combination for biogas technologies in North Sumatera of Indonesia’,
Journal of Cleaner Production, 184, pp. 1028–1040. Available at:
https://doi.org/10.1016/j.jclepro.2018.02.299.
National Academy of Sciences (2010) ‘Environmental Impacts of Renewable Electricity Generation’,
in The Power Of Renewables: Opportunities and Challenges for China and United States. Available
at: http://nap.edu/12987.
Negro, S. O., Alkemade, F. and Hekkert, M. P. (2012) ‘Why does renewable energy diffuse so
slowly? A review of innovation system problems’, Renewable and Sustainable Energy Reviews, pp.
3836–3846. doi: 10.1016/j.rser.2012.03.043.
Ng, F.-Y. et al. (2011) ‘A Renewable Future Driven with Malaysian Palm Oil-based Green
Technology’, Journal of Oil Palm & The Environment, 2(January), pp. 1–7. doi:
10.5366/jope.2011.01.
Ng, K. S. and Sadhukhan, J. (2011) ‘Process integration and economic analysis of bio-oil platform for
the production of methanol and combined heat and power’, Biomass and Bioenergy. Elsevier Ltd,
35(3), pp. 1153–1169. doi: 10.1016/j.biombioe.2010.12.003.

185
Ng, R. T. L., Ng, D. K. S. and Tan, R. R. (2013) ‘Systematic Approach for Synthesis of Integrated
Palm Oil Processing Complex. Part 2: Multiple Owners’, Industrial & Engineering Chemistry
Research, 52(30), pp. 10221–10235. doi: 10.1021/ie400846g.
Ng, W. P. Q. et al. (2012) ‘Waste-to-wealth : green potential from palm biomass in Malaysia’,
Journal of Cleaner Production, 34(September 2011), pp. 57–65. doi: 10.1016/j.jclepro.2012.04.004.
Ngadiron, Z. et al. (2015) ‘Review on Restructuring of Malaysia Electricity Supply Industry’, Applied
Mechanics and Materials, 773–774, pp. 476–480. doi: 10.4028/www.scientific.net/AMM.773-
774.476.
Nizami, A. S. et al. (2017) ‘Waste Biorefineries: Enabling Circular Economies in Developing
Countries’, Bioresource Technology. Elsevier Ltd. doi: 10.1016/j.biortech.2017.05.097.
Oh, T. H. et al. (2017a) ‘Energy policy and alternative energy in Malaysia: Issues and challenges for
sustainable growth - An update’, Renewable and Sustainable Energy Reviews. Elsevier Ltd, (May
2016), pp. 1–11. doi: 10.1016/j.rser.2017.06.112.
Oh, T. H. et al. (2017b) ‘Energy policy and alternative energy in Malaysia: Issues and challenges for
sustainable growth – An update’, Renewable and Sustainable Energy Reviews. Elsevier Ltd, 81(July
2017), pp. 3021–3031. doi: 10.1016/j.rser.2017.06.112.
Ong, C. L. et al. (2016) ‘Designing model for biomass transport cost of biofuel refinery in Malaysia’,
in Proceeding of the Burapha University International Conferences 2016, pp. 610–621. Available at:
www.buuconference.buu.ac.th.
Ong, H. C., Mahlia, T. M. I. and Masjuki, H. H. (2011) ‘A review on energy scenario and sustainable
energy in Malaysia’, Renewable and Sustainable Energy Reviews, 15(1), pp. 639–647. doi:
10.1016/j.rser.2010.09.043.
Oschmann, V. (2010) ‘A Success Story – The German Renewable Energy Act Turns Ten’, Renewable
Energy Law and Policy, 1, pp. 45–60.
Osmani, A. et al. (2013) ‘Electricity generation from renewables in the United States: Resource
potential, current usage, technical status, challenges, strategies, policies, and future directions’,
Renewable and Sustainable Energy Reviews. Elsevier, 24, pp. 454–472. doi:
10.1016/j.rser.2013.03.011.
Overend, R. P. (2004) ‘Direct Combustion of Biomass’, Renewable Energy Sources Charged With
Energy From The Sun And Originated From Earth-Moon Interaction, I.
Ozturk, M. et al. (2017) ‘Biomass and bioenergy: An overview of the development potential in
Turkey and Malaysia’, Renewable and Sustainable Energy Reviews. Elsevier Ltd, 79(April 2016), pp.
1285–1302. doi: 10.1016/j.rser.2017.05.111.
Padfield, R. et al. (2011) ‘Exploring opportunities for sustainability in the Malaysian palm oil
industry’, in International Conference on Environment Science and Engineering, pp. 175–178.
Pang, S. and Mujumdar, A. S. (2010) ‘Drying of woody biomass for bioenergy: Drying technologies
and optimization for an integrated bioenergy plant’, Drying Technology, pp. 690–701. doi:
10.1080/07373931003799236.
Panwar, N. L., Kaushik, S. C. and Kothari, S. (2011) ‘Role of renewable energy sources in
environmental protection: A review’, Renewable and Sustainable Energy Reviews. Elsevier Ltd,
15(3), pp. 1513–1524. doi: 10.1016/j.rser.2010.11.037.
Patel, C. et al. (2011) ‘Techno-economic performance analysis of energy production from biomass at

186
different scales in the UK context’, Chemical Engineering Journal. Elsevier B.V., 171(3), pp. 986–
996. doi: 10.1016/j.cej.2011.04.049.
Patel, C., Lettieri, P. and Germanà, A. (2011) ‘Techno-economic performance analysis and
environmental impact assessment of small to medium scale SRF combustion plants for energy
production in the UK’, Process Safety and Environmental Protection. Institution of Chemical
Engineers, 90(3), pp. 255–262. doi: 10.1016/j.psep.2011.06.015.
Patel, M., Zhang, X. and Kumar, A. (2016) ‘Techno-economic and life cycle assessment on
lignocellulosic biomass thermochemical conversion technologies: A review’, Renewable and
Sustainable Energy Reviews. Elsevier, 53, pp. 1486–1489. doi: 10.1016/j.rser.2015.09.070.
PEMANDU (2010) Economic Transformation Programme: A Roadmap for Malaysia. Available at:
https://policy.asiapacificenergy.org/sites/default/files/ETP.pdf.
Perez-Jimenez, J. A. (2015) ‘Gaseous Emissions from the Combustion of Biomass Pellets’, WIT
Transaction on State of the Art in Science and Engineering, 85, pp. 87–99. doi: 10.2495/978-1-84566-
062-8/006.
Perpiñá, C. et al. (2009) ‘Methodology based on Geographic Information Systems for biomass
logistics and transport optimisation’, Renewable Energy, 34(3), pp. 555–565. doi:
10.1016/j.renene.2008.05.047.
Perry, S., Klemeš, J. and Bulatov, I. (2008) ‘Integrating waste and renewable energy to reduce the
carbon footprint of locally integrated energy sectors’, Energy, 33(10), pp. 1489–1497. doi:
10.1016/j.energy.2008.03.008.
Petinrin, J. O. and Shaaban, M. (2015) ‘Renewable energy for continuous energy sustainability in
Malaysia’, Renewable and Sustainable Energy Reviews. Elsevier, 50, pp. 967–981. doi:
10.1016/j.rser.2015.04.146.
Petrenko, C., Paltseva, J. and Searle, S. (2016) Ecological impacts of palm oil expansion in Indonesia
| International Council on Clean Transportation. The International Council On Clean Transportation.
Available at: http://www.theicct.org/ecological-impacts-of-palm-oil-expansion-indonesia.
Pianosi, F. et al. (2016) ‘Sensitivity analysis of environmental models: A systematic review with
practical workflow’, Environmental Modelling and Software. Elsevier Ltd, 79, pp. 214–232. doi:
10.1016/j.envsoft.2016.02.008.
Piringer, G. et al. (2016) ‘Environmental hot spot analysis in agricultural lifecycle assessments – three
case studies’, Journal of Central European Agriculture. Journal of Central European Agriculture,
17(2), pp. 477–492. doi: 10.5513/JCEA01/17.2.1732.
Plevin, R. J. et al. (2010) ‘The greenhouse gas emissions from indirect land use change are uncertain,
but potentially much greater than previously estimated’, Environmental Science & Technology.,
44(21), pp. 8015–8021. doi: 10.1021/es101946t.
Polzin, F. et al. (2015) ‘Public policy influence on renewable energy investments-A panel data study
across OECD countries’, Energy Policy. Elsevier, 80, pp. 98–111. doi: 10.1016/j.enpol.2015.01.026.
Prado, V., Wender, B. A. and Seager, T. P. (2017) ‘Interpretation of comparative LCAs: external
normalization and a method of mutual differences’, International Journal of Life Cycle Assessment.
The International Journal of Life Cycle Assessment, 22(12), pp. 2018–2029. doi: 10.1007/s11367-
017-1281-3.
Qian, Y. et al. (2017) ‘Review of the state-of-the-art of biogas combustion mechanisms and
applications in internal combustion engines’, Renewable and Sustainable Energy Reviews. Elsevier,
187
69(October 2015), pp. 50–58. doi: 10.1016/j.rser.2016.11.059.
Rabl, A. et al. (2007) ‘How to account for CO2 emissions from biomass in an LCA’, International
Journal of Life Cycle Assessment, 12(5), p. 1. doi: 10.1065/lca2007.06.347.
Ragauskas, A. J. et al. (2006) ‘The path forward for biofuels and biomaterials’, Science, 311(5760),
pp. 484–489. doi: 10.1126/science.1114736.
Rahman, M. M., Saat, A. and Wahid, M. A. (2016) ‘Renewable energy policy in Germany and
Malaysia: Success factors’, Proceedings of the International Conference on Industrial Engineering
and Operations Management, 8–10 March, pp. 1085–1091.
Rajendran, K. (2017) ‘Effect of Moisture Content on Lignocellulosic Power Generation: Energy,
Economic and Environmental Impacts’, Processes, 5(78), pp. 1–15. doi: doi:10.3390/pr5040078.
Reeb, C. W. et al. (2014) ‘Supply Chain Analysis, Delivered Cost, and Life Cycle Assessment of Oil
Palm Empty Fruit Bunch Biomass for Green Chemical Production in Malaysia’, Bioresources, 9(3),
pp. 5385–5416. doi: 10.15376/biores.9.3.5385-5416.
Reiche, D. and Bechberger, M. (2004) ‘Policy differences in the promotion of renewable energies in
the EU member states’, Energy Policy, 32(7), pp. 843–849. doi: 10.1016/S0301-4215(02)00343-9.
Ren, X. et al. (2017) ‘Carbon, sulfur and nitrogen oxide emission from combustion of pulverized raw
and torrefied biomass’, Fuel, 188, pp. 310–323. Available at:
http://dx.doi.org/10.1016/j.fuel.2016.10.017.
REN 21 (2017) Renewables 2017: Global Status Report. Available at: http://www.ren21.net/wp-
content/uploads/2017/06/17-8399_GSR_2017_Full_Report_0621_Opt.pdf.
Rocha, M. H. et al. (2014) ‘Life cycle assessment (LCA) for biofuels in Brazilian conditions: A meta-
analysis’, Renewable and Sustainable Energy Reviews. Elsevier, 37, pp. 435–459. doi:
10.1016/j.rser.2014.05.036.
Rodriguez, L. C. et al. (2011) ‘Biomass assessment and small scale biomass fired electricity
generation in the Green Triangle, Australia’, Biomass and Bioenergy. Elsevier Ltd, 35(7), pp. 2589–
2599. doi: 10.1016/j.biombioe.2011.02.030.
Rosen, M. A., Le, M. N. and Dincer, I. (2005) ‘Efficiency analysis of a cogeneration and district
energy system’, Applied Thermal Engineering, 25(1), pp. 147–159. doi:
10.1016/j.applthermaleng.2004.05.008.
Russell, A., Ekvall, T. and Baumann, H. (2005) ‘Life cycle assessment - Introduction and overview’,
Journal of Cleaner Production, 13(13–14), pp. 1207–1210. doi: 10.1016/j.jclepro.2005.05.008.
Sadhukhan, J. et al. (2009) ‘Heat Integration Strategy for Economic Production of Combined Heat
and Power from Biomass Waste’, Energy and Fuels, 23(5), pp. 5106–5120. doi: 10.1021/ef900472s.
Sadhukhan, J. (2014) ‘Distributed and micro-generation from biogas and agricultural application of
sewage sludge: Comparative environmental performance analysis using life cycle approaches’,
Applied Energy. Elsevier Ltd, 122, pp. 196–206. doi: 10.1016/j.apenergy.2014.01.051.
Sadhukhan, J. et al. (2018) ‘Role of bioenergy, biorefinery and bioeconomy in sustainable
development:Strategic pathways for Malaysia’, Renewable and Sustainable Energy Reviews, 81, pp.
1966–1987. doi: http://dx.doi.org/10.1016/j.rser.2017.06.007.
Saidur, R. et al. (2011) ‘A review on biomass as a fuel for boilers’, Renewable and Sustainable
Energy Reviews, 15, pp. 2262–2289. doi: 10.1016/j.rser.2011.02.015.

188
Samiran, N. A. et al. (2016) ‘Progress in biomass gasification technique - With focus on Malaysian
palm biomass for syngas production’, Renewable and Sustainable Energy Reviews. Elsevier, 62, pp.
1047–1062. doi: 10.1016/j.rser.2016.04.049.
Sampattagul, S., Nutongkaew, P. and Kiatsiriroat, T. (2011) ‘Life cycle assessment of palm oil
biodiesel production in Thailand’, International Journal of Renewable Energy, 6(1), pp. 1–14.
Sarkar, S. K. et al. (2013) ‘Trends of Energy Demand and Supply as well as GHG Emissions in
Malaysia’, in 2nd International Conference on Agricultural, Environment and Biological Sciences,
pp. 89–94.
Saswattecha, K. et al. (2017) ‘Improving environmental sustainability of Thai palm oil production in
2050’, Journal of Cleaner Production. Elsevier Ltd, 147, pp. 572–588. doi:
10.1016/j.jclepro.2017.01.137.
Schmidt, J. H. (2010) ‘Comparative life cycle assessment of rapeseed oil and palm oil’, International
Journal of Life Cycle Assessment, 15(2), pp. 183–197. doi: 10.1007/s11367-009-0142-0.
Seabra, J. E. A. and Macedo, I. C. (2011) ‘Comparative analysis for power generation and ethanol
production from sugarcane residual biomass in Brazil’, Energy Policy. Elsevier, 39(1), pp. 421–428.
doi: 10.1016/j.enpol.2010.10.019.
Searcy, E. and Flynn, P. (2009) ‘The impact of biomass availability and processing cost on optimum
size and processing technology selection’, Applied Biochemistry and Biotechnology, 154(1–3), pp.
271–286. doi: 10.1007/s12010-008-8407-9.
SEDA (2018) Guideline on Biomass Power Plant Acceptance Test and Performance Assessment
(AT&PA) for Feed-in Tariff (FIT) Projects in Malaysia. Available at:
http://seda.gov.my/?omaneg=00010100000001010101000100001000000000000000000000&s=3296.
Sen, S. and Ganguly, S. (2017) ‘Opportunities, barriers and issues with renewable energy
development – A discussion’, Renewable and Sustainable Energy Reviews, pp. 1170–1181. doi:
10.1016/j.rser.2016.09.137.
Setyawan, D. (2014) ‘Formulating revolving fund scheme to support energy efficiency projects in
Indonesia’, Energy Procedia. Elsevier B.V., 47, pp. 37–46. doi: 10.1016/j.egypro.2014.01.194.
Shafie, S. . et al. (2012) ‘A review on electricity generation based on biomass residue in Malaysia’,
Renewable and Sustainable Energy Reviews. Elsevier, 16(8), pp. 5879–5889. doi:
10.1016/j.rser.2012.06.031.
Shahmohammadi, M. S. et al. (2015) ‘A decision support system for evaluating effects of Feed-in
Tariff mechanism: Dynamic modeling of Malaysia’s electricity generation mix’, Applied Energy.
Elsevier Ltd, 146, pp. 217–229. doi: 10.1016/j.apenergy.2015.01.076.
Shamsuddin, A. H. (2012) ‘Development of renewable energy in Malaysia strategic initiatives for
carbon reduction in the power generation sector’, in Procedia Engineering, pp. 384–391. doi:
10.1016/j.proeng.2012.10.150.
Shirley, R. and Kammen, D. (2015) ‘Energy planning and development in Malaysian Borneo:
Assessing the benefits of distributed technologies versus large scale energy mega-projects’, Energy
Strategy Reviews, 8, pp. 15–29. doi: 10.1016/j.esr.2015.07.001.
Shuit, S. H. et al. (2009) ‘Oil palm biomass as a sustainable energy source : A Malaysian case study’,
Energy. Elsevier Ltd, 34(9), pp. 1225–1235. doi: 10.1016/j.energy.2009.05.008.
Shuit, S. H. et al. (2009) ‘Oil palm biomass as a sustainable energy source: A Malaysian case study’,

189
Energy. Elsevier Ltd, 34(9), pp. 1225–1235. doi: 10.1016/j.energy.2009.05.008.
Silva Herran, D. and Nakata, T. (2012) ‘Design of decentralized energy systems for rural
electrification in developing countries considering regional disparity’, Applied Energy. Elsevier Ltd,
91(1), pp. 130–145. doi: 10.1016/j.apenergy.2011.09.022.
Simões-moreira, J. R. (2012) Thermal Power Plant Performance Analysis. doi: 10.1007/978-1-4471-
2309-5.
Singh, B. and Singh, V. K. (2017) ‘Fertilizer Management in Rice’, in Chauhan, B. S., Jabran, K., and
Mahajan, G. (eds) Rice Production World Wide, pp. 217–252. doi: DOI: 10.1007/978-3-319-47516-
5_10.
Singh, P. et al. (2013) ‘Using biomass residues from oil palm industry as a raw material for pulp and
paper industry: Potential benefits and threat to the environment’, Environment, Development and
Sustainability, 15(2), pp. 367–383. doi: 10.1007/s10668-012-9390-4.
Siregar, K. et al. (2015) ‘A Comparison of Life Cycle Assessment on Oil Palm (Elaeis guineensis
Jacq.) and Physic Nut (Jatropha curcas Linn.) as Feedstock for Biodiesel Production in Indonesia’,
Energy Procedia. Elsevier B.V., 65, pp. 170–179. doi: 10.1016/j.egypro.2015.01.054.
Sivers, M. Von and Zacchi, G. (1995) ‘A TECHNO-ECONOMICAL COMPARISON OF THREE
PROCESSES FOR THE PRODUCTION OF ETHANOL FROM’, 51, pp. 43–52.
von Sivers, M. and Zacchi, G. (1995) ‘A techno-economical comparison of three processes for the
production of ethanol from pine’, Bioresource Technology, pp. 43–52. doi: 10.1016/0960-
8524(94)00094-H.
Soimakallio, S. and Koponen, K. (2011) ‘How to ensure greenhouse gas emission reductions by
increasing the use of biofuels? - Suitability of the European Union sustainability criteria’, Biomass
and Bioenergy. Elsevier Ltd, 35(8), pp. 3504–3513. doi: 10.1016/j.biombioe.2011.04.041.
Song, H. et al. (2012) ‘Influence of drying process on the biomass-based polygeneration system of
bioethanol , power and heat’, Applied Energy. Elsevier Ltd, 90(1), pp. 32–37. doi:
10.1016/j.apenergy.2011.02.019.
Song, R. et al. (2017) ‘The implementation of inter-plant heat integration among multiple plants. Part
I: A novel screening algorithm’, Energy, pp. 1018–1029. doi: 10.1016/j.energy.2017.09.039.
Soraya, D. F. et al. (2014) ‘Life Cycle Assessment of Biodiesel Production from Palm Oil and
Jatropha Oil in Indonesia’, Journal of Sustainable Energy & Environment, 5, pp. 27–32.
Sovacool, B. K. and Drupady, I. M. (2011) ‘Examining the small renewable energy power (SREP)
program in Malaysia’, Energy Policy. Elsevier, 39(11), pp. 7244–7256. doi:
10.1016/j.enpol.2011.08.045.
Staffas, L., Gustavsson, M. and Mccormick, K. (2013) ‘Strategies and Policies for the Bioeconomy
and Bio-Based Economy: An Analysis of Official National Approaches’, pp. 2751–2769. doi:
10.3390/su5062751.
Ståhl, M. et al. (2004) ‘Industrial processes for biomass drying and their effects on the quality
properties of wood pellets’, Biomass and Bioenergy, 27(6), pp. 621–628. doi:
10.1016/j.biombioe.2003.08.019.
Van Stappen, F., Brose, I. and Schenkel, Y. (2011) ‘Direct and indirect land use changes issues in
European sustainability initiatives: State-of-the-art, open issues and future developments’, Biomass
and Bioenergy. Elsevier Ltd, 35(12), pp. 4824–4834. doi: 10.1016/j.biombioe.2011.07.015.

190
Steubing, B. et al. (2014) ‘Identifying environmentally and economically optimal bioenergy plant
sizes and locations: A spatial model of wood-based SNG value chains’, Renewable Energy. Elsevier,
61, pp. 57–68. doi: 10.1016/j.renene.2012.08.018.
Stichnothe, H. and Bessou, C. (2017) ‘Challenges for Life Cycle Assessment Of Palm Oil Production
System’, Indonesian Journal of Life Cycle Assessment and Sustainability, 1(2), pp. 1–9. Available at:
http://ijolcas.ilcan.or.id/index.php/IJoLCAS/article/view/28.
Stichnothe, H. and Schuchardt, F. (2011) ‘Life cycle assessment of two palm oil production systems’,
Biomass and Bioenergy. Elsevier Ltd, 35(9), pp. 3976–3984. doi: 10.1016/j.biombioe.2011.06.001.
Stougie, L. et al. (2018) ‘Environmental and exergetic sustainability assessment of power generation
from biomass’, Renewable Energy. Elsevier Ltd, 128, pp. 520–528. doi:
10.1016/j.renene.2017.06.046.
Strantzali, E. and Aravossis, K. (2016) ‘Decision making in renewable energy investments : A
review’, Renewable and Sustainable Energy Reviews. Elsevier, 55, pp. 885–898. doi:
10.1016/j.rser.2015.11.021.
Subramaniam, V. et al. (2010) ‘Life Cycle Assessment of the Production of Crude Palm Kernel Oil
(Part 3a)’, Journal of Oil Palm Research, 22(DECEMBER), pp. 895–903.
Sulaiman, F. et al. (2010) ‘A Perspective of Oil Palm and Its Wastes’, Journal of Physical Science,
21(September 2016), pp. 67–77. doi: 10.1016/j.jclepro.2012.04.004.
Sulaiman, F. et al. (2011) ‘An outlook of Malaysian energy, oil palm industry and its utilization of
wastes as useful resources’, Biomass and Bioenergy. Elsevier Ltd, 35(9), pp. 3775–3786. doi:
10.1016/j.biombioe.2011.06.018.
Sultana, A., Kumar, A. and Harfield, D. (2010) ‘Development of agri-pellet production cost and
optimum size’, Bioresource Technology. Elsevier Ltd, 101(14), pp. 5609–5621. doi:
10.1016/j.biortech.2010.02.011.
Summers, C. M. (1971) ‘The Conversion of Energy’, Scientific American, 225(3), pp. 148–160. doi:
10.1038/scientificamerican0971-148.
Tabe-Ojong, M. P. J. and Habiyaremye, N. (2017) ‘Economic Evaluation of Bioenergy Production
from Biomass in Germany: Implications for the bioeconomy?’, Journal of Sustainable Development,
10(6), p. 262. doi: 10.5539/jsd.v10n6p262.
Tan, R. R. et al. (2016) ‘An optimization-based cooperative game approach for systematic allocation
of costs and benefits in interplant process integration’, Chemical Engineering Research and Design.
Institution of Chemical Engineers, 106, pp. 43–58. doi: 10.1016/j.cherd.2015.11.009.
Tan, S. T. et al. (2014) ‘Energy and emissions benefits of renewable energy derived from municipal
solid waste: Analysis of a low carbon scenario in Malaysia’, Applied Energy. Elsevier Ltd, 136, pp.
797–804. doi: 10.1016/j.apenergy.2014.06.003.
Teh, C. B. S. (2016) Availability, use, and removal of oil palm biomass in Indonesia, Report prepared
for the International Council on Clean Transportation. doi: 10.13140/RG.2.1.4697.4485.
Tenorio, C. and Moya, R. (2012) ‘Evaluation of different approaches for the drying of lignocellulose
residues’, BioResources, 7(3), pp. 3500–3514. doi: 10.15376/biores.7.3.3500-3514.
Theo, W. L. et al. (2017) ‘Optimisation of oil palm biomass and palm oil mill effluent (POME)
utilisation pathway for palm oil mill cluster with consideration of BioCNG distribution network’,
Energy. Elsevier Ltd, 121(February), pp. 865–883. doi: 10.1016/j.energy.2017.01.021.

191
Thornley, P. et al. (2009) ‘Integrated assessment of bioelectricity technology options’, Energy Policy,
37(3), pp. 890–903. doi: 10.1016/j.enpol.2008.10.032.
Tong, S. L. and Jaafar, A. B. (2004) ‘Waste to Energy: Methane Recovery from Anaerobic Digestion
of POME’, Energy Smart, 14, pp. 4–11.
Treyer, K. and Bauer, C. (2016) ‘Life cycle inventories of electricity generation and power supply in
version 3 of the ecoinvent database—part II: electricity markets’, International Journal of Life Cycle
Assessment. The International Journal of Life Cycle Assessment, 21(9), pp. 1255–1268. doi:
10.1007/s11367-013-0694-x.
Tsai, W. T. (2012) ‘An analysis of the use of biosludge as an energy source and its environmental
benefits in Taiwan’, Energies, 5(8), pp. 3064–3073. doi: 10.3390/en5083064.
Turconi, R., Boldrin, A. and Astrup, T. (2013) ‘Life cycle assessment (LCA) of electricity generation
technologies: Overview, comparability and limitations’, Renewable and Sustainable Energy Reviews.
Elsevier, 28, pp. 555–565. doi: 10.1016/j.rser.2013.08.013.
Umar, M. S., Jennings, P. and Urmee, T. (2012) ‘The Palm Oil Biomass Renewable Energy
Development in Malaysia : How Sustainable is this Industry ?’, in International Conference on
Sustainable Energy Engineering and Application (ICSEEA) 2012.
Umar, M. S., Jennings, P. and Urmee, T. (2013a) ‘Strengthening the palm oil biomass Renewable
Energy industry in Malaysia’, Renewable Energy, 60, pp. 107–115. doi:
10.1016/j.renene.2013.04.010.
Umar, M. S., Jennings, P. and Urmee, T. (2013b) ‘Strengthening the palm oil biomass Renewable
Energy industry in Malaysia’, RENE. Elsevier, 60, pp. 107–115. doi: 10.1016/j.renene.2013.04.010.
Umar, M. S., Jennings, P. and Urmee, T. (2014a) ‘Generating renewable energy from oil palm
biomass in Malaysia : The Feed-in Tariff policy framework’, Biomass and Bioenergy. Elsevier Ltd,
62, pp. 37–46. doi: 10.1016/j.biombioe.2014.01.020.
Umar, M. S., Jennings, P. and Urmee, T. (2014b) ‘Sustainable electricity generation from oil palm
biomass wastes in Malaysia: An industry survey’, Energy. Elsevier Ltd, 67, pp. 496–505. doi:
10.1016/j.energy.2014.01.067.
Umar, M. S., Urmee, T. and Jennings, P. (2017) ‘A Policy Framework and Industry Roadmap Model
for Sustainable Oil Palm Biomass Electricity Generation in Malaysia’, Renewable Energy. Elsevier
B.V. doi: 10.1016/j.renene.2017.12.060.
UNFCC (2017) Methodological tool: Project and leakage emissions from biomass - Version 02.0.
Available at: http://cdm.unfccc.int/methodologies/PAmethodologies/tools/am-tool-16-
v1.pdf/history_view.
Varun, Bhat, I. K. and Prakash, R. (2009) ‘LCA of renewable energy for electricity generation
systems-A review’, Renewable and Sustainable Energy Reviews, 13(5), pp. 1067–1073. doi:
10.1016/j.rser.2008.08.004.
Vázquez-Rowe, I. et al. (2014) ‘Applying consequential LCA to support energy policy: Land use
change effects of bioenergy production’, Science of the Total Environment. Elsevier B.V., 472, pp.
78–89. doi: 10.1016/j.scitotenv.2013.10.097.
Verbruggen, A. et al. (2010) ‘Renewable energy costs, potentials, barriers: Conceptual issues’,
Energy Policy, 38(2), pp. 850–861. doi: 10.1016/j.enpol.2009.10.036.
Vijay, V. et al. (2016) ‘The Impacts of Oil Palm on Recent Deforestation and Biodiversity Loss The

192
Impacts of Oil Palm on Recent Deforestation and Biodiversity Loss’, PLoS ONE, 11(October), pp. 1–
19. doi: 10.5061/dryad.2v77j.
Visvanathan, C. and Chiemchaisri, C. (2006) ‘Management of Agricultural Wastes and Residues in
Thailand : Wastes to Energy Approach’, Agro Wastes. Available at:
http://www.faculty.ait.asia/visu/Prof Visu’s CV/Conferance/28/Agri-waste2energy-Thai.pdf.
Wahid, M. A. and Ujir, H. (2012) ‘Reacting shock waves characteristics for biogas compared to other
gaseous fuel’, AIP Conference Proceedings, 1440, pp. 90–99. doi: 10.1063/1.4704206.
Wain, Y. A. (2014) ‘Updating the Lang Factor and Testing its Accuracy, Reliability and Precision as
a Stochastic Cost Estimating Method’, PM World Journal Updating the Lang Factor and Testing its
Accuracy, Reliability, III(X), pp. 1–17. Available at: www.pmworldlibrary.net.
Walla, C. and Schneeberger, W. (2008) ‘The optimal size for biogas plants’, Biomass and Bioenergy,
32(6), pp. 551–557. doi: 10.1016/j.biombioe.2007.11.009.
Wan, Y. K., Sadhukhan, J., Ng, K. S. and Ng, D. K. S. (2016) ‘Techno-economic evaluations for
feasibility of sago-based biorefinery, Part 1: Alternative energy systems’, Chemical Engineering
Research and Design. Institution of Chemical Engineers, 107, pp. 263–279. doi:
10.1016/j.cherd.2015.11.001.
Wan, Y. K., Sadhukhan, J., Ng, K. S. and Ng, D. K. S. (2016) ‘Techno-economic evaluations for
feasibility of sago-based biorefinery, Part 1: Alternative energy systems’, Chemical Engineering
Research and Design. Institution of Chemical Engineers, 107, pp. 263–279. doi:
10.1016/j.cherd.2015.11.001.
Wang, L. et al. (2009) ‘Technical and economical analyses of combined heat and power generation
from distillers grains and corn stover in ethanol plants’, Energy Conversion and Management.
Elsevier Ltd, 50(7), pp. 1704–1713. doi: 10.1016/j.enconman.2009.03.025.
Warren, B., Christoff, P. and Green, D. (2016) ‘Australia’s sustainable energy transition: The
disjointed politics of decarbonisation’, Environmental Innovation and Societal Transitions. Elsevier
B.V., 21, pp. 1–12. doi: 10.1016/j.eist.2016.01.001.
WBCSD (2000) A Corporate Accounting and Reporting Standard. Available at:
https://ghgprotocol.org/sites/default/files/standards/ghg-protocol-revised.pdf.
White, A. J., Kirk, D. W. and Graydon, J. W. (2011) ‘Analysis of small-scale biogas utilization
systems on Ontario cattle farms’, Renewable Energy. Elsevier Ltd, 36(3), pp. 1019–1025. doi:
10.1016/j.renene.2010.08.034.
Wicke, B. et al. (2011) ‘Exploring land use changes and the role of palm oil production in Indonesia
and Malaysia’, Land Use Policy. Elsevier Ltd, 28(1), pp. 193–206. doi:
10.1016/j.landusepol.2010.06.001.
Wielgosinski, G., Lechtanska, P. and Namiecinska, O. (2017) ‘Emission of some pollutants from
biomass combustion in comparison to hard coal combustion’, Journal of the Energy Institute, 90, pp.
787–796. Available at: http://dx.doi.org/10.1016/j.joei.2016.06.005.
Wood, S. R. and Rowley, P. N. (2011) ‘A techno-economic analysis of small-scale, biomass-fuelled
combined heat and power for community housing’, Biomass and Bioenergy. Elsevier Ltd, 35(9), pp.
3849–3858. doi: 10.1016/j.biombioe.2011.04.040.
Wood, S. R. and Rowley, P. N. (2011) ‘A techno-economic analysis of small-scale , biomass-fuelled
combined heat and power for community housing’, Biomass and Bioenergy. Elsevier Ltd, 35(9), pp.
3849–3858. doi: 10.1016/j.biombioe.2011.04.040.
193
World Energy Council (2016a) World Energy Resources: Waste to Energy 2016. Available at:
https://www.worldenergy.org/wp-content/uploads/2013/10/WER_2013_7b_Waste_to_Energy.pdf.
World Energy Council (2016b) World Energy Resources 2016. Available at:
https://www.worldenergy.org/wp-content/uploads/2016/10/World-Energy-Resources-Full-report-
2016.10.03.pdf.
Wright, M., Brown, R. C. and August, R. (2007) ‘Establishing the optimal sizes of different kinds of
biorefi neries’, pp. 191–200. doi: 10.1002/bbb.
Wu, J. (2008) ‘Land Use Changes: Economic, Social, and Environmental Impacts’, CHOICES 4th
Quarter, 23(4).
Wu, Q. et al. (2017) ‘Sustainable and renewable energy from biomass wastes in palm oil industry: A
case study in Malaysia’, International Journal of Hydrogen Energy. Elsevier Ltd, 42(37), pp. 23871–
23877. doi: 10.1016/j.ijhydene.2017.03.147.
Wu, T. Y. et al. (2010) ‘Pollution control technologies for the treatment of palm oil mill effluent
(POME) through end-of-pipe processes’, Journal of Environmental Management. Elsevier Ltd, 91(7),
pp. 1467–1490. doi: 10.1016/j.jenvman.2010.02.008.
Wu, X. et al. (2015) ‘Steam power plant configuration, design, and control’, Wiley Interdisciplinary
Reviews: Energy and Environment, 4(6), pp. 537–563. doi: 10.1002/wene.161.
Xu, Q. and Pang, S. (2008) ‘Mathematical modeling of rotary drying of woody biomass’, Drying
Technology, 26(11), pp. 1344–1350. doi: 10.1080/07373930802331050.
Yacob, S. et al. (2006) ‘Baseline study of methane emission from anaerobic ponds of palm oil mill
effluent treatment’, Science of the Total Environment, 366(1), pp. 187–196. doi:
10.1016/j.scitotenv.2005.07.003.
Yarnal, G. S. and Puranik, V. S. (2010) ‘Energy Management Study in Sugar Industries by Various
Bagasse Drying Methods’, Strategic Planning for Energy and the Environment, 29(3), pp. 56–78. doi:
10.1080/10485231009595087.
Yarnal, G. S. and Puranik, V. S. (2016) ‘Energy Management Study in Sugar Industries by Various
Bagasse Drying Methods Energy Management Study in Sugar Industries by Various Bagasse Drying
Methods’, 5236(November). doi: 10.1080/10485231009595087.
Yatim, P. et al. (2016) ‘Energy policy shifts towards sustainable energy future for Malaysia’, Clean
Technologies and Environmental Policy. Springer Berlin Heidelberg. doi: 10.1007/s10098-016-1151-
x.
Yatim, P. et al. (2017) ‘Overview of the key risks in the pioneering stage of the Malaysian biomass
industry’, Clean Technologies and Environmental Policy. Springer Berlin Heidelberg, 19(7), pp.
1825–1839. doi: 10.1007/s10098-017-1369-2.
Yeoh, B. G. (2004) ‘A Technical and Economic Analysis of Heat and Power Generation from
Biomethanation of Palm Oil Mill Effluent’, International Energy Journal, 6(January), pp. 63–79.
Available at: https://cdm.unfccc.int/Projects/DB/DNV-
CUK1186564216.66/ReviewInitialComments/RHZT1TSAN6IZJRF5KIY8FAHND28EF9.
Yilmaz, S. and Selim, H. (2013) ‘A review on the methods for biomass to energy conversion systems
design’, Renewable and Sustainable Energy Reviews, 25, pp. 420–430. doi:
10.1016/j.rser.2013.05.015.
Yue, D., You, F. and Snyder, S. W. (2014) ‘Biomass-to-bioenergy and biofuel supply chain

194
optimization: Overview, key issues and challenges’, Computers & Chemical Engineering, 66, pp. 36–
56. doi: 10.1016/j.compchemeng.2013.11.016.
Yui, S. and Yeh, S. (2013) ‘Land use change emissions from oil palm expansion in Pará, Brazil
depend on proper policy enforcement on deforested lands’, Environmental Research Letters, 8(4), pp.
1–10. doi: 10.1088/1748-9326/8/4/044031.
Yusoff, S. (2006) ‘Renewable energy from palm oil – innovation on effective utilization of waste’,
Journal of Cleaner Production, 14(1), pp. 87–93. doi: 10.1016/j.jclepro.2004.07.005.
Zema, D. A. (2017) ‘Planning the optimal site, size, and feed of biogas plants in agricultural districts’,
Biofuels, Bioproducts and Biorefining, pp. 454–471. doi: 10.1002/bbb.1757.
Zhu, X. G., Long, S. P. and Ort, D. R. (2010) ‘Advanced technology paths to global climate stability:
Energy for a greenhouse planet’, Annu Rev Plant Biol, 61(November), pp. 235–261. doi:
10.1146/annurev-arplant-042809-112206.
Zulkifli, H. et al. (2010) ‘Life cycle assessment for oil palm fresh fruit bunch production from
continued land use for oil palm planted on mineral soil (Part 2)’, Journal of Oil Palm Research,
22(December), pp. 887–894. doi: 10.1007/s00122-012-1801-2.

195
Appendix A

Table A-1 summarises the policies and initiatives taken by the Government of Malaysia to foster the development of electricity generation from
renewable resources since it first introduced until the most recent.

Table A-1: RE Policies and Initiatives in Malaysia

Year Policy/Act/Strategy/Master Plan Targets Initiative(s) Outcome(s)


2000 Fifth Fuel Policy 5% contribution from the RE in Initiation of SREP under Special 12 MW of RE-based
To recognize the potential of the energy mixture which is Committee on RE (SCORE) in electricity was connected to
biomass, biogas, MSW, solar and equivalent to 500 MW grid- 2001 to encourage private sector to the grid at the end of 2005.
mini-hydro as the alternative fuel connected power generation by invest into small scale RE-based RE contribution was reduced
for electricity generation. 2005 power generation. to 350 MW by the end of 2010
To diversify the energy mixture for Establishment of Biomass Power 65 MW of RE-based capacity
electricity generation to reduce the Generation and Demonstration was installed at the end of
over-dependency on fossil fuel. (BioGen) in 2002 to promote and 2010
To mitigate the climate change demonstrate the biomass and biogas
effects. grid-connected power generation
project
Procurement of Pioneer Status (PS)
and entitlement for Investment Tax
Allowance (ITA)
2010 NREAP Short Term Target Institution of Sustainable Energy 100 MW installed capacity as
To enhance the usage of RE towards 6% contribution from the RE in Development Authorities Act of 2012
national electricity supply security the energy mixture which is (2011) that appoint Sustainable
while reducing carbon emission. equivalent to 985 MW grid- Energy Development Authorities
connected power generation by (SEDA) Malaysia as the
2015. governance body to promote and
Medium Term Target implement the NREAP objective. It
also act as a sole manager of the
196
11% contribution from the RE Feed-in Tariff (FiT) system in
in the energy mixture which is Peninsular Malaysia and Sabah.
equivalent to 2080 MW grid- Institution of RE Act (2011) that
connected power generation by incorporate Feed-in Tariff (FiT)
2020. system to obligate the National
Long Term Target Utility to purchase the RE-based
17% contribution from the RE electricity at fixed price for a
in the energy mixture which is specific duration.
equivalent to 4000 MW grid-
connected power generation by
2030.
73% contribution from the RE
in the energy mixture which is
equivalent to 21.4 GW grid-
connected power generation by
2050.

Electricity Generation from


Palm Oil Biomass Target 12MW palm oil biomass
electricity generation capacity
Short Term Target (STR) at the end of 2014
330 MW grid-connected power
generation by 2015.
Medium Term Target(MTR)
800 MW grid-connected power
generation by 2020.
Long Term Target (LTR)
1340 MW grid-connected
power generation by 2030.

197
2011 National Biomass Strategy 2020 Focusing on the MTR and Introduction of 1 Malaysia Biomass
(NBS 2020) LTR of electricity generation Strategy (2012) to promote the
To promote higher value downstream from palm oil biomass benefits of utilizing biomass across
uses of biomass with an initial focus Additional of US$ 6.9 billion sectors and to enhance the
on palm oil industries. gross national income sustainability of biomass industries.
To allow the palm oil industries to Creating 66,000 new high
capture the opportunities in value job opportunities
electricity generation sector Potential 12% reduction of
CO2 emission from electricity
generation sector.
2011 Economic Transformation Under National Key Economic National Biogas Implementation
Programme (ETP) Area : Palm Oil (EPP5) to encourage the palm oil
To boost the economic growth and Installation of biogas facilities mills in Malaysia to install the
achieve a high income status by 2020 at almost all mills across biogas capture system and utilize it
Malaysia to meet the target of to achieve the target.
410 MW installed capacity by
2030
2017 Green Technology Master Plan The aspirational target of 30% The on-going initiatives introduced
Malaysia 2017-2030 RE-based installed electricity earlier to promote the utilisation of
To facilitates the mainstreaming of capacity by 2030 biomass for electricity generation
green technology

198
Appendix B

This appendix contains the data checklist form used during the field visits to collect the required
data, statistical information on the availability of the EFB and biogas, and the spreadsheets
used to model the POMR-RES in Chapter 4.

Appendix B-1 – Data Checklist Form

199
Appendix B-2 – Statistic information of available EFB and biogas in Peninsular Malaysia

Table B-1: Availability of EFB and biogas in the mill according to their FFB processing capacity
Processing Capacity (t/hr) 10 20 30 40 50 60 70 80 90 100 110 120
Amount of Available EFB 12.00 24.00 36.00 48.00 60.00 72.00 84.00 96.00 108.00 120.00 132.00 144.00
(k t/yr)
Amount of Available biogas 0.05 0.11 0.16 0.21 0.27 0.32 0.37 0.42 0.48 0.53 0.58 0.64
(k t/yr)
Amount of available semi- 9.24 18.48 27.72 36.96 46.20 55.44 64.68 73.92 83.16 92.40 101.64 110.88
dry EFB for electricity
generation1
(k t/yr)
Amount of available biogas 0.04 0.10 1.43 1.91 2.39 2.86 3.34 3.82 4.29 4.77 5.25 5.72
for electricity generation2
(k t/yr)

Note:
1
the mass of semi-dry EFB is reduced by 23% from the raw EFB at the collection point
2
the mass of biogas is subjected to the methane leakage factor of 0.1.

200
Appendix B-3 – Mathematical Modelling Spreadsheet

The Excel spreadsheets were developed in two parts. The first part is to estimate the availability
of EFB and biogas as in Tables 4.6 and 4.7. The electricity generation capacity of the EFB and
biogas is calculated in the later part of the spreadsheet using Eq. (1) – Eq. (5) in Section 4.3.1.
The data types involved in the spreadsheet to model the POMR-RES power plant are further
explained in Table B-2.

Table B-2: Description of Data Used for Modelling the POMR-RES Power Plant
Data Types Unit Remarks
Mill Processing t/hr The mill processing capacity was measured referring to the
Capacity amount of FFB process in hourly basis typically varies
from 10 t/hr to 120 t/hr.
Amount of FFB t/yr The amount of the FFB processed was obtained from the
processed FFB processing log or can be estimated by multiplying the
mill processing capacity with the mill operation hours.
Amount of EFB t/yr The mass amount of raw EFB available was obtained from
available the in-house log or can be estimated at 15% from the amount
of FFB processed. The mass amount of the semi-dry EFB is
77% of the mass amount of raw EFB.
Amount of POME m3/yr The mass amount of POME available was obtained from the
available in-house log or can be estimated at 0.65m3 for every tonne
of FFB processed.
COD POME mg/l The standard COD value of 51000 mg/l was used to
represent the organic content of every litre of POME.
COD converted m3/yr The amount of organic content produced from POME in the
anaerobic digestion process was calculated by multiplying
the amount of POME available with the COD value and the
efficiency of the AD process (80%).
Methane Produced t/yr The amount of methane produced from the organic content
of POME was calculated by multiplying the COD converted
value with the methane conversion factor of 0.25 kg CH4/kg
COD. The methane produced was subjected to 0.1 leakage
factor.
Mill Operation Hours hrs/yr The operation hours for the mills and proposed energy
plant was set at 8000 hours.
Feedstock Feed Rate kg/s The mass flow rate of EFB and biogas into the boiler was
calculated by dividing the amount of available EFB and
methane produced with the POMR-RES operation time.
Calorific Value MJ/kg The CV of 8.5 MJ/kg was used for semi-dry EFB and 50
MJ/kg for biogas.
Extractable Energy MW The amount of superheated steam (50 bar, 500°C) produces
from residues (𝑸𝑪) from combusting the residues in the boiler was calculated
using Eq. (1).
Mass Flow Rate of kg/s The mass of the superheated steam that flowed to the turbine
Superheated Steam was calculated using Eq. (2). The enthalpy properties of the
(𝒎𝑯𝑷𝑺𝑯𝑺 ) superheated steam, saturated steam, vaporization of water,
201
saturated temperature of steam, the temperature of boiler
feed water and the specific heat of water from Wan et al.
(2016) were used for this estimation.
Total Superheated t/yr The total amount of superheated steam generated from the
Steam Generated boiler used for the electricity generation process was
(𝑯𝑷𝑺𝑯𝑺𝑨𝑪𝑪 ) calculated by multiplying the 𝑚𝐻𝑃𝑆𝐻𝑆 with the operation
time of the POMR-RES as in Eq. (3).
Electricity MW The electricity generation capacity from the expansion of
Generation Capacity superheated steam in the back pressure steam turbine was
(𝑬𝑪𝑨𝑷 ) calculated using Eq. (4).

Total Electricity MWh/yr The amount of electricity generated in the POMR-RES is the
Generated (𝑬𝑮𝑬𝑵 ) multiplication product of electricity generation capacity
with the operation time as in Eq. (5).

Appendix B-3-1 – Semi-dry EFB-based POMR-RES Power Plant

Table B-3: Modelling of semi-dry EFB-based Power Plant in 10 t/hr mill


Mill Processing Capacity (t/hr) 10
Amount of FFB process (t/yr) 80000
Amount EFB Available (t/yr) 9240
Operation Hour (hrs/yr) 8000
Feedstock Feed rate (kg/s) 0.3208
Calorific Value (MJ/kg) 8.5000

Mathematical Modelling
Extractable Energy from Residues (MW) 2.18
Mass Flow Rate of Superheated Steam (kg/s) 0.74
Total Superheated Steam Generated (t/yr) 21336
Total Available Steam for Electricity Generation (t/yr) 21336
Installed Electricity Generation Capacity (MW) 0.89
Total Electricity Generated (MWh/yr) 7122.76

Table B-4: Modelling of semi-dry EFB-based Power Plant in 20 t/hr mill


Mill Processing Capacity (t/hr) 20
Amount of FFB process (t/yr) 160000
Amount EFB Available (t/yr) 18480
Operation Hour (hrs/yr) 8000
Feedstock Feed rate (kg/s) 0.6417
Calorific Value (MJ/kg) 8.5000

202
Mathematical Modelling
Extractable Energy from Residues (MW) 4.36
Mass Flow Rate of Superheated Steam (kg/s) 1.48
Total Superheated Steam Generated (t/yr) 42673
Total Available Steam for Electricity Generation (t/yr) 42673
Installed Electricity Generation Capacity (MW) 1.78
Total Electricity Generated (MWh/yr) 14245.51

Table B-5: Modelling of semi-dry EFB-based Power Plant in 30 t/hr mill


Mill Processing Capacity (t/hr) 30
Amount of FFB process (t/yr) 240000
Amount EFB Available (t/yr) 27720
Operation Hour (hrs/yr) 8000
Feedstock Feed rate (kg/s) 0.9625
Calorific Value (MJ/kg) 8.5000

Mathematical Modelling
Extractable Energy from Residues (MW) 6.55
Mass Flow Rate of Superheated Steam (kg/s) 2.22
Total Superheated Steam Generated (t/yr) 64009
Total Available Steam for Electricity Generation (t/yr) 64009
Installed Electricity Generation Capacity (MW) 2.67
Total Electricity Generated (MWh/yr) 21368.27

Table B-6: Modelling of semi-dry EFB-based Power Plant in 40 t/hr mill


Mill Processing Capacity (t/hr) 40
Amount of FFB process (t/yr) 320000
Amount EFB Available (t/yr) 36960
Operation Hour (hrs/yr) 8000
Feedstock Feed rate (kg/s) 1.2833
Calorific Value (MJ/kg) 8.5000

Mathematical Modelling
Extractable Energy from Residues (MW) 8.73
Mass Flow Rate of Superheated Steam (kg/s) 2.96
Total Superheated Steam Generated (t/yr) 85345
Total Available Steam for Electricity Generation (t/yr) 85345
Installed Electricity Generation Capacity (MW) 3.56
Total Electricity Generated (MWh/yr) 28491.03

203
Table B-7: Modelling of semi-dry EFB-based Power Plant in 50 t/hr mill
Mill Processing Capacity (t/hr) 50
Amount of FFB process (t/yr) 400000
Amount EFB Available (t/yr) 46200
Operation Hour (hrs/yr) 8000
Feedstock Feed rate (kg/s) 1.6042
Calorific Value (MJ/kg) 8.5000

Mathematical Modelling
Extractable Energy from Residues (MW) 10.91
Mass Flow Rate of Superheated Steam (kg/s) 3.70
Total Superheated Steam Generated (t/yr) 106682
Total Available Steam for Electricity Generation (t/yr) 106682
Installed Electricity Generation Capacity (MW) 4.45
Total Electricity Generated (MWh/yr) 35613.78

Table B-8: Modelling of semi-dry EFB-based Power Plant in 60 t/hr mill


Mill Processing Capacity (t/hr) 60
Amount of FFB process (t/yr) 480000
Amount EFB Available (t/yr) 55440
Operation Hour (hrs/yr) 8000
Feedstock Feed rate (kg/s) 1.9250
Calorific Value (MJ/kg) 8.5000

Mathematical Modelling
Extractable Energy from Residues (MW) 13.09
Mass Flow Rate of Superheated Steam (kg/s) 4.45
Total Superheated Steam Generated (t/yr) 128018
Total Available Steam for Electricity Generation (t/yr) 128018
Installed Electricity Generation Capacity (MW) 5.34
Total Electricity Generated (MWh/yr) 42736.54

Table B-9: Modelling of semi-dry EFB-based Power Plant in 70 t/hr mill


Mill Processing Capacity (t/hr) 70
Amount of FFB process (t/yr) 560000
Amount EFB Available (t/yr) 64680
Operation Hour (hrs/yr) 8000
Feedstock Feed rate (kg/s) 2.2458
Calorific Value (MJ/kg) 8.5000

204
Mathematical Modelling
Extractable Energy from Residues (MW) 15.27
Mass Flow Rate of Superheated Steam (kg/s) 5.19
Total Superheated Steam Generated (t/yr) 149354
Total Available Steam for Electricity Generation (t/yr) 149354
Installed Electricity Generation Capacity (MW) 6.23
Total Electricity Generated (MWh/yr) 49859.30

Table B-10: Modelling of semi-dry EFB-based Power Plant in 80 t/hr mill


Mill Processing Capacity (t/hr) 80
Amount of FFB process (t/yr) 640000
Amount EFB Available (t/yr) 73920
Operation Hour (hrs/yr) 8000
Feedstock Feed rate (kg/s) 2.5667
Calorific Value (MJ/kg) 8.5000

Mathematical Modelling
Extractable Energy from Residues (MW) 17.45
Mass Flow Rate of Superheated Steam (kg/s) 5.93
Total Superheated Steam Generated (t/yr) 170691
Total Available Steam for Electricity Generation (t/yr) 170691
Installed Electricity Generation Capacity (MW) 7.12
Total Electricity Generated (MWh/yr) 56982.05

Table B-11: Modelling of semi-dry EFB-based Power Plant in 90 t/hr mill


Mill Processing Capacity (t/hr) 90
Amount of FFB process (t/yr) 720000
Amount EFB Available (t/yr) 83160
Operation Hour (hrs/yr) 8000
Feedstock Feed rate (kg/s) 2.8875
Calorific Value (MJ/kg) 8.5000

Mathematical Modelling
Extractable Energy from Residues (MW) 19.64
Mass Flow Rate of Superheated Steam (kg/s) 6.67
Total Superheated Steam Generated (t/yr) 192027
Total Available Steam for Electricity Generation (t/yr) 192027
Installed Electricity Generation Capacity (MW) 8.01

205
Total Electricity Generated (MWh/yr) 64104.81

Table B-12: Modelling of semi-dry EFB-based Power Plant in 100 t/hr mill
Mill Processing Capacity (t/hr) 100
Amount of FFB process (t/yr) 800000
Amount EFB Available (t/yr) 92400
Operation Hour (hrs/yr) 8000
Feedstock Feed rate (kg/s) 3.2083
Calorific Value (MJ/kg) 8.5000

Mathematical Modelling
Extractable Energy from Residues (MW) 21.82
Mass Flow Rate of Superheated Steam (kg/s) 7.41
Total Superheated Steam Generated (t/yr) 213364
Total Available Steam for Electricity Generation (t/yr) 213364
Installed Electricity Generation Capacity (MW) 8.90
Total Electricity Generated (MWh/yr) 71227.57

Table B-13: Modelling of semi-dry EFB-based Power Plant in 110 t/hr mill
Mill Processing Capacity (t/hr) 110
Amount of FFB process (t/yr) 880000
Amount EFB Available (t/yr) 101640
Operation Hour (hrs/yr) 8000
Feedstock Feed rate (kg/s) 3.5292
Calorific Value (MJ/kg) 8.5000

Mathematical Modelling
Extractable Energy from Residues (MW) 24.00
Mass Flow Rate of Superheated Steam (kg/s) 8.15
Total Superheated Steam Generated (t/yr) 234700
Total Available Steam for Electricity Generation (t/yr) 234700
Installed Electricity Generation Capacity (MW) 9.79
Total Electricity Generated (MWh/yr) 78350.32

206
Table B-14: Modelling of semi-dry EFB-based Power Plant in 120 t/hr mill
Mill Processing Capacity (t/hr) 120
Amount of FFB process (t/yr) 960000
Amount EFB Available (t/yr) 110880
Operation Hour (hrs/yr) 8000
Feedstock Feed rate (kg/s) 3.8500
Calorific Value (MJ/kg) 8.5000

Mathematical Modelling
Extractable Energy from Residues (MW) 26.18
Mass Flow Rate of Superheated Steam (kg/s) 8.89
Total Superheated Steam Generated (t/yr) 256036
Total Available Steam for Electricity Generation (t/yr) 256036
Installed Electricity Generation Capacity (MW) 10.68
Total Electricity Generated (MWh/yr) 85473.08

Appendix B-3-2 – Biogas-based POMR-RES Power Plant

Table B-15: Modelling of biogas-based Power Plant in 10 t/hr mill


Mill Processing Capacity (t/hr) 10
Mill Operation Hour (hrs/yr) 8000
Amount of FFB process (t/yr) 80000
Amount POME Available (m3/yr) 52000
COD POME (mg/l) 51000
COD Converted (m3/yr) 2121600
COD Converted (t/yr) 2121.6
Methane Produced (t/yr) 477.36
Feedstock Feed rate (kg/s) 0.0166
Calorific Value (MJ/kg) 50.0000

Mathematical Modelling
Extractable Energy from Residues (MW) 0.75
Mass Flow Rate of Superheated Steam (kg/s) 0.25
Total Superheated Steam Generated (t/yr) 7295
Total Available Steam for Electricity Generation (t/yr) 7295
Installed Electricity Generation Capacity (MW) 0.30
Total Electricity Generated (MWh/yr) 2435.15

207
Table B-16: Modelling of biogas-based Power Plant in 20 t/hr mill
Mill Processing Capacity (t/hr) 20
Mill Operation Hour (hrs/yr) 8000
Amount of FFB process (t/yr) 160000
Amount POME Available (m3/yr) 104000
COD POME (mg/l) 51000
COD Converted (m3/yr) 4243200
COD Converted (t/yr) 4243.2
Methane Produced (t/yr) 954.72
Feedstock Feed rate (kg/s) 0.0332
Calorific Value (MJ/kg) 50.0000

Mathematical Modelling
Extractable Energy from Residues (MW) 1.49
Mass Flow Rate of Superheated Steam (kg/s) 0.51
Total Superheated Steam Generated (t/yr) 14589
Total Available Steam for Electricity Generation (t/yr) 14589
Installed Electricity Generation Capacity (MW) 0.61
Total Electricity Generated (MWh/yr) 4870.30

208
Table B-17: Modelling of biogas-based Power Plant in 30 t/hr mill
Mill Processing Capacity (t/hr) 30
Mill Operation Hour (hrs/yr) 8000
Amount of FFB process (t/yr) 240000
Amount POME Available (m3/yr) 156000
COD POME (mg/l) 51000
COD Converted (m3/yr) 6364800
COD Converted (t/yr) 6364.8
Methane Produced (t/yr) 1432.08
Feedstock Feed rate (kg/s) 0.0497
Calorific Value (MJ/kg) 50.0000

Mathematical Modelling
Extractable Energy from Residues (MW) 2.24
Mass Flow Rate of Superheated Steam (kg/s) 0.76
Total Superheated Steam Generated (t/yr) 21884
Total Available Steam for Electricity Generation (t/yr) 21884
Installed Electricity Generation Capacity (MW) 0.91
Total Electricity Generated (MWh/yr) 7305.45

Table B-18: Modelling of biogas-based Power Plant in 40 t/hr mill


Mill Processing Capacity (t/hr) 40
Mill Operation Hour (hrs/yr) 8000
Amount of FFB process (t/yr) 320000
Amount POME Available (m3/yr) 208000
COD POME (mg/l) 51000
COD Converted (m3/yr) 8486400
COD Converted (t/yr) 8486.4
Methane Produced (t/yr) 1909.44
Feedstock Feed rate (kg/s) 0.0663
Calorific Value (MJ/kg) 50.0000

Mathematical Modelling
Extractable Energy from Residues (MW) 2.98
Mass Flow Rate of Superheated Steam (kg/s) 1.01
Total Superheated Steam Generated (t/yr) 29178
Total Available Steam for Electricity Generation (t/yr) 29178
Installed Electricity Generation Capacity (MW) 1.22
Total Electricity Generated (MWh/yr) 9740.60

209
Table B-19: Modelling of biogas-based Power Plant in 50 t/hr mill
Mill Processing Capacity (t/hr) 50
Mill Operation Hour (hrs/yr) 8000
Amount of FFB process (t/yr) 400000
Amount POME Available (m3/yr) 260000
COD POME (mg/l) 51000
COD Converted (m3/yr) 10608000
COD Converted (t/yr) 10608
Methane Produced (t/yr) 2386.8
Feedstock Feed rate (kg/s) 0.0829
Calorific Value (MJ/kg) 50.0000

Mathematical Modelling
Extractable Energy from Residues (MW) 3.73
Mass Flow Rate of Superheated Steam (kg/s) 1.27
Total Superheated Steam Generated (t/yr) 36473
Total Available Steam for Electricity Generation (t/yr) 36473
Installed Electricity Generation Capacity (MW) 1.52
Total Electricity Generated (MWh/yr) 12175.75

Table B-20: Modelling of biogas-based Power Plant in 60 t/hr mill


Mill Processing Capacity (t/hr) 60
Mill Operation Hour (hrs/yr) 8000
Amount of FFB process (t/yr) 480000
Amount POME Available (m3/yr) 312000
COD POME (mg/l) 51000
COD Converted (m3/yr) 12729600
COD Converted (t/yr) 12729.6
Methane Produced (t/yr) 2864.16
Feedstock Feed rate (kg/s) 0.0995
Calorific Value (MJ/kg) 50.0000

Mathematical Modelling
Extractable Energy from Residues (MW) 4.48
Mass Flow Rate of Superheated Steam (kg/s) 1.52
Total Superheated Steam Generated (t/yr) 43767
Total Available Steam for Electricity Generation (t/yr) 43767
Installed Electricity Generation Capacity (MW) 1.83
Total Electricity Generated (MWh/yr) 14610.90

210
Table B-21: Modelling of biogas-based Power Plant in 70 t/hr mill
Mill Processing Capacity (t/hr) 70
Mill Operation Hour (hrs/yr) 8000
Amount of FFB process (t/yr) 560000
Amount POME Available (m3/yr) 364000
COD POME (mg/l) 51000
COD Converted (m3/yr) 14851200
COD Converted (t/yr) 14851.2
Methane Produced (t/yr) 3341.52
Feedstock Feed rate (kg/s) 0.1160
Calorific Value (MJ/kg) 50.0000

Mathematical Modelling
Extractable Energy from Residues (MW) 5.22
Mass Flow Rate of Superheated Steam (kg/s) 1.77
Total Superheated Steam Generated (t/yr) 51062
Total Available Steam for Electricity Generation (t/yr) 51062
Installed Electricity Generation Capacity (MW) 2.13
Total Electricity Generated (MWh/yr) 17046.05

Table B-22: Modelling of biogas-based Power Plant in 80 t/hr mill


Mill Processing Capacity (t/hr) 80
Mill Operation Hour (hrs/yr) 8000
Amount of FFB process (t/yr) 640000
Amount POME Available (m3/yr) 416000
COD POME (mg/l) 51000
COD Converted (m3/yr) 16972800
COD Converted (t/yr) 16972.8
Methane Produced (t/yr) 3818.88
Feedstock Feed rate (kg/s) 0.1326
Calorific Value (MJ/kg) 50.0000

Mathematical Modelling
Extractable Energy from Residues (MW) 5.97
Mass Flow Rate of Superheated Steam (kg/s) 2.03
Total Superheated Steam Generated (t/yr) 58356
Total Available Steam for Electricity Generation (t/yr) 58356
Installed Electricity Generation Capacity (MW) 2.44
Total Electricity Generated (MWh/yr) 19481.20

211
Table B-23: Modelling of biogas-based Power Plant in 90 t/hr mill
Mill Processing Capacity (t/hr) 90
Mill Operation Hour (hrs/yr) 8000
Amount of FFB process (t/yr) 720000
Amount POME Available (m3/yr) 468000
COD POME (mg/l) 51000
COD Converted (m3/yr) 19094400
COD Converted (t/yr) 19094.4
Methane Produced (t/yr) 4296.24
Feedstock Feed rate (kg/s) 0.1492
Calorific Value (MJ/kg) 50.0000

Mathematical Modelling
Extractable Energy from Residues (MW) 6.71
Mass Flow Rate of Superheated Steam (kg/s) 2.28
Total Superheated Steam Generated (t/yr) 65651
Total Available Steam for Electricity Generation (t/yr) 65651
Installed Electricity Generation Capacity (MW) 2.74
Total Electricity Generated (MWh/yr) 21916.35

Table B-24: Modelling of biogas-based Power Plant in 100 t/hr mill


Mill Processing Capacity (t/hr) 100
Mill Operation Hour (hrs/yr) 8000
Amount of FFB process (t/yr) 800000
Amount POME Available (m3/yr) 520000
COD POME (mg/l) 51000
COD Converted (m3/yr) 21216000
COD Converted (t/yr) 21216
Methane Produced (t/yr) 4773.6
Feedstock Feed rate (kg/s) 0.1658
Calorific Value (MJ/kg) 50.0000

Mathematical Modelling
Extractable Energy from Residues (MW) 5.97
Mass Flow Rate of Superheated Steam (kg/s) 2.03
Total Superheated Steam Generated (t/yr) 58356
Total Available Steam for Electricity Generation (t/yr) 58356
Installed Electricity Generation Capacity (MW) 2.44
Total Electricity Generated (MWh/yr) 19481.20

212
Table B-25: Modelling of biogas-based Power Plant in 110 t/hr mill
Mill Processing Capacity (t/hr) 110
Mill Operation Hour (hrs/yr) 8000
Amount of FFB process (t/yr) 880000
Amount POME Available (m3/yr) 572000
COD POME (mg/l) 51000
COD Converted (m3/yr) 23337600
COD Converted (t/yr) 23337.6
Methane Produced (t/yr) 5250.96
Feedstock Feed rate (kg/s) 0.1823
Calorific Value (MJ/kg) 50.0000

Mathematical Modelling
Extractable Energy from Residues (MW) 8.20
Mass Flow Rate of Superheated Steam (kg/s) 2.79
Total Superheated Steam Generated (t/yr) 80240
Total Available Steam for Electricity Generation (t/yr) 80240
Installed Electricity Generation Capacity (MW) 3.35
Total Electricity Generated (MWh/yr) 26786.65

Table B-26: Modelling of biogas-based Power Plant in 120 t/hr mill


Mill Processing Capacity (t/hr) 120
Mill Operation Hour (hrs/yr) 8000
Amount of FFB process (t/yr) 960000
Amount POME Available (m3/yr) 624000
COD POME (mg/l) 51000
COD Converted (m3/yr) 25459200
COD Converted (t/yr) 25459.2
Methane Produced (t/yr) 5728.32
Feedstock Feed rate (kg/s) 0.1989
Calorific Value (MJ/kg) 50.0000

Mathematical Modelling
Extractable Energy from Residues (MW) 8.95
Mass Flow Rate of Superheated Steam (kg/s) 3.04
Total Superheated Steam Generated (t/yr) 87534
Total Available Steam for Electricity Generation (t/yr) 87534
Installed Electricity Generation Capacity (MW) 3.65
Total Electricity Generated (MWh/yr) 29221.80
213
Appendix B-4 – Dryer Enthalpy Balance and Kinematics Modelling Spreadsheet

The spreadsheet used to estimate the heat requirement for drying the biomass to reduce its
moisture content from 60% to 11% are presented in Table B-27 to Table B-38. The detail
description of the parameter used in the spreadsheet is given in Table B-26.

Table B-27: Description of Data Used for Modelling the EFB Dryer
Data Types Unit Remarks
Mass flow rate of EFB on dried kg/s The 𝑚𝑑𝑎𝑓 was calculated by dividing the amount of
and ash free basis (𝑚𝑑𝑎𝑓 ) available dried EFB in the mills (35% of the raw EFB)
with the operation hours.
Mass flow rate of ash (𝑚𝑎𝑠ℎ ) kg/s The 𝑚𝑎𝑠ℎ was calculated by dividing the amount of ash
(5% of the raw EFB) with the operation hours.
Mass of water remaining in kg The amount of water that needs to be removed from
EFB (𝑀𝑊,𝐸𝐹𝐵 ) EFB until it reached the final moisture content of 11%
was identified using Eq. (9).
Water in EFB Evaporation kg/s The evaporation rate of the water in the EFB is
Rate calculated using Eq. (10).
Sensible Heat MW The amount of heat that required to change the
temperature for the drying process is calculated using
Eq. (7). The modelling parameters from Luk et al.
(2013) and Gereegziabher, Oyedun and Hui (2013) are
used in the calculation.

Latent Heat MW The amount of heat that absorbed by the EFB during
the drying process due to its moisture content is
calculated using Eq. (8).

Table B-28: Heat Requirement to Dry EFB in 10 t/hr mill

Mass flow rate of EFB on dried and ash free basis (kg/s) 0.15
Mass flow rate of ash (kg/s) 0.02
Mass of water remaining in EFB (kg) 0.02
Water in EFB Evaporation Rate (kg/s) 0.21
Sensible Heat (MW) 0.03
Latent Heat(MW) 0.57
Total Heat (MW) 0.60

214
Table B-29: Heat Requirement to Dry EFB in 20 t/hr mill

Mass flow rate of EFB on dried and ash free basis (kg/s) 0.29
Mass flow rate of ash (kg/s) 0.04
Mass of water remaining in EFB (kg) 0.04
Water in EFB Evaporation Rate (kg/s) 0.40
Sensible Heat (MW) 0.05
Latent Heat(MW) 1.25
Total Heat (MW) 1.30

Table B-30: Heat Requirement to Dry EFB in 30 t/hr mill

Mass flow rate of EFB on dried and ash free basis (kg/s) 0.44
Mass flow rate of ash (kg/s) 0.06
Mass of water remaining in EFB (kg) 0.06
Water in EFB Evaporation Rate (kg/s) 0.65
Sensible Heat (MW) 0.07
Latent Heat(MW) 1.73
Total Heat (MW) 1.80

Table B-31: Heat Requirement to Dry EFB in 40 t/hr mill


Mass flow rate of EFB on dried and ash free basis (kg/s) 0.58
Mass flow rate of ash (kg/s) 0.08
Mass of water remaining in EFB (kg) 0.08
Water in EFB Evaporation Rate (kg/s) 0.83
Sensible Heat (MW) 0.10
Latent Heat(MW) 2.20
Total Heat (MW) 2.30

215
Table B-32: Heat Requirement to Dry EFB in 50 t/hr mill

Mass flow rate of EFB on dried and ash free basis (kg/s) 0.73
Mass flow rate of ash (kg/s) 0.10
Mass of water remaining in EFB (kg) 0.10
Water in EFB Evaporation Rate (kg/s) 1.00
Sensible Heat (MW) 0.12
Latent Heat(MW) 2.68
Total Heat (MW) 2.80

Table B-33: Heat Requirement to Dry EFB in 60 t/hr mill


.
Mass flow rate of EFB on dried and ash free basis (kg/s) 0.88
Mass flow rate of ash (kg/s) 0.13
Mass of water remaining in EFB (kg) 0.12
Water in EFB Evaporation Rate (kg/s) 1.15
Sensible Heat (MW) 0.14
Latent Heat(MW) 3.16
Total Heat (MW) 3.20

216
Table B-34: Heat Requirement to Dry EFB in 70 t/hr mill

Mass flow rate of EFB on dried and ash free basis (kg/s) 1.02
Mass flow rate of ash (kg/s) 0.15
Mass of water remaining in EFB (kg) 0.14
Water in EFB Evaporation Rate (kg/s) 1.29
Sensible Heat (MW) 0.17
Latent Heat(MW) 3.53
Total Heat (MW) 3.70

Table B-35: Heat Requirement to Dry EFB in 80 t/hr mill

Mass flow rate of EFB on dried and ash free basis (kg/s) 1.17
Mass flow rate of ash (kg/s) 0.17
Mass of water remaining in EFB (kg) 0.16
Water in EFB Evaporation Rate (kg/s) 1.41
Sensible Heat (MW) 0.19
Latent Heat(MW) 3.81
Total Heat (MW) 4.00

Table B-36: Heat Requirement to Dry EFB in 90 t/hr mill

Mass flow rate of EFB on dried and ash free basis (kg/s) 1.31
Mass flow rate of ash (kg/s) 0.19
Mass of water remaining in EFB (kg/s) 0.19
Water in EFB Evaporation Rate (kg/s) 1.54
Sensible Heat (MW) 0.20
Latent Heat(MW) 4.20
Total Heat (MW) 4.40

Table B-37: Heat Requirement to Dry EFB in 100 t/hr mill

Mass flow rate of EFB on dried and ash free basis (kg/s) 1.46
Mass flow rate of ash (kg/s) 0.20
Mass of water remaining in EFB (kg) 0.20
Water in EFB Evaporation Rate (kg/s) 1.65
Sensible Heat (MW) 0.24
Latent Heat(MW) 4.46
Total Heat (MW) 4.70

217
Table B-38: Heat Requirement to Dry EFB in 110 t/hr mill

Mass flow rate of EFB on dried and ash free basis (kg/s) 1.60
Mass flow rate of ash (kg/s) 0.23
Mass of water remaining in EFB (kg) 0.23
Water in EFB Evaporation Rate (kg/s) 1.75
Sensible Heat (MW) 0.26
Latent Heat(MW) 4.74
Total Heat (MW) 5.00

Table B-39: Heat Requirement to Dry EFB in 120 t/hr mill

Mass flow rate of EFB on dried and ash free basis (kg/s) 1.75
Mass flow rate of ash (kg/s) 0.25
Mass of water remaining in EFB (kg) 0.25
Water in EFB Evaporation Rate (kg/s) 1.85
Sensible Heat (MW) 0.29
Latent Heat(MW) 5.01
Total Heat (MW) 5.30

Appendix B-5 – Mathematical Modelling Spreadsheet for Dried EFB Power Plant

The power plant was modelled following the procedure in Appendix B-3 for the semi-dry EFB-
based power plant. The CV of the dried EFB was used to estimate the amount of heat extracted
from the combustion process with an efficiency of 90%.

Firstly, the spreadsheet assumed that the amount of heat required for drying the EFB was
channelled from the heat extracted from the combustion process. A 16-25% reduction in the
amount of heat available for electricity generation was observed when a certain amount of heat
from the combustion was used for drying the EFB. This reduction was found to be similar to
what was expressed by Luk et al. (2013). Next, the modelling process was repeated using the
second assumption where the flue gas is used for drying the EFB. The heat extracted from the
combustion process was used for the electricity generation process.

218
Appendix C

This appendix compiles the technical and techno-economic assessment spreadsheets used to
assess the technical and economic performance of the POMR-RES.

Appendix C-1 – Technical Assessment Spreadsheet

The technical assessment spreadsheet in Tables C1 – C12 were used to assess the feasibility
of the POMR-RES according to Eq. (13).

Table C-1: Technical Feasibility Assessment of 0.9 MW Power Plant


Total Electricity Generated (MWh/yr) 7122
Mill Electricity Demand (MWh/yr) 2160
Parasitic Load of Generation Plant (MWh/yr) 783
Net Electricity Available for the Grid (MWh/yr) 4179

Table C-2: Technical Feasibility Assessment of 1.8 MW Power Plant


Total Electricity Generated (MWh/yr) 14246
Mill Electricity Demand (MWh/yr) 4320
Parasitic Load of Generation Plant (MWh/yr) 1567
Net Electricity Available for the Grid (MWh/yr) 8359

Table C-3: Technical Feasibility Assessment of 2.7 MW Power Plant


Total Electricity Generated (MWh/yr) 21368
Mill Electricity Demand (MWh/yr) 6480
Parasitic Load of Generation Plant (MWh/yr) 2351
Net Electricity Available for the Grid (MWh/yr) 12538

Table C-4: Technical Feasibility Assessment of 3.6 MW Power Plant


Total Electricity Generated (MWh/yr) 28491
Mill Electricity Demand (MWh/yr) 8640
Parasitic Load of Generation Plant (MWh/yr) 3134
Net Electricity Available for the Grid (MWh/yr) 16717

Table C-5: Technical Feasibility Assessment of 4.4 MW Power Plant


Total Electricity Generated (MWh/yr) 35614
Mill Electricity Demand (MWh/yr) 10800
Parasitic Load of Generation Plant (MWh/yr) 3918
Net Electricity Available for the Grid (MWh/yr) 20896

219
Table C-6: Technical Feasibility Assessment of 5.3 MW Power Plant
Total Electricity Generated (MWh/yr) 42737
Mill Electricity Demand (MWh/yr) 12960
Parasitic Load of Generation Plant (MWh/yr) 4701
Net Electricity Available for the Grid (MWh/yr) 25076

Table C-7: Technical Feasibility Assessment of 6.2 MW Power Plant


Total Electricity Generated (MWh/yr) 49859
Mill Electricity Demand (MWh/yr) 15120
Parasitic Load of Generation Plant (MWh/yr) 5485
Net Electricity Available for the Grid (MWh/yr) 29255

Table C-8: Technical Feasibility Assessment of 7.1 MW Power Plant


Total Electricity Generated (MWh/yr) 56982
Mill Electricity Demand (MWh/yr) 17280
Parasitic Load of Generation Plant (MWh/yr) 6268
Net Electricity Available for the Grid (MWh/yr) 33434

Table C-9: Technical Feasibility Assessment of 8.0 MW Power Plant


Total Electricity Generated (MWh/yr) 64105
Mill Electricity Demand (MWh/yr) 19440
Parasitic Load of Generation Plant (MWh/yr) 7052
Net Electricity Available for the Grid (MWh/yr) 37613

Table C-10: Technical Feasibility Assessment of 8.9 MW Power Plant


Total Electricity Generated (MWh/yr) 71228
Mill Electricity Demand (MWh/yr) 21600
Parasitic Load of Generation Plant (MWh/yr) 7835
Net Electricity Available for the Grid (MWh/yr) 41793

Table C-11: Technical Feasibility Assessment of 9.8 MW Power Plant


Total Electricity Generated (MWh/yr) 78350
Mill Electricity Demand (MWh/yr) 23760
Parasitic Load of Generation Plant (MWh/yr) 8619
Net Electricity Available for the Grid (MWh/yr) 45972

Table C-12: Technical Feasibility Assessment of 10.7 MW Power Plant


Total Electricity Generated (MWh/yr) 85473
Mill Electricity Demand (MWh/yr) 25920
Parasitic Load of Generation Plant (MWh/yr) 9402
Net Electricity Available for the Grid (MWh/yr) 50151

220
Appendix C-2 – Techno-Economic Spreadsheet

The techno-economic assessment spreadsheet consists of six sections: (1) capital cost
estimation; (2) operation cost estimation; (3) revenue interpretation; (4) ROI and payback
period analysis; and (5) LCOE estimation. The techno-economic assessment was conducted
using the second cost estimation method which is the direct cost to capacity correlation
estimation. A cost correlation of US$ 2.3 million/MW was used to estimate the capital cost for
the POMR-RES. The spreadsheet in Tables C13 – C24 present the techno-economic
assessment for the on-site private-funded EFB-based power plant and Tables C25 – C36 for
the self-funded power plant. It can be seen that most of the private-funded power plant has a
negative net cash flow indicating that the cash outflows are higher than the cash inflows. The
negative net cash flow also demonstrated that the installation of the private-funded power plant
is not economically viable. The sensitivity analysis for the power plant was carried out using
the same spreadsheets by changing the sensitivity variables according to the analysis
requirements. The procedure to conduct the techno-economic assessment was repeated for on-
site CHP plant, off-site POMR-RES installation, co-operative POMR-RES installation, and
dried EFB-based power plant. For the dried EFB-based power plant, the installation cost for
the dryer and the operating cost of the drying process were included in the capital cost
estimation and operation cost estimation section respectively. The feedstock cost and feedstock
transportation costs were added as the variable operating cost when assessing the techno-
economic performance of the off-site POMR-RES.

221
Table C-13: Techno-Economic Assessment of the Private-Funded 0.9 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 0.90 2.11 2.10
Total Delivered Cost of Equipment (mill US$) 2.10

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 3.31

Fixed Interest Rate (%) 0.12


No of Years 15
Interest Payment (mil US$) 5.95
Annualization Factor 0.15
Annual Capital Cost (mil US$) 0.49
Annual Interest Payment (mil US$) 0.87

Fixed Cost

No Items Amount
1 10% of Capital Cost (mil US$/yr) 0.33
Total Fixed Cost (mil US$) 0.33
Variable Cost
No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0
Total Operation Cost (mil US$) 0.33
Total Annual Cost with Grid Cost
(mil US$) 1.69

222
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 0.45


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 0.22
Total Revenue (mil US$/y) 0.67

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI -473

Payback Period Analysis

Payback Period of 0.9 MW


0.0000
-2.0000 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Net Cash Flow (mil US$)

-4.0000
-6.0000
-8.0000
-10.0000
-12.0000
-14.0000
-16.0000
-18.0000
-20.0000
Year

LCOE Estimation

Annual Cost (mil US$) 1.69


Electricity Generated (kWh/yr) 92500000.00
LCOE (US$/kWh) 0.18

223
Table C-14: Techno-Economic Assessment of the Private-Funded 1.8 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 1.80 3.68 3.66
Total Delivered Cost of Equipment (mill US$) 3.66

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 4.87

Fixed Interest Rate (%) 0.12


No of Years 15
Interest Payment (mil US$) 8.76
Annualization Factor 0.15
Annual Capital Cost (mil US$) 0.71
Annual Interest Payment (mil US$) 1.29

Fixed Cost
No Items Amount
1 10% of Capital Cost (mil US$/yr) 0.49
Total Fixed Cost (mil US$) 0.49
Variable Cost
No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0
Total Operation Cost (mil US$) 0.49
Total Annual Cost with Grid Cost
(mil US$) 2.49

224
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 0.90


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 0.44
Total Revenue (mil US$/y) 1.33

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI -401

Payback Period Analysis

Payback Period 1.8 MW


0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Net Cash Flow (mil US$)

-5.0000

-10.0000

-15.0000

-20.0000

-25.0000
Year

LCOE Estimation

Annual Cost (mil US$) 2.48


Electricity Generated (kWh/yr) 18500000.00
LCOE (US$/kWh) 0.13

225
Table C-15: Techno-Economic Assessment of the Private-Funded 2.7 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 2.70 5.09 5.07
Total Delivered Cost of Equipment (mill US$) 5.07

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 6.27

Fixed Interest Rate (%) 0.12


No of Years 15
Interest Payment (mil US$) 11.29
Annualization Factor 0.15
Annual Capital Cost (mil US$) 0.92
Annual Interest Payment (mil US$) 1.66

Fixed Cost
No Items Amount
1 10% of Capital Cost (mil US$/yr) 0.63
Total Fixed Cost (mil US$) 0.63
Variable Cost
No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 0.63


Total Annual Cost with Grid Cost
(mil US$) 3.21

226
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 1.35


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 0.65
Total Revenue (mil US$/y) 2.00

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI -357

Payback Period Analysis

Payback Period 2.7 MW


0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

-5.0000
Net Cash Flow (mil US$)

-10.0000

-15.0000

-20.0000

-25.0000
Year

LCOE Estimation

Annual Cost (mil US$) 3.20


Electricity Generated (kWh/yr) 21368000.00
LCOE (US$/kWh) 0.15

227
Table C-16: Techno-Economic Assessment of the Private-Funded 3.6 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 3.6 6.41 6.38
Total Delivered Cost of Equipment (mill US$) 6.38

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 7.58

Fixed Interest Rate (%) 0.12


No of Years 15
Interest Payment (mil US$) 13.65
Annualization Factor 0.15
Annual Capital Cost (mil US$) 1.11
Annual Interest Payment (mil US$) 2.00

Fixed Cost
No Items Amount
1 10% of Capital Cost (mil US$/yr) 0.76
Total Fixed Cost (mil US$) 0.76
Variable Cost
No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 0.76


Total Annual Cost with Grid Cost
(mil US$) 3.88

228
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 1.80


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 0.87
Total Revenue (mil US$/y) 2.67

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI -324

Payback Period Analysis

Payback Period 3.6 MW


0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-5.0000
Net Cash Flow (mil US$)

-10.0000

-15.0000

-20.0000

-25.0000

-30.0000
Year

LCOE Estimation

Annual Cost (mil US$) 3.90


Electricity Generated (kWh/yr) 37001000.00
LCOE (US$/kWh) 0.10

229
Table C-17: Techno-Economic Assessment of the Private-Funded 4.4 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 4.40 7.52 7.49
Total Delivered Cost of Equipment (mill US$) 7.49

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 8.69

Fixed Interest Rate (%) 0.12


No of Years 15
Interest Payment (mil US$) 15.65
Annualization Factor 0.15
Annual Capital Cost (mil US$) 1.28
Annual Interest Payment (mil US$) 2.30

Fixed Cost
No Items Amount
1 10% of Capital Cost (mil US$/yr) 0.87
Total Fixed Cost (mil US$) 0.87
Variable Cost
No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 0.87


Total Annual Cost with Grid Cost
(mil US$) 4.44

230
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 2.25


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 1.09
Total Revenue (mil US$/y) 3.34

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI -293

Payback Period Analysis

Payback Period 4.4 MW


0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Net Cash Flow (mil US$)

-5.0000

-10.0000

-15.0000

-20.0000

-25.0000

-30.0000
Year

LCOE Estimation

Annual Cost (mil US$) 4.44


Electricity Generated (kWh/yr) 46251000.00
LCOE (US$/kWh) 0.09

231
Table C-18: Techno-Economic Assessment of the Private-Funded 5.3 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 5.30 8.73 8.69
Total Delivered Cost of Equipment (mill US$) 8.69

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 9.90

Fixed Interest Rate (%) 0.12


No of Years 15
Interest Payment (mil US$) 17.81
Annualization Factor 0.15
Annual Capital Cost (mil US$) 1.45
Annual Interest Payment (mil US$) 2.62

Fixed Cost
No Items Amount
1 10% of Capital Cost (mil US$/yr) 0.99
Total Fixed Cost (mil US$) 0.99
Variable Cost
No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 0.99


Total Annual Cost with Grid Cost
(mil US$) 5.06

232
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 2.70


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 1.30
Total Revenue (mil US$/y) 4.00

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI -272

Payback Period Analysis

Payback Period 5.3 MW


0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-5.0000
Net Cash Flow (mil US$)

-10.0000

-15.0000

-20.0000

-25.0000

-30.0000
Year

LCOE Estimation

Annual Cost (mil US$) 4.44


Electricity Generated (kWh/yr) 46251000.00
LCOE (US$/kWh) 0.10

233
Table C-19: Techno-Economic Assessment of the Private-Funded 6.2 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 6.20 9.90 9.86
Total Delivered Cost of Equipment (mill US$) 9.86

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 11.06

Fixed Interest Rate (%) 0.12


No of Years 15
Interest Payment (mil US$) 19.91
Annualization Factor 0.15
Annual Capital Cost (mil US$) 1.62
Annual Interest Payment (mil US$) 2.92

Fixed Cost
No Items Amount
1 10% of Capital Cost (mil US$/yr) 1.11
Total Fixed Cost (mil US$) 1.11
Variable Cost
No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 1.11


Total Annual Cost with Grid Cost
(mil US$) 5.65

234
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 3.14


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 1.53
Total Revenue (mil US$/y) 4.67

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI -254

Payback Period Analysis

Payback Period 6.2 MW


0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-5.0000
Net Cash Flow (mil US$)

-10.0000

-15.0000

-20.0000

-25.0000

-30.0000
Year

LCOE Estimation

Annual Cost (mil US$) 5.65


Electricity Generated (kWh/yr) 64752000.00
LCOE (US$/kWh) 0.09

235
Table C-20: Techno-Economic Assessment of the Private-Funded 7.1 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 7.10 11.03 10.99
Total Delivered Cost of Equipment (mill US$) 10.99

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 12.19

Fixed Interest Rate (%) 0.12


No of Years 15
Interest Payment (mil US$) 21.94
Annualization Factor 0.15
Annual Capital Cost (mil US$) 1.79
Annual Interest Payment (mil US$) 3.22

Fixed Cost
No Items Amount
1 10% of Capital Cost (mil US$/yr) 1.22
Total Fixed Cost (mil US$) 1.22
Variable Cost
No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 1.22


Total Annual Cost with Grid Cost
(mil US$) 6.23

236
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 3.59


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 1.75
Total Revenue (mil US$/y) 5.34

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI -239

Payback Period Analysis

Payback Period 7.1 MW


0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-5.0000
Net Cash Flow (mil US$)

-10.0000

-15.0000

-20.0000

-25.0000

-30.0000
Year

LCOE Estimation

Annual Cost (mil US$) 6.23


Electricity Generated (kWh/yr) 74002000.00
LCOE (US$/kWh) 0.08

237
Table C-21: Techno-Economic Assessment of the Private-Funded 8.0 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 8.00 12.14 12.09
Total Delivered Cost of Equipment (mill US$) 12.09

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 13.29

Fixed Interest Rate (%) 0.12


No of Years 15
Interest Payment (mil US$) 23.92
Annualization Factor 0.15
Annual Capital Cost (mil US$) 1.95
Annual Interest Payment (mil US$) 3.51

Fixed Cost
No Items Amount
1 10% of Capital Cost (mil US$/yr) 1.33
Total Fixed Cost (mil US$) 1.33
Variable Cost
No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 1.33


Total Annual Cost with Grid Cost
(mil US$) 6.79

238
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 4.04


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 1.96
Total Revenue (mil US$/y) 6.00

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI -224

Payback Period Analysis

Payback Period 8.0 MW


0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-5.0000
Net Cash Flow (mil US$)

-10.0000

-15.0000

-20.0000

-25.0000

-30.0000
Year

LCOE Estimation

Annual Cost (mil US$) 6.79


Electricity Generated (kWh/yr) 83253000.00
LCOE (US$/kWh) 0.08

239
Table C-22: Techno-Economic Assessment of the Private-Funded 8.9 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 8.90 13.22 13.16
Total Delivered Cost of Equipment (mill US$) 13.16

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 14.36

Fixed Interest Rate (%) 0.12


No of Years 15
Interest Payment (mil US$) 25.86
Annualization Factor 0.15
Annual Capital Cost (mil US$) 2.11
Annual Interest Payment (mil US$) 3.80

Fixed Cost
No Items Amount
1 10% of Capital Cost (mil US$/yr) 1.44
Total Fixed Cost (mil US$) 1.44
Variable Cost
No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 1.44


Total Annual Cost with Grid Cost
(mil US$) 7.34

240
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 4.49


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 2.18
Total Revenue (mil US$/y) 6.67

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI -212

Payback Period Analysis

Payback Period 8.9 MW


0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

-5.0000
Net Cash Flow (mil US$)

-10.0000

-15.0000

-20.0000

-25.0000
Years

LCOE Estimation

Annual Cost (mil US$) 7.34


Electricity Generated (kWh/yr) 92503000.00
LCOE (US$/kWh) 0.08

241
Table C-23: Techno-Economic Assessment of the Private-Funded 9.8 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 9.80 14.28 14.22
Total Delivered Cost of Equipment (mill US$) 14.22

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 15.47

Fixed Interest Rate (%) 0.12


No of Years 15
Interest Payment (mil US$) 27.84
Annualization Factor 0.15
Annual Capital Cost (mil US$) 2.27
Annual Interest Payment (mil US$) 4.09

Fixed Cost
No Items Amount
1 10% of Capital Cost (mil US$/yr) 1.55
Total Fixed Cost (mil US$) 1.55
Variable Cost
No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 1.55


Total Annual Cost with Grid Cost
(mil US$) 7.91

242
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 4.94


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 2.40
Total Revenue (mil US$/y) 7.34

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI -202

Payback Period Analysis

Payback Period 9.8 MW


0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

-5.0000
Net Cash Flow (mil US$)

-10.0000

-15.0000

-20.0000

-25.0000
Years

LCOE Estimation

Annual Cost (mil US$) 7.91


Electricity Generated (kWh/yr) 101753000.00
LCOE (US$/kWh) 0.08

243
Table C-24: Techno-Economic Assessment of the Private-Funded 10.7 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 10.70 15.32 15.25
Total Delivered Cost of Equipment (mill US$) 15.25

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 16.45

Fixed Interest Rate (%) 0.12


No of Years 15
Interest Payment (mil US$) 26.62
Annualization Factor 0.15
Annual Capital Cost (mil US$) 2.42
Annual Interest Payment (mil US$) 4.35

Fixed Cost
No Items Amount
1 10% of Capital Cost (mil US$/yr) 1.65
Total Fixed Cost (mil US$) 1.65
Variable Cost
No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 1.65


Total Annual Cost with Grid Cost
(mil US$) 8.41

244
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 5.16


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 2.62
Total Revenue (mil US$/y) 7.78

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI -204

Payback Period Analysis

Payback Period 10.7 MW


0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-5.0000
Net Cash Flow (mil US$)

-10.0000

-15.0000

-20.0000

-25.0000

-30.0000
Years

LCOE Estimation

Annual Cost (mil US$) 8.41


Electricity Generated (kWh/yr) 111004000.00
LCOE (US$/kWh) 0.08

245
Table C-25: Techno-Economic Assessment of the Self-Funded 0.9 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 0.90 2.11 2.10
Total Delivered Cost of Equipment (mill US$) 2.10

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 3.31

No of Years 15
Annual Capital Cost (mil US$) 0.22

Fixed Cost
No Items Amount
1 10% of Capital Cost (mil US$/yr) 0.33
Total Fixed Cost (mil US$) 0.33
Variable Cost
No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 0.33


Total Annual Cost with Grid Cost
(mil US$) 0.55

246
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 0.45


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 0.22
Total Revenue (mil US$/y) 0.67

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI -141

Payback Period Analysis

Payback Period 0.9 MW


0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-0.5000
Net Cash Flow (mil US$)

-1.0000

-1.5000

-2.0000

-2.5000

-3.0000

-3.5000
Year

LCOE Estimation

Annual Cost (mil US$) 0.55


Electricity Generated (kWh/yr) 92500000.00
LCOE (US$/kWh) 0.06

247
Table C-26: Techno-Economic Assessment of the Self-Funded 1.8 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 1.80 3.68 3.66
Total Delivered Cost of Equipment (mill US$) 3.66

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 4.87

No of Years 15
Annual Capital Cost (mil US$) 0.32

Fixed Cost

No Items Amount
1 10% of Capital Cost (mil US$/yr) 0.49
Total Fixed Cost (mil US$) 0.49
Variable Cost

No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 0.49


Total Annual Cost with Grid Cost
(mil US$) 0.81

248
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 0.90


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 0.44
Total Revenue (mil US$/y) 1.33

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI -47

Payback Period Analysis

Payback Period 1.8 MW


3.0000

2.0000
Net Cash Flow (mil US$)

1.0000

0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-1.0000

-2.0000

-3.0000

-4.0000

-5.0000
Years

LCOE Estimation

Annual Cost (mil US$) 0.81


Electricity Generated (kWh/yr) 18500000.00
LCOE (US$/kWh) 0.04

249
Table C-27: Techno-Economic Assessment of the Self-Funded 2.7 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 2.70 5.09 5.07
Total Delivered Cost of Equipment (mill US$) 5.07

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 6.27

No of Years 15
Annual Capital Cost (mil US$) 0.42

Fixed Cost

No Items Amount
1 10% of Capital Cost (mil US$/yr) 0.63
Total Fixed Cost (mil US$) 0.63
Variable Cost

No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 0.63


Total Annual Cost with Grid Cost
(mil US$) 1.05

250
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 1.35


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 0.65
Total Revenue (mil US$/y) 2.00

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI 11

Payback Period Analysis

Payback Period 2.7 MW


8.0000

6.0000
Net Cash Flow (mil US$)

4.0000

2.0000

0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-2.0000

-4.0000

-6.0000
Year

LCOE Estimation

Annual Cost (mil US$) 1.05


Electricity Generated (kWh/yr) 21368000.00
LCOE (US$/kWh) 0.04

251
Table C-28: Techno-Economic Assessment of the Self-Funded 3.6 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 3.6 6.41 6.38
Total Delivered Cost of Equipment (mill US$) 6.38

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 7.58

No of Years 15
Annual Capital Cost (mil US$) 0.51

Fixed Cost

No Items Amount
1 10% of Capital Cost (mil US$/yr) 0.76
Total Fixed Cost (mil US$) 0.76
Variable Cost

No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 0.76


Total Annual Cost with Grid Cost
(mil US$) 1.26

252
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 1.80


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 0.87
Total Revenue (mil US$/y) 2.67

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI 53

Payback Period Analysis

Payback Period 3.6 MW


15.0000
Net Cash Flow (mil US$)

10.0000

5.0000

0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-5.0000

-10.0000
Year

LCOE Estimation

Annual Cost (mil US$) 1.26


Electricity Generated (kWh/yr) 37001000.00
LCOE (US$/kWh) 0.03

253
Table C-29: Techno-Economic Assessment of the Self-Funded 4.4 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 4.40 7.52 7.49
Total Delivered Cost of Equipment (mill US$) 7.49

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 8.69

No of Years 15
Annual Capital Cost (mil US$) 0.58

Fixed Cost

No Items Amount
1 10% of Capital Cost (mil US$/yr) 0.87
Total Fixed Cost (mil US$) 0.87
Variable Cost

No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 0.87


Total Annual Cost with Grid Cost
(mil US$) 1.45

254
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 2.25


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 1.09
Total Revenue (mil US$/y) 3.34

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI 94

Payback Period Analysis

Payback Period 4.4 MW


20.0000

15.0000
Net Cash Flow (mil US$)

10.0000

5.0000

0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-5.0000

-10.0000
Year

LCOE Estimation

Annual Cost (mil US$) 1.45


Electricity Generated (kWh/yr) 46251000.00
LCOE (US$/kWh) 0.03

255
Table C-30: Techno-Economic Assessment of the Self-Funded 5.3 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 5.30 8.73 8.69
Total Delivered Cost of Equipment (mill US$) 8.69

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 9.90

No of Years 15
Annual Capital Cost (mil US$) 0.66

Fixed Cost

No Items Amount
1 10% of Capital Cost (mil US$/yr) 0.99
Total Fixed Cost (mil US$) 0.99
Variable Cost

No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 0.99


Total Annual Cost with Grid Cost
(mil US$) 1.65

256
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 2.70


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 1.30
Total Revenue (mil US$/y) 4.00

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI 120

Payback Period Analysis

Payback Period 5.3 MW


25.0000

20.0000
Net Cash Flow (mil US$)

15.0000

10.0000

5.0000

0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-5.0000

-10.0000
Year

LCOE Estimation

Annual Cost (mil US$) 1.65


Electricity Generated (kWh/yr) 46251000.00
LCOE (US$/kWh) 0.03

257
Table C-31: Techno-Economic Assessment of the Self-Funded 6.2 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 6.20 9.90 9.86
Total Delivered Cost of Equipment (mill US$) 9.86

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 11.06

No of Years 15
Annual Capital Cost (mil US$) 0.74

Fixed Cost

No Items Amount
1 10% of Capital Cost (mil US$/yr) 1.11
Total Fixed Cost (mil US$) 1.11
Variable Cost

No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 1.11


Total Annual Cost with Grid Cost
(mil US$) 1.84

258
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 3.14


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 1.53
Total Revenue (mil US$/y) 4.67

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI 143

Payback Period Analysis

Payback Period 6.2 MW


35.0000
30.0000
25.0000
Net Cash Flow (mil US$)

20.0000
15.0000
10.0000
5.0000
0.0000
-5.0000 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

-10.0000
-15.0000
Years

LCOE Estimation

Annual Cost (mil US$) 1.84


Electricity Generated (kWh/yr) 64752000.00
LCOE (US$/kWh) 0.03

259
Table C-32: Techno-Economic Assessment of the Self-Funded 7.1 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 7.10 11.03 10.99
Total Delivered Cost of Equipment (mill US$) 10.99

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 12.19

No of Years 15
Annual Capital Cost (mil US$) 0.81

Fixed Cost

No Items Amount
1 10% of Capital Cost (mil US$/yr) 1.22
Total Fixed Cost (mil US$) 1.22
Variable Cost

No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 1.22


Total Annual Cost with Grid Cost
(mil US$) 2.03

260
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 3.59


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 1.75
Total Revenue (mil US$/y) 5.34

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI 163

Payback Period Analysis

Payback Period 7.1 MW


40.0000
35.0000
30.0000
Net Cash Flow (mil US$)

25.0000
20.0000
15.0000
10.0000
5.0000
0.0000
-5.0000 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-10.0000
-15.0000
Year

LCOE Estimation

Annual Cost (mil US$) 2.03


Electricity Generated (kWh/yr) 74002000.00
LCOE (US$/kWh) 0.03

261
Table C-33: Techno-Economic Assessment of the Self-Funded 8.0 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 8.00 12.14 12.09
Total Delivered Cost of Equipment (mill US$) 12.09

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 13.29

No of Years 15
Annual Capital Cost (mil US$) 0.89

Fixed Cost

No Items Amount
1 10% of Capital Cost (mil US$/yr) 1.33
Total Fixed Cost (mil US$) 1.33
Variable Cost

No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 1.33


Total Annual Cost with Grid Cost
(mil US$) 2.21

262
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 4.04


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 1.96
Total Revenue (mil US$/y) 6.00

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI 181

Payback Period Analysis

Payback Period 8.0 MW


50.0000

40.0000
Net Cash Flow (mil US$)

30.0000

20.0000

10.0000

0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-10.0000

-20.0000
Year

LCOE Estimation

Annual Cost (mil US$) 2.21


Electricity Generated (kWh/yr) 83253000.00
LCOE (US$/kWh) 0.03

263
Table C-34: Techno-Economic Assessment of the Self-Funded 8.9 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 8.90 13.22 13.16
Total Delivered Cost of Equipment (mill US$) 13.16

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 14.36

No of Years 15
Annual Capital Cost (mil US$) 0.96

Fixed Cost

No Items Amount
1 10% of Capital Cost (mil US$/yr) 1.44
Total Fixed Cost (mil US$) 1.44
Variable Cost

No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 1.44


Total Annual Cost with Grid Cost
(mil US$) 2.39

264
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 4.49


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 2.18
Total Revenue (mil US$/y) 6.67

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI 197

Payback Period Analysis

Payback Period of 8.9 MW


50.0000

40.0000
Net Cash Flow (mil US$)

30.0000

20.0000

10.0000

0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-10.0000

-20.0000
Years

LCOE Estimation

Annual Cost (mil US$) 2.39


Electricity Generated (kWh/yr) 92503000.00
LCOE (US$/kWh) 0.03

265
Table C-35: Techno-Economic Assessment of the Self-Funded 9.8 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 9.80 14.28 14.22
Total Delivered Cost of Equipment (mill US$) 14.22

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 15.47

No of Years 15
Annual Capital Cost (mil US$) 2.27

Fixed Cost

No Items Amount
1 10% of Capital Cost (mil US$/yr) 1.55
Total Fixed Cost (mil US$) 1.55
Variable Cost

No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 1.55


Total Annual Cost with Grid Cost
(mil US$) 2.58

266
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 4.94


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 2.40
Total Revenue (mil US$/y) 7.34

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI 210

Payback Period Analysis

Payback Period 9.8 MW


60.0000

50.0000
Net Cash Flow (mil US$)

40.0000

30.0000

20.0000

10.0000

0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-10.0000

-20.0000
Years

LCOE Estimation

Annual Cost (mil US$) 2.58


Electricity Generated (kWh/yr) 101753000.00
LCOE (US$/kWh) 0.03

267
Table C-36: Techno-Economic Assessment of the Self-Funded 10.7 MW Power Plant
POMR-RES Capital Cost Estimation

CEPCI Base Year 574.4 2014


CEPCI Design Year 571.9 Aug-17
Scaling Factor 0.8

Base Base Design Design Present


No Equipment
Size Cost Size Cost DOE
1 Per MW Cost 1 2.3 10.70 15.32 15.25
Total Delivered Cost of Equipment (mill US$) 15.25

Additional Cost :
No Items Factor Amount
1 Grid Connection Cost (mil US$) 0.23 mil US$/km 1.15
2 Administration Fees (mil US$) 0.05

Total Capital Cost with Grid Cost


(mil US$) 16.45

No of Years 15
Annual Capital Cost (mil US$) 1.10

Fixed Cost

No Items Amount
1 10% of Capital Cost (mil US$/yr) 1.65
Total Fixed Cost (mil US$) 1.65
Variable Cost

No Items Amount
1 Raw Materials
2 Utilities
3 Transportation of Feedstock
Total Variable Cost (mil US$) 0

Total Operation Cost (mil US$) 1.65


Total Annual Cost with Grid Cost
(mil US$) 2.74

268
POMR-RES Revenue Interpretation
Revenue = FiT Rate Grant + Steam Cost Saving + Electricity
Cost Saving

FiT Rate Grant (mil US$/y) 5.16


Steam Cost Saving (mil US$/y) 0.00
Electricity Cost Saving (mil US$/y) 2.62
Total Revenue (mil US$/y) 7.78

ROI
ROI = ((Net Cash Flow – Total Cash Flow) / Total Cash Flow)/100

ROI 208

Payback Period Analysis

Payback Period 10.7 MW


60.0000

50.0000
Net Cash Flow (mill US$)

40.0000

30.0000

20.0000

10.0000

0.0000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-10.0000

-20.0000
Year

LCOE Estimation

Annual Cost (mil US$) 2.74


Electricity Generated (kWh/yr) 111004000.00
LCOE (US$/kWh) 0.02

269
Table C-37: DCF of the Private-Funded 0.90 MW Power Plant

Discounted Cash Flow (DCF)

Year Simple Cash Flow Net Present Value


(mill RM) (mill RM)
0 -3.30 -3.30
1 -0.76 -3.98
2 -0.76 -4.58
3 -0.76 -5.12
4 -0.76 -5.60
5 -0.76 -6.03
6 -0.76 -6.42
7 -0.76 -6.76
8 -0.76 -7.06
9 -0.76 -7.34
10 -0.76 -7.58
11 -0.76 -7.80
12 -0.76 -7.99
13 -0.76 -8.16
14 -0.76 -8.32
15 -0.76 -8.46

DCF
0.00
0 1 2 3 4 5 6 7 8 9 10
-1.00

-2.00
NPV (mill RM)

-3.00

-4.00

-5.00

-6.00

-7.00

-8.00
Year

270
Table C-38: DCF of the Private-Funded 1.80 MW Power Plant

Discounted Cash Flow (DCF)

Year Simple Cash Flow Net Present Value


(mill RM) (mill RM)

0 -4.86 -4.86
1 -0.55 -5.36
2 -0.55 -5.80
3 -0.55 -6.19
4 -0.55 -6.54
5 -0.55 -6.85
6 -0.55 -7.13
7 -0.55 -7.38
8 -0.55 -7.61
9 -0.55 -7.80
10 -0.55 -7.98
11 -0.55 -8.14
12 -0.55 -8.28
13 -0.55 -8.41
14 -0.55 -8.52
15 -0.55 -8.62

DCF
0.00
0 1 2 3 4 5 6 7 8 9 10
-1.00
-2.00
-3.00
NPV (mill RM)

-4.00
-5.00
-6.00
-7.00
-8.00
-9.00
Year

271
Table C-39: DCF of the Private-Funded 2.70 MW Power Plant

Discounted Cash Flow (DCF)

Year Simple Cash Flow Net Present Value


(mill RM) (mill RM)

0 -6.27 -6.27
1 -0.43 -6.65
2 -0.43 -6.99
3 -0.43 -7.30
4 -0.43 -7.57
5 -0.43 -7.81
6 -0.43 -8.03
7 -0.43 -8.23
8 -0.43 -8.40
9 -0.43 -8.55
10 -0.43 -8.69
11 -0.43 -8.81
12 -0.43 -8.92
13 -0.43 -9.02
14 -0.43 -9.11
15 -0.43 -9.19

DCF
0.00
-1.00 0 1 2 3 4 5 6 7 8 9 10

-2.00
-3.00
NPV (mill RM)

-4.00
-5.00
-6.00
-7.00
-8.00
-9.00
-10.00
Year

272
Table C-40: DCF of the Private-Funded 3.60 MW Power Plant

Discounted Cash Flow (DCF)

Year Simple Cash Flow Net Present Value


(mill RM) (mill RM)

0 -7.58 -7.58
1 -0.27 -7.82
2 -0.27 -8.04
3 -0.27 -8.23
4 -0.27 -8.40
5 -0.27 -8.55
6 -0.27 -8.69
7 -0.27 -8.81
8 -0.27 -8.92
9 -0.27 -9.02
10 -0.27 -9.11
11 -0.27 -9.18
12 -0.27 -9.25
13 -0.27 -9.31
14 -0.27 -9.37
15 -0.27 -9.42

273
Table C-41: DCF of the Private-Funded 4.50 MW Power Plant

Discounted Cash Flow (DCF)

Year Simple Cash Flow Net Present Value


(mill RM) (mill RM)

0 -8.69 -8.69
1 -0.04 -8.72
2 -0.04 -8.75
3 -0.04 -8.78
4 -0.04 -8.80
5 -0.04 -8.82
6 -0.04 -8.84
7 -0.04 -8.86
8 -0.04 -8.87
9 -0.04 -8.89
10 -0.04 -8.90
11 -0.04 -8.91
12 -0.04 -8.92
13 -0.04 -8.93
14 -0.04 -8.93
15 -0.04 -8.94

DCF
-8.55
0 1 2 3 4 5 6 7 8 9 10
-8.60

-8.65
NPV (mill RM)

-8.70

-8.75

-8.80

-8.85

-8.90

-8.95
Year

274
Table C-42: DCF of the Private-Funded 5.40 MW Power Plant

Discounted Cash Flow (DCF)

Year Simple Cash Flow Net Present Value


(mill RM) (mill RM)

0 -9.89 -9.89
1 0.17 -9.75
2 0.17 -9.61
3 0.17 -9.49
4 0.17 -9.39
5 0.17 -9.29
6 0.17 -9.21
7 0.17 -9.13
8 0.17 -9.06
9 0.17 -9.00
10 0.17 -8.95
11 0.17 -8.90
12 0.17 -8.86
13 0.17 -8.82
14 0.17 -8.78
15 0.17 -8.75

DCF
-8.40
0 1 2 3 4 5 6 7 8 9 10
-8.60

-8.80
NPV (mill RM)

-9.00

-9.20

-9.40

-9.60

-9.80

-10.00
Year

275
Table C-43: DCF of the Private-Funded 6.30 MW Power Plant

Discounted Cash Flow (DCF)

Year Simple Cash Flow Net Present Value


(mill RM) (mill RM)

0 -11.06 -11.06
1 0.38 -10.71
2 0.38 -10.41
3 0.38 -10.13
4 0.38 -9.89
5 0.38 -9.67
6 0.38 -9.48
7 0.38 -9.30
8 0.38 -9.15
9 0.38 -9.01
10 0.38 -8.89
11 0.38 -8.78
12 0.38 -8.68
13 0.38 -8.59
14 0.38 -8.51
15 0.38 -8.44

DCF
0.00
0 1 2 3 4 5 6 7 8 9 10
-2.00

-4.00
NPV (mill RM)

-6.00

-8.00

-10.00

-12.00
Year

276
Table C-44: DCF of the Private-Funded 7.20 MW Power Plant

Discounted Cash Flow (DCF)

Year Simple Cash Flow Net Present Value


(mill RM) (mill RM)

0 -12.19 -12.19
1 0.61 -11.64
2 0.61 -11.15
3 0.61 -10.71
4 0.61 -10.32
5 0.61 -9.97
6 0.61 -9.66
7 0.61 -9.38
8 0.61 -9.14
9 0.61 -8.91
10 0.61 -8.72
11 0.61 -8.54
12 0.61 -8.38
13 0.61 -8.24
14 0.61 -8.12
15 0.61 -8.00

DCF
0.00
0 1 2 3 4 5 6 7 8 9 10
-2.00

-4.00
NPV (mill RM)

-6.00

-8.00

-10.00

-12.00

-14.00
Year

277
Table C-45: DCF of the Private-Funded 8.00 MW Power Plant

Discounted Cash Flow (DCF)

Year Simple Cash Flow Net Present Value


(mill RM) (mill RM)

0 -13.29 -13.29
1 0.85 -12.52
2 0.85 -11.84
3 0.85 -11.23
4 0.85 -10.69
5 0.85 -10.21
6 0.85 -9.77
7 0.85 -9.39
8 0.85 -9.04
9 0.85 -8.73
10 0.85 -8.46
11 0.85 -8.21
12 0.85 -7.99
13 0.85 -7.80
14 0.85 -7.62
15 0.85 -7.47

DCF
0.00
0 1 2 3 4 5 6 7 8 9 10
-2.00

-4.00
NPV (mill RM)

-6.00

-8.00

-10.00

-12.00

-14.00
Year

278
Table C-46: DCF of the Private-Funded 8.90 MW Power Plant

Discounted Cash Flow (DCF)

Year Simple Cash Flow Net Present Value


(mill RM) (mill RM)

0 -14.36 -14.36
1 1.10 -13.38
2 1.10 -12.50
3 1.10 -11.71
4 1.10 -11.01
5 1.10 -10.38
6 1.10 -9.82
7 1.10 -9.32
8 1.10 -8.87
9 1.10 -8.48
10 1.10 -8.12
11 1.10 -7.80
12 1.10 -7.52
13 1.10 -7.27
14 1.10 -7.04
15 1.10 -6.84

DCF
0.00
0 1 2 3 4 5 6 7 8 9 10
-2.00

-4.00
NPV (mill RM)

-6.00

-8.00

-10.00

-12.00

-14.00

-16.00
Year

279
Table C-47: DCF of the Private-Funded 9.80 MW Power Plant

Discounted Cash Flow (DCF)

Year Simple Cash Flow Net Present Value


(mill RM) (mill RM)

0 -15.42 -15.42
1 1.36 -14.20
2 1.36 -13.11
3 1.36 -12.14
4 1.36 -11.28
5 1.36 -10.50
6 1.36 -9.81
7 1.36 -9.20
8 1.36 -8.65
9 1.36 -8.15
10 1.36 -7.71
11 1.36 -7.32
12 1.36 -6.97
13 1.36 -6.66
14 1.36 -6.38
15 1.36 -6.13

DCF
0.00
0 1 2 3 4 5 6 7 8 9 10
-2.00
-4.00
-6.00
NPV (mill RM)

-8.00
-10.00
-12.00
-14.00
-16.00
-18.00
Year

280
Table C-48: DCF of the Private-Funded 10.70 MW Power Plant

Discounted Cash Flow (DCF)

Year Simple Cash Flow Net Present Value


(mill RM) (mill RM)

0 -16.45 -16.45
1 1.40 -15.20
2 1.40 -14.09
3 1.40 -13.09
4 1.40 -12.20
5 1.40 -11.41
6 1.40 -10.70
7 1.40 -10.07
8 1.40 -9.50
9 1.40 -9.00
10 1.40 -8.55
11 1.40 -8.15
12 1.40 -7.79
13 1.40 -7.47
14 1.40 -7.18
15 1.40 -6.93

DCF
0.00
0 1 2 3 4 5 6 7 8 9 10
-2.00
-4.00
-6.00
NPV (mill RM)

-8.00
-10.00
-12.00
-14.00
-16.00
-18.00
Year

281
Appendix D

This appendix presents the life cycle inventories (LCI) used to conduct the cradle-to-grave
LCA for POMR-RES. As mentioned in Section 6.3, the LCI are gathered from the Eco-invent
v 3 database (Wernet et al., 2016). The allocation default dataset are used following the
attributional LCA modelling that aim to assess the direct EIs from the life cycle (Brander et
al., 2008). Furthermore, unit process inventories are used to describe a single process step to
provide transparency (Goedkoop et al., 2016). Appendix D-1 and Appendix D-2 define the
inventories for oil palm plantation and CPO milling processes relevant to Malaysia,
respectively. The inventories for the combustion process for wood waste in the industrial boiler
based on surrogate date from the United State of America are summarised in Appendix D-3.
Lastly, the inventories for the 5.7 MW power plant (RoW) are described in Appendix D-4.

Appendix D-1 – Oil Palm Plantation Stage

The dataset presented in Table D-1 are the inventories used for the oil palm plantation stage to
produce 1 kg of FFB (Palm fruit bunch {MY} |production|Alloc Def,U). The plantation lifetime
includes establishment, operation, clearing the plantation and transporting the harvested FFB
to the farm processes. The plantation modelled in the base case are considered matured
(established for more than 20 years) and the database was adapted to represent a without LUC
effect (a scenario with LUC was also modelled). The input for the process includes mineral
fertilisers and pesticides while the input from the oil palm nursery is consider negligible as
quoted by Muhamad et al. (2014). The distance to transport the harvested FFB from the
plantation area to the mills is kept at 25 km.

Table D-1: Life Cycle Inventories for Oil Palm Plantation to Produce 1 kg of FFB
Inventories Value Unit
Input from Nature
Carbon dioxide, in air 1.15 kg
Transformation, from permanent crop 0.00 m2
Transformation, to permanent crop 0.00 m2
Occupation, permanent crop 0.40 m2

Materials/Fuels
Ammonium Sulphate 0.006 kg
Phenoxy-Compound 5.840E-6 kg
Transport, tractor and trailer 0.026 tkm
Pesticide 1.142E-7 kg
282
Organophosphorus-compound 4.041E-5 kg
Phosphate fertiliser, as P205 0.001 kg
Pyrethroid-compound 2.882E-6 kg
Dolomite 0.003 kg
Potassium chloride, as K20 0.009 kg
Carbamate-compound 2.041E-5 kg
Lime 0.002 kg
Land Use Change, perennial crop 0.000 ha
Wood chipping, chipper, mobile, diesel 1.390 hr
Benzimidazole-compound 3.723E-8 kg
Output
Emission to air
Ammonia 6.126E-4 kg
Nitrogen oxides 3.844E-5 kg
Dinitrogen monoxides 1.830E-4 kg
Water 0.006 m3
Carbon Dioxide, land transformation 0.000 kg
Emission to water
Nitrate, groundwater 0.023 kg
Water, river 0.002 m3
Phosphorus, river 7.788E-6 kg
Phosphorus, groundwater 2.802E-6 kg
Water, groundwater 0.009 kg
Emission to Soil
Benomyl 3.723E-8 kg
Copper -3.833E-6 kg
Nickel 7.692E-6 kg
Lead 7.480E-6 kg
Zinc -4.217E-6 kg
Thiram 6.605E-8 kg
Chromium 5.845E-8 kg
Cadmium 1.447E-9 kg
Cypermethrin 2.882E-6 kg
Glyphosate 4.047E-5 kg
Carbofuran 2.041E-5 kg

Appendix D-2 – PO Milling Process Stage

Table D-2 provides the inventories for PO milling process in Malaysia in order to produce 1
kg PO and EFB as the co-products (Palm Oil, crude {MY}|palm oil mill operation|Alloc
Def,U). The activities in the CPO wet milling process includes the sterilization, stripping, oil
extraction (digestion and pressing) and oil clarification as presented in Figure 3.1 and Figure
6.3. The heat and energy supply for the mills is considered to be extracted from PKS and fibres
and treatment of POME and other wastewater are taken into account.

283
Table D-2: Life Cycle Inventories for Wet Milling Process (to Produce 1 kg of PO)
Inventories Value Unit
Inputs from Nature
Carbon dioxide, in air 0.00 kg

Materials/Fuels
Chemical 8.133E-5 kg
Phosphoric acid 7.625E-4 kg
Sodium chloride 5.809 E-5 kg
Heat and Power co-generation unit 1.367 E-8 p
Ammonia 1.161 E-7 kg
Hexane 0.003 kg
Palm fruit bunch 3.940 kg
Lubricating Oil 4.647 E-5 kg
Chlorines 4.647 E-6 kg
Water 0.011 kg
Tap water 12.066 kg
Oil Mill 8.684E-10 p

Output
Emission to air
Phenol 1.108 E-10 kg
Benzene 3.844E-5 kg
Potassium 1.830E-4 kg
Bromine 8.501 E-7 m3
Mercury 4.250 E-9 kg
Water 0.002
Ammonia 2.465 E-5 kg
Dioxin 4.241E-13 kg
Formaldehyde 1.778 E-6 kg
Manganese 2.422 E-6 kg
Nickel 8.501 E-8 kg
Zinc 4.250 E-6
Sulphur Dioxide 3.528 E-5 kg
Particulates 0.001 kg
Chromium 5.667E-10 kg
Sodium 1.842 E-5 kg
Acetaldehyde 8.642 E-7 kg
Carbon Monoxide 9.577E-5 kg
m-Xylene 1.641 E-6 kg
Toluene 4.104 E-6 kg
Nitrogen Oxides 0.001 kg
Hexane 0.002 kg
Methane 5.937 E-6 kg
Fluorine 7.084 E-7 kg
Hydrocarbons 4.241 E-5 kg
Dinitrogen Monoxide 3.258 E-5 kg
Hydrocarbons 1.245 E-5 kg
Calcium 8.289 E-5 kg
Cadmium 9.918 E-9 kg
NMVOC 8.345 E-6 kg
284
Polycyclic Aromatic Hydrocarbons 1.504 E-7 kg
Chromium 5.610E-8 kg
Benzopyrene 6.840 E-9 kg
Lead 3.258 E-7 kg
Carbon Dioxide 1.825 kg
Phosphorus 4.250 E-6 kg
Magnesium 5.115 E-6 kg
Chlorine 2.550 E-6 kg
Arsenic 1.417E-8 kg
Copper 3.117 E-7 kg
Emission to water
Water 0.006 m3
Waste to Treatment
Waste mineral oil 4.647E-5 kg
Municipal Solid Waste 0 kg
Wood ash mixture 0.007 kg
Wastewater 1.115 E-5 m3

Appendix D-3 – Combustion Process Stage

The data set for wood waste combustion process with the average moisture content of nearly
50% (however varies from 5% to 75% depending from residues types) in the industrial boiler
from the United States of America (Wood waste, unspecified, combusted in industrial
boiler/US) to produce heat energy are tabulated in Table D-3. No specification of types of wood
waste is given in the inventory description, however, the emission profile is found to be
appropriate to represent the EFB combustion process. The combustion process is assumed to
take place in the industrial boiler that provide nearly complete combustion of organic matter of
the residues (EPA, 2003).

Table D-3: Life Cycle Inventories for Combustion of 1 kg EFB

Inventories Value Unit


Materials/Fuels
EFB 1.44 kg

Emission to air
Acetaldehyde 7.47E-6 kg
Acrolein 36E-6 kg
Antimony 7E-8 kg
Arsenic 19.8E-8 kg
Benzene 0.000037 kg
Beryllium 4.08E-9 kg
Cadmium 3.6E-8 kg
Carbon Monoxide 0.0054 kg
Methane 40.5E-8 kg
285
Chlorine 7.11E-6 kg
Chromium 18.9E-8 kg
Cobalt 5.8E-8 kg
Dioxin 1.53E-8 kg
Formaldehyde 39.6E-6 kg
Hydrogen Chloride 0.000171 kg
Lead 43.8E-8 kg
Manganese 14.4E-6 kg
Mercury 30.2E-8 kg
Metals, Unspecified 0.000384 kg
Methane 0.000189 kg
Naphthalene 87.9E-8 kg
Nickel 29.4E-6 kg
Nitrogen Oxides 29.43E-8 kg
Particulates >2.5mm<10mm 0.0045 kg
Phenols 45.6E-8 kg
Selenium 2.5E-8 kg
Sulfur Oxides 0.000225 kg
TOC 36.81E-8 kg

Emission to Water
BOD 5 0.0175 kg

Appendix D-4 – EFB- based Power Plant

The dataset used for construction of gas power plant (Gas power plant, 100 MW electrical
{RoW}|construction|Alloc Def, U) includes the material used and the energy requirements are
presented in Table D-4. The data presented have been scaled down to 5.7 MW from 100 MW
using power law (Eq. 14) with 0.8 scaling factor.

Table D-2: Life Cycle Inventories for 5.7 MW POMR-RES


Inventories Value Unit
Resources
Transformation 2021 m2
Occupation 72783 m2a
Transformation, to industrial area 2021 m2

Materials/Fuels
Aluminium, cast alloy 2426 kg
Concrete, normal (RoW) 183 m3
Polyethylene, low density (GLO) 15163 kg
Aluminium, wrought alloy (GLO) 5155 kg
Stone wool, packed (GLO 15163 kg
Reinforcing steel (GLO) 55598 kg
Copper (GLO) 7581 kg

286
Electricity/heat
Diesel, burning in building machine 1172622 MJ
(GLO) kWh
Electricity medium voltage (AU) 305 kWh
Electricity medium voltage (RU) 1299 kWh
Electricity medium voltage (TR) 293 kWh
Electricity medium voltage (RAF) 330 kWh
Electricity medium voltage (RAS) 11074 kWh
Electricity medium voltage (RLA) 1221 kWh
Electricity medium voltage (RNA) 5974 kWh
Electricity medium voltage (RoW) 3355 kWh
Heat, district, other than natural gas 477 MJ
Heat, district, other than natural gas 1113513 MJ

287
Appendix E – Publication

This appendix compiled the publications related to the research as listed below:

1. Md Jaye, I. F., Sadhukhan, J. and Murphy, R. J. (2016) ‘Renewable, local electricity


generation from palm oil mills: a case study from Peninsular Malaysia’, International
Journal of Smart Grid and Clean Energy, 44, pp. 106–111. doi: 10.12720/sgce.5.2.106-
111.
2. Md Jaye, I. F., Sadhukhan, J. and Murphy, R. J. (2018) ‘Integrated Assessment of Palm
Oil Mill Residues to Sustainable Electricity System ( POMR-SES ): A Case Study from
Peninsular Malaysia’, IOP Conference Series: Materials Sceince and Engineering, 358,
pp. 1–7. doi: 10.1088/1757-899X/358/1/012002.
3. Sadhukhan, J. et al. (2018) ‘Role of bioenergy, biorefinery and bioeconomy in
sustainable development:Strategic pathways for Malaysia’, Renewable and Sustainable
Energy Reviews, 81, pp. 1966–1987. doi: http://dx.doi.org/10.1016/j.rser.2017.06.007.

288

Potrebbero piacerti anche