153
VALUE CHAIN OPTIMIZATION OF A FOREST BIOMASS POWER PLANT CONSIDERING UNCERTAINTIES by Nazanin Shabani B.Sc., Civil Engineering, Sharif University of Technology, 2003 M.Sc., Civil Engineering, Iran University of Science and Technology, 2005 M.A.Sc., Civil Engineering, University of British Columbia, 2009 A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY in The Faculty of Graduate and Postdoctoral Studies (Forestry) THE UNIVERSITY OF BRITISH COLUMBIA (Vancouver) April 2014 © Nazanin Shabani, 2014

2014 Shabani VALUE CHAIN OPTIMIZATION OF A FOREST BIOMASS POWER PLANT CONSIDERING UNCERTAINTIES

Embed Size (px)

DESCRIPTION

VALUE CHAIN OPTIMIZATION OF A FOREST BIOMASS POWER PLANT CONSIDERING UNCERTAINTIES

Citation preview

VALUE CHAIN OPTIMIZATION OF A FOREST BIOMASS POWER PLANT

CONSIDERING UNCERTAINTIES

by

Nazanin Shabani

B.Sc., Civil Engineering, Sharif University of Technology, 2003

M.Sc., Civil Engineering, Iran University of Science and Technology, 2005

M.A.Sc., Civil Engineering, University of British Columbia, 2009

A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF

THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

in

The Faculty of Graduate and Postdoctoral Studies

(Forestry)

THE UNIVERSITY OF BRITISH COLUMBIA

(Vancouver)

April 2014

© Nazanin Shabani, 2014

ii

Abstract

Mathematical modeling has been employed to improve the cost competitiveness of forest

bioenergy supply chains. Most of the studies done in this area are at the strategic level, focus on

one part of the supply chain and ignore uncertainties. The objective of this thesis is to optimize

the value generated in a forest biomass power plant at the tactical level considering uncertainties.

To achieve this, first the supply chain configuration of a power plant is presented and a nonlinear

model is developed and solved to maximize its overall value. The model considers procurement,

storage, production and ash management in an integrated framework and is applied to a real case

study in Canada. The optimum solution forecasts $1.74M lower procurement cost compared to

the actual cost of the power plant. Sensitivity analysis and Monte Carlo simulation are performed

to identify important uncertain parameters and evaluate their impacts on the solution.

The model is reformulated into a linear programming model to facilitate incorporating

uncertainty in the decision making process. To include uncertainty in the biomass availability,

biomass quality and both of them simultaneously, a two-stage stochastic programming model, a

robust optimization model and a hybrid stochastic programming-robust optimization model are

developed, respectively. The results show that including uncertainty in the optimization model

provides a solution which is feasible for all realization of uncertain parameters within the defined

scenario sets or uncertainty ranges, with a lower profit compared to the deterministic model.

Including uncertainty in biomass availability using the stochastic model decreases the profit by

$0.2M. In the robust optimization model, there is a trade-off between the profit and the selected

range of biomass quality. Profit decreases by up to $3.67M when there are ±13% variation in

moisture content and ±5% change in higher heating value. The hybrid model takes advantage of

iii

both modeling approaches and balances the profit and model tractability. With the cost of only

$30,000, an implementable solution is provided by the hybrid model with unique first stage

decision variables. These models could help managers of a biomass power plant to achieve

higher profit by better managing their supply chains.

iv

Preface

This dissertation is original and presents the work of Nazanin Shabani during her Ph.D. program.

The research was conducted by the author under the supervision of her academic adviser, Dr.

Taraneh Sowlati. Dr. Sowlati advised Shabani during the process of defining the research topic,

gathering data and information, developing and validating the models and preparing manuscripts.

This thesis presents a background on the research topic, research objectives, a review of the

literature, several optimization models to achieve the research objectives with their application to

a real case study in Canada, analysis of the obtained results, and the main findings and

conclusions. The author visited the power plant several times, had close collaboration with the

managers of the power plant, obtained information and detailed data on the power plant supply

chain, presented the model results to the power plant managers and had the model validated by

them. Five scientific papers were generated from this research, and in all of them Shabani was

the first author. The list of papers generated from this research is provided below.

A version of Chapter 2 is published. Shabani, N., Akhtari, S., Sowlati, T. 2013. Value chain

optimization of forest biomass for bioenergy production: A review. Renewable and

Sustainable Energy Reviews, 110(3): 280-290.

A version of Chapter 3 is published. Shabani N, Sowlati T. 2013. A mixed integer non-linear

programming model for tactical value chain optimization of a wood biomass power plant.

Applied Energy, 104:353-361.

A version of Chapter 4 is submitted for publication. Shabani N, Sowlati T.Evaluating the

impact of uncertainty and variability on the value chain optimization of a forest biomass

power plant using Monte Carlo Simulation.

v

A version of Chapter 5 is submitted for publication. Shabani N, Sowlati T., Ouhimmou M.,

Rönnqvist M. Tactical supply chain planning for a forest biomass power plant under supply

uncertainty.

A version of Chapter 6 is submitted for publication. Shabani N, Sowlati T. A hybrid

stochastic programming-robust optimization model for maximizing the value chain of a

forest biomass power plant under uncertainty.

vi

Table of Contents

Abstract ........................................................................................................................................... ii

Preface............................................................................................................................................ iv

Table of Contents ........................................................................................................................... vi

List of Tables ................................................................................................................................. ix

List of Figures ................................................................................................................................ xi

Acknowledgements ...................................................................................................................... xiii

Dedication ..................................................................................................................................... xv

Chapter 1 Introduction .................................................................................................................. 1

1.1 Background ...................................................................................................................... 1

1.2 Research objectives .......................................................................................................... 6

1.3 Organization of the dissertation ....................................................................................... 7

Chapter 2 Literature review ........................................................................................................... 9

2.1 Synopsis ........................................................................................................................... 9

2.2 Deterministic optimization models .................................................................................. 9

2.2.1 Power plants ............................................................................................................ 10

2.2.2 District heating plants ............................................................................................. 13

2.2.3 Co-generation plants ............................................................................................... 16

2.2.4 Biofuel plants .......................................................................................................... 20

2.3 Optimization models with uncertainties ......................................................................... 23

2.3.1 Modeling approaches .............................................................................................. 27

2.3.2 Sensitivity analysis and Monte Carlo simulation.................................................... 31

2.3.3 Stochastic programming ......................................................................................... 34

2.3.4 Robust optimization model ..................................................................................... 38

vii

2.4 Discussion and conclusions ............................................................................................ 41

Chapter 3 Deterministic model ................................................................................................... 43

3.1 Synopsis ......................................................................................................................... 43

3.2 The power plant supply chain ........................................................................................ 43

3.3 The optimization model ................................................................................................. 49

3.4 Case study ...................................................................................................................... 55

3.5 Results ............................................................................................................................ 60

3.5.1 Scenario analysis ..................................................................................................... 61

3.5.2 Sensitivity analysis.................................................................................................. 63

3.6 Discussion and conclusions ............................................................................................ 65

Chapter 4 Monte Carlo simulation .............................................................................................. 68

4.1 Synopsis ......................................................................................................................... 68

4.2 Uncertainty and Monte Carlo simulation ....................................................................... 68

4.2.1 Uncertainty in biomass quality ............................................................................... 70

4.2.2 Uncertainty in biomass availability and cost and electricity price ......................... 76

4.3 Results ............................................................................................................................ 77

4.4 Discussion and conclusions ............................................................................................ 82

Chapter 5 Stochastic programming ............................................................................................. 85

5.1 Synopsis ......................................................................................................................... 85

5.2 The mixed integer programming model of the power plant supply chain ..................... 85

5.3 The stochastic mixed integer programming model of the power plant supply chain .... 88

5.4 Managing the risk ........................................................................................................... 92

5.4.1 Variability index ..................................................................................................... 93

5.4.2 Downside risk ......................................................................................................... 94

5.5 Results ............................................................................................................................ 95

viii

5.5.1 Result of deterministic models................................................................................ 95

5.5.2 Results of the stochastic model ............................................................................... 95

5.5.3 Results for the variability index ............................................................................ 101

5.5.4 Results for the downside risk ................................................................................ 103

5.6 Discussion and conclusions .......................................................................................... 103

Chapter 6 Hybrid stochastic programming-robust optimization model .................................... 106

6.1 Synopsis ....................................................................................................................... 106

6.2 Robust optimization formulation ................................................................................. 106

6.3 Hybrid stochastic programming-robust optimization model ....................................... 111

6.4 Results .......................................................................................................................... 113

6.5 Discussion and conclusions .......................................................................................... 118

Chapter 7 Conclusions, strength points, limitations and future research .................................. 121

7.1 Conclusions .................................................................................................................. 121

7.2 Strengths points ............................................................................................................ 122

7.3 Limitations ................................................................................................................... 124

7.4 Future research ............................................................................................................. 125

References ................................................................................................................................... 127

ix

List of Tables

Table 2-1: Summary of studies on deterministic optimization of forest biomass power plants ... 13

Table 2-2: Summary of studies on deterministic optimization of district heating plants ............. 15

Table 2-3: Summary of studies on deterministic optimization of co-generation plants ............... 19

Table 2-4: Summary of studies on deterministic optimization of biofuel plants .......................... 23

Table 2-5: Summary of studies on sensitivity analysis, scenario analysis and Monte Carlo

simulation applied to bioenergy supply chain with uncertainty ................................................... 33

Table 2-6: Summary of studies on stochastic programming of forest and bioenergy supply chains

....................................................................................................................................................... 38

Table 2-7: Summary of studies on robust optimization of forest and bioenergy supply chain .... 41

Table 3-1: List of indices and decision variables used in the optimization model ....................... 49

Table 3-2: List of parameters used in the optimization model ..................................................... 50

Table 3-3: Characteristics of the case study ................................................................................. 56

Table 3-4: Variables and parameters of the case study ................................................................. 58

Table 3-5: Results of cost, revenue and profit for the optimization model (in $M) ..................... 60

Table 3-6: Total profit and biomass procurement cost for four different scenarios ..................... 63

Table 4-1: Product type of suppliers and their contract type ........................................................ 70

Table 4-2: Average and standard deviation of bark, sawdust and shavings MC for Suppliers 1 to

5..................................................................................................................................................... 71

Table 4-3: Average and standard deviation of biomass MC for Supplier 6 ................................. 73

Table 4-4: Average and standard deviation of HHV for different biomass types ........................ 75

Table 4-5: Minimum, maximum, average and standard deviation of profit for considering

uncertainty in different parameters ............................................................................................... 78

Table 4-6: Results of Monte Carlo simulation-optimization model for scenarios of electricity

price and biomass availability and cost ........................................................................................ 79

Table 4-7: Probability of having profit within different ranges when considering uncertainty in

different model parameters ........................................................................................................... 81

x

Table 4-8: Ranges of biomass purchase from suppliers without contract, biomass consumption

and storage levels when considering uncertainty in different parameters (1000 green tonnes) ... 82

Table 5-1: Decision variables of the linear programming model ................................................. 85

Table 5-2: Stochastic model decision variables ............................................................................ 90

Table 5-3: Expected value of profit for scenario analysis, stochastic and average scenario models

($M) .............................................................................................................................................. 96

Table 5-4: Biomass procurement cost for each scenario of stochastic and deterministic models

($M) .............................................................................................................................................. 99

Table 5-5: Average profit if the first stage decision variables of each scenario is implemented

and other scenarios happen ($M) ................................................................................................ 100

Table 5-6: Average and standard deviation of the monthly biomass consumption for

deterministic model with scenario analysis and stochastic models (1000 green tonnes) ........... 101

Table 6-1: Profit ($M) for different ranges of MCs,p,t and HHVs,p,t used in the robust optimization

model........................................................................................................................................... 114

Table 6-2: Profit for different ranges of MCs,p,t and HHVs,p,t used in the robust optimization and

hybrid models.............................................................................................................................. 118

xi

List of Figures

Figure 3-1: Schematic of supply chain configuration of a forest biomass power plant ............... 43

Figure 3-2: The amount of firm and surplus electricity production in each month ...................... 61

Figure 3-3: Optimum amount of biomass stored, purchased and consumed in each month based

on the 2011 data ............................................................................................................................ 61

Figure 3-4: Variations in profit with 20% change in different parameters ................................ 64

Figure 3-5: Variations in profit for different initial storage levels ............................................... 64

Figure 4-1: Histogram and probability distribution of MC of bark (a), sawdust (b), shavings (c)72

Figure 4-2: Histogram and probability distribution for MC of RLD in January (a), February (b),

March (c), April (d), June (e), July (f), August (g), September (h), October (i), November (j), and

December (k) ................................................................................................................................ 75

Figure 4-3: Histogram and probability distribution of HHV of sawdust (a) and RLD (b) ........... 76

Figure 4-4: Histogram of profit when MC varies ......................................................................... 78

Figure 4-5: Histogram of profit when HHV varies ....................................................................... 78

Figure 4-6: Histogram of profit when electricity price and biomass availability and cost vary for

a) low, b) average and c) high scenarios ....................................................................................... 80

Figure 5-1: Histogram of profit distribution for the deterministic model with first stage decisions

based on average scenario and the stochastic model .................................................................... 97

Figure 5-2: Profit mean and standard deviation for different weights (ρ) .................................. 102

Figure 5-3: Histogram of profit distribution for different weights (ρ) associated with variability

index ............................................................................................................................................ 102

Figure 5-4: Histogram of profit distribution for before and after managing the downside risk

(Ω=$14M) ................................................................................................................................... 103

Figure 6-1: Solution of the robust optimization model for different ranges of moisture content114

Figure 6-2: Solution of the robust optimization model for different ranges of higher heating value

..................................................................................................................................................... 115

Figure 6-3: Solution of the robust optimization model for different ranges of energy value ..... 115

Figure 6-4: The optimum storage level in different months from the robust optimization model

with HHVs,p,t∈[8100,8900] and MCs,p,t∈[26,34] and the deterministic model ........................... 116

xii

Figure 6-5: The optimum biomass consumption level in different months from the robust

optimization model with HHVs,p,t∈[8100,8900] and MCs,p,t∈[26,34] and the deterministic model

..................................................................................................................................................... 117

Figure 6-6: Optimum storage level of 27 scenarios for the first five months for a) robust

optimization model, and b) hybrid stochastic programming-robust optimization model when

HHVs,p,t∈[8100,8900] and MCs,p,t∈[26,34] ................................................................................ 118

xiii

Acknowledgements

First of all, I offer my enduring gratitude to my dear supervisor, Dr. Taraneh Sowlati, for her

help, guidance, and consideration during my Ph.D. studies. She put endless effort into

supervising and motivating me, and never seemed to be tired or reluctant to help. Her enthusiasm

and dedication for research always inspire me.

I am indeed grateful to my supervisory committee, Dr. Farrokh Sassani and Dr. Philip D. Evans

for their time and invaluable feedback on my research during my PhD studies. Special thanks to

Dr. John D. Nelson, Dr. Harish Krishnan and Dr. Chander K. Shahi for critically reviewing the

thesis and providing constructive comments on it. I am also grateful to Dr. Mustapha Ouhimmou

and Dr. Mikael Rönnqvist for their comments and collaboration on reformulating the non-linear

model to a linear model presented in Chapter 5.

I would also like to thank the Natural Sciences and Engineering Research Council of Canada

(NSERC) for providing the funding for this research. I also acknowledge the partial funding

provided by the power plant and sincerely thank the Fiber Supply Manager and Finance Manager

for their time and support in providing the required data and information for the modeling and

validating the results.

I express my appreciation to my lab-mates and friends in the Industrial Engineering Research

Group for their comments and discussion during my Ph.D. studies. I am also thankful of my

friends at UBC for their kindness and support during my stay in Canada and creating so much

fun in my life. Special thanks to my dear friend, Nina, who assisted me in proofreading the

thesis.

xiv

I would like to express my deepest gratitude to my beloved family in Iran. My sincere

appreciation is due to my lovely parents for their unconditional love and support throughout my

life.

Last but not least, I express my profound affection and appreciation to my love and my best

friend, Hamid, for his continuous love, kindness, encouragement and support during my years of

education. This accomplishment would not have been done without him.

xv

Dedication

To my mother, Faranak, who taught me how to be happy by immersing myself in every

joyful moment of my life.

To my father, Mohammad, my polestar, who cultivated in me a hunger for learning.

1

Chapter 1 Introduction

1.1 Background

Using forest biomass as an energy source is not a new idea; before the twentieth century, wood

was one of the main sources of energy, but it has been substituted partly by coal, oil and natural

gas during the past century (Bowyer et al., 2007). Recently, there has been an interest in the

utilization of more sustainable and secure energy sources such as forest biomass due to the high

levels of emissions associated with fossil fuels, increasing energy demands and volatility of

international energy market (Ragauskas et al., 2006).

Biomass is defined as biological material derived from living, or recently living organisms.

Depending on the source of biomass, there are three main biomass categories: forest (woody)

biomass, agricultural biomass, and bio-wastes (animal waste and municipal solid waste)

(Rentizelas et al., 2009b). This research focuses on utilizing forest biomass to generate

electricity in power plants using direct combustion.

Although the conversion and transportation of forest biomass for energy generation affects the

air quality negatively (Hall & Scrase, 1998), there is a potential to decrease emissions

significantly when it is used as a substitute for fossil fuels (Hall (2002), Ahtikoski et al. (2008),

International Energy Agency (2012)). Generally, if managed (produced, transported and used)

in a sustainable manner, converting biomass to energy or fuel is considered to be a carbon

neutral process, meaning that the amount of CO2 released during biomass combustion is equal

to the amount of CO2 taken from the atmosphere by the plant during the growing stage (Saidur

et al., 2011). Increasing the contribution of biomass in energy generation is an important step in

developing sustainable communities, managing greenhouse gas emissions effectively according

2

to Ragauskas et al. (2006), and decreasing the gap between the actual emission levels and the

international protocol targets such as those in the Kyoto and Copenhagen Accords (Bradley D.,

2010). In areas covered with large forest lands and large forest industries, such as Canada, a

large amount of forest biomass is available for energy generation (Bradley D., 2010). In

Canada, after hydro, biomass has the highest share in the production of renewable energy

(2.9% in 2008 according to Ralevic (2008)).

Forest biomass is a flexible energy source, capable of generating electricity, heat, biofuels or a

combination of them. Compared to other renewable energy sources such as wind or solar, the

advantage of using forest biomass for energy generation is that it can be stored and used on

demand (Hall & Scrase (1998), Demirbaş (2001)). It also has the potential to: increase the

recovery of forest residues that would otherwise be disposed of in landfills or incinerated,

create jobs, and provide local and sustainable energy for communities which in turn will

decrease their dependency on the international fuel market (BIOCAP Canada, 2008).

Forest biomass for the purpose of energy generation can be supplied from forest residues

including branches and tops left in the harvest areas, by-products of other forest product mills,

such as sawdust, bark and shavings (Demirbaş, 2001), and fast growing crops such as poplar

and willow grown specifically for energy purposes (Rockwood et al., 2004). Chipping,

handling, transporting, storing and pre-processing operations, such as drying for improving the

quality of biomass, are usually needed before using forest biomass for energy generation

(Flisberg et al., 2012). Forest biomass can be used for energy generation either directly, such as

in direct heat and power generation, or indirectly such as in biofuels (pellet, bioethanol)

production. The energy production process depends on the conversion technology used in the

3

energy plant (Rentizelas et al., 2009b). The conversion technologies include pelletization,

combustion, co-combustion, gasification, pyrolysis, digestion and fermentation (Demirbaş

(2001), Frombo et al. (2009a)). Different types of products are then sent to customers through

the grid, networks or channels of distributors, wholesalers and retailers. Profit in each stage of

the forest bioenergy supply chain is a function of procurement, transportation, operating,

capital and other costs, and depends also on the availability and quality of biomass (Gunn,

2009).

Despite the advantages of using forest biomass for energy generation, there are several barriers

to its efficient utilization, including biomass availability, cost and quality, conversion

efficiency, transportation cost, and the efficiency of the supply logistics system. Forest biomass

is a bulky material with relatively low density (400-900 kg/m3 (Demirbaş, 2001)) and high

moisture content (Hall, 2002). The quality of raw material plays an important role in the

performance of the production process (Rentizelas et al., 2009b). It is difficult to collect,

transport, handle and store low-density materials. Moreover, unlike fossil fuels, forest biomass

is usually spread over large areas rather than being concentrated. Transportation could

contribute as much as 50% to the total biomass cost in some cases (Allen et al., 1998) and may

involve the use of a large amount of equipment and different transportation modes (Hall &

Scrase, 1998). Inaccessibility of forests in some months during the year, when energy demand

is quite high, raises concerns about the secure supply of biomass to energy plants. Therefore,

storage, which affects the quality of material (Fuller, 1985), is also important in this supply

chain. Comminution and storage of residues can take place either in the forest, at the plant or at

an intermediate point. Another challenge in using forest biomass for energy generation is the

existence of uncertainty in this supply chain due to several factors, such as market instability,

4

natural disasters, policy and climate change as well as the nature of the industry (for example

heterogeneous raw material and unpredictable quality (Hall, 2002)). Uncertainty makes this

supply chain volatile and risk vulnerable, which in turn makes proper planning difficult.

All these challenges contribute to a higher cost of energy generation from forest biomass

compared to that of other sources of energy. Utilizing more advanced technologies, for

example to improve raw material quality or system efficiency, is one way to deal with some of

these challenges. A complementary approach to reduce the cost of energy produced from forest

biomass and increase its competitiveness is to improve its supply chain and optimize its design

and production planning (Bowyer et al., 2007).

A supply chain model can be developed to help decision makers in their decisions and manage

the supply chain more efficiently. Operations research and mathematical programming have

been used in forest biomass supply chain planning and management. Modeling is effective

particularly if it integrates different parts of the supply chain, such as procurement, production,

transportation, and distribution, and at different decision levels, such as strategic, tactical and

operational levels. Using optimization techniques in designing and managing forest bioenergy

supply chains could result in better performance which could help to make this energy source

economically viable according to Bowyer et al. (2007).

Optimization models have been developed and used in the literature to determine the optimal

material flow, transportation, storage and chipping location of energy systems, mainly heating

plants (Eriksson & Björheden (1989), Gunnarsson et al. (2004), Kanzian et al. (2009), Freppaz

et al. (2004) and Van Dyken et al. (2010)). There are also some studies that evaluated the

conversion technology and the possibility of co-generation in the design of district heating

5

systems using mathematical programming (Nagel (2000), Frombo et al. (2009a), Difs et al.

(2010), Wetterlund & Söderström (2010), Börjesson & Ahlgren (2010) and Keirstead et al.

(2012)). Biomass supply chains for generating biofuels have also been studied (Chinese &

Meneghetti (2009), Ekşioğlu et al. (2009), Kim et al. (2011a), and a iba e -Aguilar et al.

(2011)). Although using forest biomass for electricity generation is not as common as using it

for generating heat, there are some studies that considered the supply chain design of biomass

power plants in an optimization framework, i.e. Reche et al. (2008), Alam et al. (2009 and

2012a) and Vera et al. (2010). These studies focused on the strategic design of a forest

biomass power plant and did not consider the tactical planning with multiple time steps in their

models. Alam et al. (2012b) suggested an optimization model for biomass procurement to meet

the monthly electricity demand of a forest biomass power plant over a one-year time horizon.

Decision variables of the model included monthly harvesting levels from several forest cells.

This study focused on the procurement of biomass and did not include the whole supply chain

in an integrated framework. Moreover, none of the above mentioned studies included the

impact of biomass quality on the electricity production and its cost. Specifically, the quality of

different types of biomass in different months of the year, the quality of the mix of biomass in

the storage and the impact of storing biomass on the amount and cost of generated electricity

have not been studied previously.

There are some studies in the literature that consider uncertainty in biofuel supply chain

optimization models. Some of them evaluated the impact of uncertainty on the model solution

through sensitivity analysis and Monte Carlo simulation (Kim et al. (2011a), Rauch & Gronalt

(2011)). More advanced optimization techniques that incorporate uncertainty in the modeling,

e.g. stochastic programming (Kim et al. (2011b), Chen & Fan (2012), Awudu & Zhang (2013),

6

Kazemzadeh & Hu (2013)) and robust optimization (Tay et al. (2013), Bredström et al.

(2013)), have also been used in this supply chain. Uncertainty in the supply chain of forest

biomass power plants and its effect on electricity production cost are ignored in previous

studies. Alam et al. (2012b) have only addressed uncertainty in the supply chain of a forest

biomass power plant through sensitivity analysis and concluded that the impact of uncertainty

in moisture content on the production cost was significant. However, there is no study in this

area that include uncertainty into the decision making process. Instead, optimization models in

the literature provided results only based on the expected value of the uncertain parameters.

Ignoring uncertainty in deterministic optimization models may result in non-optimal and/or

infeasible solutions for real world case studies. Hence, optimization models need to be

extended to incorporate uncertainty and variations in the input parameters of the supply chain.

1.2 Research objectives

The main objective of this research is to develop tools (models) to help managers of biomass

power plants to achieve greater efficiencies (profit) by better managing their supply chains,

especially by considering uncertainty in the supply chain. The specific objectives of this study

are as follows:

1. To optimize the supply chain of a forest biomass power plant considering biomass

supply, storage and electricity production in an integrated framework. To achieve this

objective a new mathematical programming model is developed to provide decisions on

the amount of biomass to be purchased, stored and consumed in each month over a one-

year time horizon to maximize the profit.

2. To apply the developed model to a real case study in Canada.

7

3. To evaluate the impact of changes in uncertain parameters on the solution of the

mathematical programming model. This objective is achieved by examining historical

data and performing sensitivity analysis, scenario analysis and Monte Carlo simulation.

4. To incorporate uncertainty in different parameters into the supply chain decision

making process. This objective is reached by developing stochastic programming and

robust optimization models, and a hybrid model. Using different modeling approaches

allowed the inclusion of uncertainty in different parameters, i.e. biomass quality and

availability. The hybrid model should incorporate uncertainties in different parameters

in the model simultaneously.

1.3 Organization of the dissertation

In addition to the introduction chapter, this dissertation includes a chapter on the literature

review, a chapter on the deterministic optimization model, a chapter on the Monte Carlo

simulation model, a chapter on stochastic programming model, a chapter on hybrid stochastic

programming-robust optimization model and a chapter on conclusions, limitations of the study,

and suggestions for future research.

The previous studies on forest bioenergy supply chain in the literature are discussed in Chapter

2. They are categorized based on the inclusion of uncertainty.

In Chapter 3, an optimization model is proposed for the supply chain planning of a forest

biomass power plant over one year. Later in this Chapter, the model is applied to a real case

study in Canada and different scenarios are evaluated and sensitivity analysis is also performed

to assess the impact of different scenarios as well as variations in input parameters on the

generated profit.

8

Monte Carlo simulation along with the optimization model is performed and presented in

Chapter 4 to provide the ranges of generated profit and its distribution when input parameters

vary. Historical data on input parameters (biomass quality, cost, availability and electricity

price) as well as data analysis are presented in this Chapter. From the results of the Monte

Carlo simulation model, risks of having low profit and low or high storage levels associated

with uncertainty in model parameters are identified.

In Chapter 5, uncertainty in the available monthly supply is incorporated in the decision

making by developing a two-stage stochastic model and two different risk measures, variability

index and downside risk, are also considered.

In Chapter 6, uncertainty in biomass quality is also added to the decision making process

through developing a robust optimization model. Then, a hybrid stochastic programming-

robust optimization model is proposed to include uncertainty in different parameters

simultaneously. The final conclusions, strengths and limitations of the study and some

suggestions for future research are presented in the last chapter, Chapter 7.

9

Chapter 2 Literature review

2.1 Synopsis

In several previous studies, optimization techniques have been employed to manage the forest

bioenergy supply chain for heat, electricity and biofuels production from strategic, tactical and

operational points of view. Most of these studies were deterministic and ignored uncertainty,

while there are examples that included uncertainty in the supply chain models especially during

the past few years. This chapter covers major relevant studies on optimization of forest

bioenergy supply chains. It also discusses the issue of uncertainty in this supply chain,

uncertainty sources and the methods used for dealing with it. The studies are categorized into

two groups: 1) studies that used deterministic mathematical programming for modeling forest

bioenergy supply chains, and 2) studies that considered uncertainty in the analysis of forest

biomass supply chains. Studies that used deterministic models are categorized based on the

type of bioenergy plants into those related to power plants, district heating plants, co-

generation plants and biofuel plants. Studies that considered uncertainty are categorized based

on the modeling approaches. Some examples from other forest product industries as well as

other biofuel industries that included uncertainty in their supply chain optimization are also

presented. The strengths and shortcomings of the relevant literature are highlighted at the end

of this chapter.

2.2 Deterministic optimization models

Different optimization techniques, such as linear programming (LP) and mixed integer linear

programming (MILP), have been used for supply chain design and management. LP is a

mathematical method which includes a set of variables to be determined, a linear objective

function to be optimized, and a set of linear equality or inequality constraints to be met. The

10

main advantages of using this optimization method are its ability to solve large scale problems,

its assured convergence to global optimum solutions, having no need to have an initial solution

and its use of a well-developed duality theory for sensitivity analysis and the ease of problem

formulation (Labadie, 1997). If some of the variables in LP are integers, the model is called

mixed integer linear programming (MILP). In this section, the studies that optimized the supply

chain of electricity plants, district heating systems, co-generation plants and biofuel plants

using forest biomass are reviewed.

2.2.1 Power plants

Forest biomass can be used in power plants directly or in combination with fossil fuels for

generating electricity. It can be burnt at a constant rate in a boiler furnace to heat water and

produce steam. The steam is then carried through the furnace using pipes to raise its

temperature and pressure further. Finally, the steam passes through the multiple blades of a

turbine, spinning the shaft, which runs an electricity generator which produces an alternating

current to use locally or to supply the national grid (Demirbaş, 2001).

The optimal supply area and location of a forest biomass power plant in a distributed power

generation system was determined by Reche et al. (2008). The objective function was to

maximize the profitability index as a function of the net present value of benefits from the sale

of electrical energy minus the initial investment, collection, transportation, maintenance and

operating costs. The authors used an artificial intelligence method, called particle swarm

optimization. They concluded that it is important in distributed generation systems to consider

the technical constraints of the network and the voltage regulation. Finally, they evaluated the

model performance using simulation.

11

Alam et al. (2009) constructed a three–objective model for optimizing the amount of each

individual type of biomass from each of the harvesting zones, and then applied their model to a

50 MWh biomass power plant using both harvesting residues and poplar trees collected from

three management zones in Northwestern Ontario, Canada. To optimize the supply chain of

energy plants, it is sometimes necessary to formulate a problem with more than one objective

since single objective models cannot always represent the problem accurately. The objectives

are often in conflict (minimizing and maximizing objectives) and it might not be possible to

achieve an optimal solution that optimizes all the objectives simultaneously. In this situation,

the trade-off between objectives can be shown and the most efficient solution is selected. In

Alam et al. (2009), pre-emptive goal programming was applied to give priority to the

objectives as follows: 1) minimizing the procurement cost of feedstock, 2) minimizing the

transportation distance of biomass supply to the plant, and 3) minimizing the feedstock

moisture content. Alam et al. (2012b) developed a GIS based integrated optimization model to

optimize the supply chain of the forest biomass power plant with 230 MW capacity. The power

plant was fed by two biomass types: harvesting residues (leftover tops, branches and other parts

of the trees harvested mainly for lumber and pulp and paper industries) and unutilized biomass

(non-merchantable trees, and trees damaged by wildfire, wind and insects). GIS data were used

to estimate transportation costs from each forest cell to the power plant. The decision variables

were the harvest levels of two types of woody biomass in each month. The objective function

was to minimize the total piling, processing, felling, extraction and transportation costs.

Finding the optimal size, location, supply area and net present value of an electricity plant in

Spain was studied in Vera et al. (2010). The raw material of the power plant was olive tree

pruning residues and the technology for electricity generation was gasifier with gas turbine.

12

The authors used GIS data for the location and number of olive trees per km2, roads,

topographical features, electric line locations, etc. Different plant sizes and locations were

considered and the optimal one with the highest net present value was determined using three

metaheuristic methods. These methods were Genetic Algorithms (GA), Binary Honey Bee

Foraging (BHBF) and Binary Particle Swarm Optimization (BPSO). It was concluded that

BHBF algorithm converged to the optimal solution better than BPSO and GA. The results

indicated that the optimal plant size was 2 MW and the predicted optimal location of the plant

was in the area with highest available biomass.

Pérez-Fortes et al. (2014) developed an optimization model to determine location-allocation

and the selection/capacity of different pre-treatment technologies for feeding biomass to an

already existing coal combustion power plant. They included biomass transportation and

storage in their model. Different pre-treatments technologies were considered including

torrefaction, torrefaction combined with pelletization, pelletization, fast pyrolysis and fast

pyrolysis combined with char grinding. Changes in biomass quality (moisture content, dry

matter, energy density and bulk density) through the use of pretreatment processes were also

studied.

Table 2-1 summarizes the studies on optimization models used for modeling forest biomass

power plant supply chain.

13

Table 2-1: Summary of studies on deterministic optimization of forest biomass power plants

Author-Year-

Region Objective Function Decision Variables Method

Reche et al.

(2008)

Spain

Maximizing profitability index (net

present value of revenue from selling

electricity minus initial investment,

biomass collection and transportation

costs, and maintenance and operation

costs)

Location and supply area

of the biomass power plant

Particle

swarm

optimization

Vera et al.

(2010) Spain

Maximizing net present value (revenue

from the sale of electrical energy minus

initial investment and collection,

transportation, maintenance and operation

costs)

Plant size and location

Supply area

Several

metaheuristic

methods

Alam et al.

(2009)

Canada

Minimizing total biomass procurement

cost

Minimizing total distance for procurement

of biomass

Maximizing the quality of biomass

(minimizing moisture content)

Quantity of biomass

procured from each supply

location to each plant

Biomass procurement zone

selection (Binary variable)

Multi

Objective

Programming

Alam et al.

(2012b)

Canada

Minimize the total piling, processing,

felling, extraction and transportation

costs.

The harvest levels of two

types of woody biomass in

each month

Non-linear

dynamic

programming

Pérez-Fortes

et al. (2014)

Spain

Minimizing net present value of

investment and operational costs

Maximizing the environmental impact of

adding biomass to a coal power plant

Location/allocation and

selection/capacity of pre-

treatment technology

Fraction of coal replaced

by biomass

Material flow

Multi-

objective

Mixed

Integer

Programming

(MIP)

2.2.2 District heating plants

Forest biomass for energy generation is mainly used in district heating systems. These systems

consist of a central plant producing heat which is sent to a group of users (customers) through a

network of pipelines in the form of hot water or steam (Gilmour & Warren, 2007). Several

authors developed optimization models for supply chain design and management of heating

14

plants. Eriksson & Björheden (1989) developed a model with decision variables related to

storage and the chipping locations for a heating plant. Gunnarsson et al. (2004) developed a

mixed integer programming model for tactical-strategic supply chain management of forest

fuel used in a heating plant in Sweden by focusing on supply procurement decisions rather than

the production process. Multiple time steps were considered in this model. It was used to solve

six generated problems rather than being applied to a real case study. The results of using

different solution methods (LP and IP, and IP heuristic) to solve the problems were compared

in this work. Strategic decisions such as plant size and location were studied in Chinese &

Meneghetti (2005) and Schmidt et al. (2010). The most profitable configuration (plant size) of

a multi-source biomass district heating plant in Italy was considered in Chinese & Meneghetti

(2005). The model developed by Kanzian et al. (2009) included 16 combined heat and power

plants and 8 terminal storages in Austria. Optimum locations of bio-energy plants were studied

in Schmidt et al. (2010) with a case study in Austria. In another study, done by Van Dyken et

al. (2010), a linear mixed-integer model was developed for biomass supply chains with

transportation, storage and processing operations over 12 weekly time steps considering

supply, constant demand, three different biomass products and two demand loads for chips and

heat. This study focused on operational supply chain planning and the developed model was

not applied to a real case study. A truck scheduling optimization model was developed in Han

& Murphy (2012) for transportation of four types of forest biomass to energy plants in Oregon,

US. This study only considered the transportation part of the supply chain.

Table 2-2 summarizes all of these studies with their objective functions and decision variables.

15

Table 2-2: Summary of studies on deterministic optimization of district heating plants

Author -Year-

Region Objective Function Decision Variables Method

Eriksson &

Björheden

(1989)

Sweden

Minimizing forest biomass

supply cost (chipping, storing

and transportation costs)

Flow of biomass direct or via storage

Chipping location

Linear

Programming

(LP)

Nagel (2000)

Germany

Maximizing annual profit

(revenue from sale of energy

minus investment cost, fixed

and variable costs, fuel cost

and waste disposal cost)

Level of heat produced by each

boiler at each time period

The capacity of the system

Whether or not to integrate a boiler

into the heating system (Binary

variable)

Mixed Integer

Programming

(MIP)

Gunnarsson et

al. (2004)

Sweden

Minimizing biomass supply

cost (transportation, chipping

and storage costs)

Flow of biomass within the supply

network

Quantity of biomass chipped and

stored at roadside and terminal

If biomass is forwarded to or is

chipped at each roadside location,

each sawmill is contracted or not,

each terminal is used or not (Binary

variables)

Mixed Integer

Programming

(MIP)

Chinese &

Meneghetti

(2005)

Italy

Maximizing profit (revenues

from sale of energy and

charging customers with

connection fees minus boiler

investment, construction and

operating costs)

Heat produced by each boiler at each

time period

The capacity of the system

If a boiler would integrate to the

heating system or not (Binary

variable)

Mixed Integer

Programming

(MIP)

Frombo et al.

(2009a) Italy

Maximizing net annual profit

(revenue from sale of heat

and power minus felling and

processing, skidding,

highway transportation, plant

installation and management

costs)

Annual quantity of biomass

harvested from each supply area

The plant capacity for different

conversion technologies

Linear

Programming

(LP)

16

Author -Year-

Region Objective Function Decision Variables Method

Frombo et al.

(2009b) Italy

Maximizing net annual profit

(revenue from sale of heat

and power minus felling and

processing, skidding,

highway transportation, plant

installation and management

costs)

The quantity of biomass harvested at

each harvesting location and to be

used at each plant location.

The capacity of each plant

Selection of the conversion

technology (Binary variables)

Mixed Integer

Programming

(MIP)

Kanzian et al.

(2009) Austria

Minimizing biomass supply

cost to the heating plants

(chipping, storing and

transporting costs)

Volume of wood chips transported

from each terminal to each plant

Location of terminals and plants

(Binary variable)

Mixed Linear

Programming

(MIP)

Van Dyken et

al. (2010)

Norway

Minimizing the present value

of the costs (investment and

operating costs and salvage

value)

Biomass, product and energy flow

within the supply network

Emissions from storing and drying

biomass

Biomass input and output moisture

content to and from dryer

Biomass storage duration (Binary

variable)

Linear and

Mixed Integer

Programming

(LP and MIP)

Keirstead et

al. (2012) UK

Minimizing system cost

(biomass purchase, storage,

transportation and conversion

costs)

Optimal capacity of boilers

Whether chipped forest biomass

should be imported from neighbor

area or non-chipped residues should

be imported and then chipped within

the area (Binary variable)

Mixed Integer

Programming

(MIP)

Han &

Murphy

(2012) US

Minimize the weighted sum

of transportation costs

Minimize the total working

time

Truck schedules Simulated

Annealing

2.2.3 Co-generation plants

Combined heat and power (CHP) systems are energy plants that use cogeneration technology

to produce both heat and power in a district heating system (Gilmour & Warren, 2007). In

some studies, mathematical programming was used to compare the cost of generating either

energy or biofuels from biomass and evaluate the possibility of co-generation.

17

Some studies indicated that utilizing biomass for energy generation is more cost effective than

for biofuel production. Azar et al. (2003) concluded that utilizing biomass for generating heat

was the most economical scenario. Wahlund et al. (2004) showed that using wood biomass for

pelletization would have a lower cost and higher CO2 reduction than using it for biofuel

production. Feng et al. (2010) investigated the possibility of having bioenergy facilities (they

called them biorefineries) in typical existing sawmills, pulp and paper mills, wood panel

facilities, biochemical, energy, and pellet facilities. Then, the authors developed a

mathematical model to design this integrated supply chain optimally.

A methodology for optimizing the utilization of distributed biomass resources for energy

production was proposed by Alfonso et al. (2009). The main focus of the paper was to optimize

the logistics, but economic and environmental analyses of different bioenergy alternatives were

also performed. The authors indicated that the methodology would provide the optimal

locations of the biomass plant, energy application (electricity, heat and/or standardized biofuels

such as pellets), and the employed technology. This methodology was applied to three districts

in Spain. Based on the results, the authors concluded that the shortest payback period and

highest CO2 savings were attained from cogeneration plants, followed by pellet plants. The

least ranked option was power-only power plants.

Difs et al. (2010) analyzed different biomass gasification scenarios, and determined the

optimum configuration with the current fossil fuel price and green energy policies. Wetterlund

& Söderström (2010) considered two scenarios of co-generating Synthetic Natural Gas (SNG)

and district heat, and co-generation of heat and power. The authors determined the policy

support levels (tradable biofuel certificates) that would make the SNG scenario cost

competitive with CHP, while maximizing the annual profit over a 20-year time period. The

18

cost-effectiveness of different applications of biomass gasification was analyzed by Börjesson

& Ahlgren (2010). The focus of this study was to determine whether CHP generation in

biomass integrated gasification combined cycle (BIGCC) plants, and biofuels production in

biomass gasification biorefineries in a case study in Sweden were cost efficient. Schmidt et al.

(2010) used a mixed integer linear programming model for optimizing the location of

bioenergy plants using forest biomass in Austria. The bioenergy plants included integrated

gasification combined cycle (IGCC) system and biomass CHP plants with carbon capture

storage (CCS), pellet mill, and transportation fuel (methanol and ethanol) plants.

The problem of indicating whether to produce electricity in addition to heat at biomass

combustion plants was studied by Freppaz et al. (2004). A decision support system (DSS) was

presented by Rentizelas et al. (2009a) to optimize a multi-biomass energy conversion system to

generate electricity, heating and cooling in an area in Greece. The authors concluded that

considering multi-biomass supply chain reduced the cost by decreasing warehousing

requirements, especially for seasonal types of biomass. The developed model was non-linear

and a hybrid optimization method was used to solve that. Rauch & Gronalt (2011) developed a

model for designing a forest fuel CHP plant supply chain in Austria.

The summary of studies on modeling co-generation plants is provided in Table 2-3.

19

Table 2-3: Summary of studies on deterministic optimization of co-generation plants

Author- Year-

Region Objective Function Decision Variables Product Method

Freppaz et al.

(2004) Italy

Maximizing annual profit

(revenues from sale of

energy minus harvesting,

transportation, installation

and maintenance, and

energy distribution costs)

Annual quantity of biomass

harvested at each collection

area and transported from

each collection area to each

of six district energy systems

Capacity and the percentage

of thermal energy generated

at each plants

If the plant produces

electricity or not (Binary

variable)

Heat/

Electricity

Mixed

Integer

Programming

(MIP)

Alfonso et al.

(2009) Spain

Minimize transport

duration, optimize the

location, etc.

Biomass resources, logistics

structure, bioenergy plants

size and location, technology

type, etc.

Co-

generation

Did not

mention

Rentizelas et

al. (2009a)

Greece

Maximizing the financial

yield of the investment

Location and size of the

bioenergy facility

The biomass types and

quantities

The maximum collection

distance for each type

Co-

generation

Hybrid

optimization

Börjesson &

Ahlgren

(2010)

Sweden

Not discussed in the paper

The optimal production

capacity at different subsidy

levels.

Selection of alternative

technologies for district heat

generation (Binary variable)

Biofuel/

Heat

Mixed

Integer

Programming

(MIP)

Difs et al.

(2010)

Sweden

Maximizing annual profit

(revenues from sale of

energy products minus

investment, fuel and

maintenance costs)

Capacity of new investment

Selection of investment

alternatives for future

(Binary variable)

Co-

generation

Mixed

Integer

Programming

(MIP)

20

Author- Year-

Region Objective Function Decision Variables Product Method

Schmidt et al.

(2010)

Austria

Minimizing total cost of

energy generation (costs of

biomass supply and

transportation, energy

generation, carbon capture

and storage, plant building

and distribution network

investment and

distribution)

The annual amount of

energy commodities

produced at plants: heat,

power, pellets, and

transportation fuels.

Plant size and location,

pipeline networks selection

and district heating networks

selection (Binary variables)

Fuel/

Energy

Mixed

Integer

Programming

(MIP)

Wetterlund &

Söderström

(2010)

Sweden

Maximizing annual profit

(revenues from sale of

electricity and synthetic

natural gas minus

investment, fuel and

maintenance costs)

The optimal government

support level (subsidy)

Selection of new investment

alternatives (Binary

variable)

Co-

generation

Mixed

Integer

Programming

(MIP)

Rauch &

Gronalt

(2011)

Austria

Minimizing total

procurement cost

(transport, chipping

investment, operations and

maintenance costs)

The annual volume of fuel

transported between

districts, terminals, regional

departure railway, and the

CHP plant

Open or close a terminal

(binary variable)

CHP

Mixed

Integer

Programming

(MIP)

2.2.4 Biofuel plants

Bioethanol is a type of fuel that is extracted from biomass through fermentation (Limayem &

Ricke, 2012). The bioethanol production has increased in recent years in many countries, such

as the U.S. Although most of the bioethanol is produced from agricultural biomass, the

controversial issue of using plants as fuel instead of food made it necessary to look for more

acceptable sources, namely forest biomass (Limayem & Ricke, 2012). Generating biofuels

from forest biomass is still in the developing phase and has not been commercialized yet. The

main challenges in commercialization of this technology include high energy or chemical

consumption for woody biomass pretreatment, even when compared to agricultural biomass,

21

low system efficiency, process scalability and intensive capital investment (Zhu & Pan, 2010).

In most of the studies presented here, forest biomass combined with agricultural biomass was

used for biofuel production

Some previous studies considered biomass supply chain management for generating biofuels.

Chinese & Meneghetti (2009) considered a real-life problem of supplying a biofuel plant with

forest fuel. A mixed-integer linear programming model was proposed to determine the optimal

configuration of the supply chain. It was mentioned that the model could be useful in resolving

trade-offs between decentralized early treatment of biofuels, resulting in lower transportation

costs, and centralized final treatment, allowing to reap the benefits of economies of scale. It

was therefore advised to apply integrated supply chain planning concepts to design biofuel

logistics systems and to support policy making in the energy field. An MIP model was also

developed by Ekşioğlu e al. (2009) for designing the biorefinery supply chain producing

cellulosic ethanol from agricultural and woody biomass. The model outputs were the number,

size and location of biorefinery plants with the objective of minimizing the total cost of annual

harvesting, storing, transporting and processing biomass; storing and transporting ethanol; and

locating and operating bio-refineries. The model included constraints on biomass availability,

flow, conversion, production and inventory capacities, and demand. The data from the State of

Mississippi was used to validate the model. The author concluded that transportation costs,

biomass availability, technology type, and planting and harvesting costs are important factors

in supply chain design decisions.

Kim et al. (2011a) developed a mixed integer linear programming optimization model for the

supply chain design of bio-gasoline and biodiesel production from six forestry resources

(logging residuals, thinnings, prunings, inter-cropped grasses, and chips/shavings). The first set

22

of conversion plants could be from a set of candidate sites with four capacity options to convert

biomass to three product types: bio-oil, char and fuel gas. These intermediate products could be

used either as local fuel sources or as feedstock to produce final products (gasoline and

biodiesel) at the second conversion plants, which could be from a set of candidate sites with

four capacity options. There were several possible markets for the final products with certain

maximum demands. The objective of the model was to maximize the overall profit by

determining the number, location, and size of the conversion plants, biomass supply locations,

the logistics and the amount of materials to be transported between the various nodes of the

designed network, while satisfying the demand constraints. The considered case study was

based on an industrial database related to a case in the Southeastern United States. The authors

evaluated the trade-off between centralized and distributed network designs.

The trade-off between economic and environmental objectives in the optimal planning of a

biorefinery in Mexico was evaluated in a iba e -Aguilar et al. (2011). The authors used a

multi-objective optimization model for selecting the feedstock type, processing technology,

and a set of products in a biorefinery supply chain. The raw material contained different types

of agricultural biomass, wood chips, sawdust, commercial wood for producing ethanol,

hydrogen, and biodiesel (generated only from agricultural biomass). The objectives were: 1) to

maximize the profit considering the costs of feedstock, products, and processing, and 2) to

minimize the life cycle environmental impacts. The authors applied their model to a case study

in Mexico. The decision makers could select from the output the solutions that fit the specific

requirements and compensate for both objectives simultaneously.

Table 2-4 summarizes the studies on optimization models used for modeling the supply chain

of forest biofuel plants.

23

Table 2-4: Summary of studies on deterministic optimization of biofuel plants

Author-Year-

Region Objective Function Decision Variables Method

Chinese &

Meneghetti

(2009) Italy

Minimizing total cost of

supply chain (harvesting,

transportation, processing

and facility installation

costs)

Flow of biomass within the supply

network

Whether to use a preprocessing

equipment or not (Binary variable)

Mixed Integer

Programming

(MIP)

Ekşioğlu e al.

(2009) USA

Minimizing total annual

cost (investment,

harvesting, storing and

transportation costs)

Number, size, and location of bio-

refineries required

Quantity of biomass harvested,

shipped, processed and stored

Whether a biorefinery and a collection

facility with specific size are located

in each site (Binary variables)

Mixed Integer

Programming

(MIP)

Kim et al.

(2011) US

Maximizing the overall

profit

Number, location, and size of the

conversion plants

Biomass supply locations

Logistics and the amount of materials

to be transported between the various

nodes of the designed network

Mixed Integer

Programming

(MIP)

a iba e -

Aguilar et al.

(2011)

Mexico

Maximizing profit (revenue

from sale of products minus

investment, process,

operating and transportation

costs)

Minimizing environmental

impacts

The quantity of products produced

from different biomass feedstock

using different processing routes

The quantity of each biomass

feedstock used for producing different

products through different processing

routes

Multi-objective

Programming

2.3 Optimization models with uncertainties

Uncertainty refers to the lack of information or the lack of certainty in the validity of the

information about the existing or future state of a system (Kangas & Kangas, 2004). It can

result from measurement errors and ignorance, which is to some extent inevitable and might be

reduced by further studies or investing in improved technology to acquire high quality data

(Petrovic (2001), Ells et al. (1997)). It may result from variability in random future events due

24

to their inherent nature (such as feedstock characteristics) (Bowyer et al. 2012), which can be

controlled to some extend by employing better forecasting methods and/or using expert

judgment. It can also result from lack of reliable historical data or lack of certainty in historical

data, for example lack of data on the demand of a new product. Other sources of uncertainty

include imprecision in judgment, vagueness, and ambiguity related to the known objects, which

belong to poorly defined sets so they cannot be classified well (Kangas & Kangas (2004),

Petrovic (2001), Ells et al. (1997)). From the system boundary point of view, the source of

uncertainty may exist outside the production process, called environmental uncertainty, such as

uncertainty in demand and supply. It may also be within the production process, called system

uncertainty, such as uncertainty in lead time due to machine failure (Chrwan-JYH, 1989). In

terms of time horizon, uncertainty may be the result of short term variations, such as day-to-

day processing variations, cancelled/rushed orders and equipment failure, or long-term

variations, such as raw material/final product unit price fluctuations, seasonal demand

variations and technology changes. Therefore, uncertainty exists in supply chains at strategic,

tactical and operational levels and should be considered in supply chain decisions.

There are several reasons why uncertainties exist in the biomass supply chain. Some of the

sources of uncertainty in forest biomass supply chains are similar to those in other industries,

such as economic fluctuation and instability, raw material supplies, manufacturing process

time, machine breakdowns, reliability of transportation channels, and exchange rates. However,

there are other sources of uncertainty that are related specifically to the characteristics of forest

biomass supply chains which are summarized here:

Interdependency between different forest sectors: There are interdependencies between

different sectors and markets within the forest industry supply chains. This means that

25

raw material of one sector could be the product of another sector. Consequently,

variations in one part of the supply chain usually propagate into the other parts.

Variations in feedstock supply: The need for having a continuous supply of raw

material for a bioenergy facility necessitates the use of a mixture of materials or even to

have new sources of material. Even when one type of biomass is used for energy

production, the quality of biomass varies over time. Therefore, this industry must have

a dynamic supply chain.

Wood is a heterogeneous natural material: Its physical and chemical characteristics

affect the quality and quantity of the products (Bowyer et al. 2012). In the bioenergy

industry, the moisture content and heating value of raw material play an important role

in the amount of produced energy and its costs (Saidur et al., 2011). Heating value and

moisture content vary from one tree stands or species to another (Demirbaş (2001),

Carlsson et al. (2009), Demirbaş (2003)) and also differ in different types of biomass

(e.g. bark, sawdust, shavings) (Lehtikangas, 2001). Wood properties may be affected by

external factors, such as growth condition, climate, harvesting methods, storage and

transportation methods. Biomass quality, such as moisture content, can also change

during storage, production, and transportation.

Divergent production structure: Unlike most of other manufacturing industries, which

have an assembly structure, forest products industries generally have a divergent

production structure. This means that multiple products, by-products, and co-products

are made from a single product simultaneously. Consequently, it is difficult to

completely control the manufacturing processes. Moreover, it is challenging to forecast

the quality and quantity of outputs due to this production structure and the use of

26

heterogeneous natural raw material in the production. This fact can impact the amount

of raw material available for bioenergy plants, which are supplied by other forest

product mills.

Ambiguous values and objectives: Most forest areas include large areas with diverse

geographical and ecological characteristics. In forestry, it is usually needed to

i corpora e differe values a d s akeholders’ prefere ces a d i eres s which

sometimes cannot be understood, interpreted or quantified completely (Ells et al.,

1997). Therefore, it is likely to have vague factors, values and objectives which can also

exist in the forest bioenergy supply chain. This aspect of uncertainty cannot be dealt

with like other sources of uncertainty. To some extent, it is possible to spend time and

money in some form of consulting with the stakeholders to get a better understanding of

their preferences, opinions, and values. However, sometimes the stakeholders may not

be able to express their preferences before a specific decision is made.

New markets and new production technologies: Investment grants, carbon and energy

taxes, green certificate schemes, conversion technologies, and availability and quality

of biomass resources may not be known with certainty (Mccormic, 2011). For example,

in designing and planning a biomass power plant, it may be hard to estimate the long

term availability, quality and cost of biomass. Alternatively, market demand for

biofuels may not be lucid from the beginning.

In general, uncertainty can be dealt with at the source, or it can be dealt with during the process

of decision making. When uncertainty is ignored, decision making is based on the expected

values of stochastic parameters, which may be different from their actual values and may lead

to non-optimal or infeasible results and solutions. Considering uncertainty in decision making

27

usually helps companies safeguard against threats while simultaneously taking advantage of the

opportunities that higher levels of uncertainty would provide. It also makes decisions robust

and mitigates the effect of the variations and perturbation on the optimal solution. The

modeling approaches for dealing with uncertainty in optimization models are presented next.

2.3.1 Modeling approaches

The method for dealing with uncertainty depends on the type of uncertain parameter, the source

of uncertain parameter, available data on the uncertain parameter, the computational effort

needed for each method, and the degree of sophistication that can be handled and accepted by

the users and decision makers. In the literature, several approaches were used to incorporate

uncertainty in the supply chain design including sensitivity analysis, scenario-based

approaches, Monte Carlo simulation, stochastic programming, and robust optimization. Mula et

al. (2006) provided an overview of the uncertainty in production planning. Based on their

classification, the supply chain planning problems with uncertainty are usually solved by

conceptual, analytical or artificial intelligent based approaches.

I de ermi is ic models, a si gle bes guess or a si gle fu ure “sce ario” represe s a

uncertain parameter based on its expected values. Through sensitivity analysis, model

sensitivities are tested to determine variations in model outcomes when input parameters

fluctuate around best guesses. In the scenario analysis approach, multiple scenarios for

uncertain parameters are generated and then the optimization model is solved for each

individual scenario. This provides an extensive what-if analysis which also helps in evaluating

the outcomes based on different realizations of the stochastic parameters. Scenario generation

itself is a challenging task which can be done using historical data, forecasting methods,

managerial and expert judgment, etc. (Benders (1962), Shapiro (2004)).

28

Similar to scenarios analysis, in Monte Carlo simulation a number of scenarios are generated,

but one further step is taken in this method by considering every possible value that each

stochastic parameter could take using its probability distribution. Moreover, each scenario is

weighted by the probability of its occurrence. In other words, in this method the deterministic

model is solved repeatedly with its stochastic input parameter based on a probability

distribution function instead of a single value. The Monte Carlo simulation method 1)

determines a possible distribution of the model outcomes, 2) evaluates model robustness and

behavior in the presence of uncertainty in input parameters, 3) determines regions of input

parameters that result in particular levels of the optimal solutions, and 4) identifies possible

risks and opportunities that result from uncertainties in the system. The process of developing

and implementing Monte Carlo simulation involve: 1) determining the ranges and distributions

of each stochastic input parameter, 2) generating samples from the specified ranges and

distributions, 3) running the model for these samples, and 4) evaluating and analyzing the

outputs (Pannell (1997), Vose (2008), Saltelli et al. (2008), Kim et al. (2011b)).

The problem with all of these approaches is that they do not provide a single overall optimal

solution for all scenarios. Stochastic programming is an approach which overcomes this

problem. In stochastic programming, it is assumed that accurate probabilistic descriptions of

the random variables such as probability distributions, densities or other probability measures

are available. In this method, the expected objective value of different potential scenarios is

optimized. In a two-stage stochastic model, decision variables are divided into two groups

called the first stage variables (control, here-and-now), which are made before the realization

of the uncertain parameters, and the second stage variables (state, wait-and-see), which are

taken after the realization of the uncertain variables. The output of such a model is the optimal

29

single first-stage policy and a set of recourse decision rules that determine which second-stage

action should be taken in response to each random variable. One of the advantages of

developing stochastic programming models is in their capability to manage the risk associated

with the supply chain performance (Birge & Louveaux (1997), Shapiro (2004)).

While stochastic programming seems to be an adequate and attractive method for addressing

uncertainty when it is possible to define potential scenarios, it is computationally intractable

when the value of an uncertain parameter covers a continuous range, unless it is approximated

by a set of scenarios derived from discretizing the uncertainty sets. The problem, hence, is that

even for a small number of discretized scenarios, the total number of scenarios will grow

exponentially when one deals with a sequence of scenarios, e.g. a scenario tree. This again

results in being computationally intractable and not being able to have a ready-to-use model

(Ben-Tal et al. 2000). Approximation and decomposition methods are being used to address

this issue (Benders, 1962). Another alternative, however, is robust optimization which is

attractive since it can be solved effectively and efficiently using the current powerful solvers if

a tractable uncertainty set is selected (Ben-Tal & Nemirovski, 2000). In stochastic

programming, the objective is to find solutions that are feasible for all realization of uncertain

parameters while optimizing the expected value of the objective function over all scenarios. In

robust optimization, the objective is to find solutions that are feasible for all realization of

uncertain parameters while optimizing the worst case performance of the system. Moreover,

contrary to stochastic programming, to incorporate uncertainty in robust optimization only a

range of uncertain parameters is required (Gabrel et al. 2013).

30

The formulation of robust optimization depends on the definition of a robust solution. A robust

solution as defined by Mulvey et al. (1995) is a solution that is not far from the optimum

solution. The authors developed a model based on scenario analysis with the objective of

providing less sensitive solutions to the realization of the model data and solved it using goal

programming. In another definition, a robust solution is a solution that must be feasible for any

realization of the uncertain parameter, or it is a solution that its objective function value must

be guaranteed. This approach was originally proposed by Soyster (1973) and later on was

co sidered as “ul raco serva ive s ra egies”. Bertsimas & Sim (2003) and (2004) suggested an

approach ha uses he idea of “budge of u cer ai y” o co rol he level of co serva ive ess.

In this method, only some of the uncertain parameters deviate from their nominal values

simultaneously. Using this definition, a constraint is immunized against uncertainty by

de ermi i g he si e of he buffer or a “pro ec io fu c io ” of i . This pro ec io fu c io is

itself an optimization model and its dual is embedded in the original model. Given the linearity

of the original problem, the robust counterpart is also a linear problem with a modified feasible

region. Ben-Tal & Nemirovski (2000) suggested less conservative approaches which were

nonlinear. In all of these methods, the solution is optimized based on the worst case, which is

the most unfavourable realization of the uncertainty. The worst case can be selected differently

too, either from a finite number of scenarios, such as historical data, or continuous, convex

uncertainty sets, such as polyhedral or ellipsoids. For a recent review of robust optimization the

reader is referred to Gabrel et al. (2013).

There are few studies which considered uncertainty in the forest biomass supply chain. These

studies as well as some other studies which included uncertainty in other forest industries and

biofuel supply chain are reviewed here.

31

2.3.2 Sensitivity analysis and Monte Carlo simulation

Sensitivity analysis and scenario analysis are usually done after an optimization model is

developed. One study which used sensitivity analysis in the forest biomass supply chain was

conducted by Kim et al. (2011a). They ran their model for different demands (100%, 90%,

75% and 60%) to evaluate the effect of changes in demand on the optimal network design. The

results showed the total profit for the distributed system was higher than that for the centralized

design at 100% demand. When the demand decreased, the profit difference between the two

systems reduced as well. In this study, although demand uncertainty was evaluated through a

very simple sensitivity analysis method, it showed the importance of considering uncertainty as

it affected the optimal result. Alam et al. (2012b) performed sensitivity analysis by testing

sixteen scenarios for different harvesting levels, processing and felling costs, conversion

efficiency, moisture content, energy density and equivalency of energy. They concluded that

moisture content and conversion efficiency had more impact on the cost compared to the other

parameters. This paper incorporated variations in biomass quality, however, a fixed range was

used to capture variations and only sensitivity analysis was performed. In Rauch & Gronalt

(2011), the effect of changes in forest fuel supply (domestic resources or imports), transport

modes (truck only, truck and ship, or truck, ship and rail), energy price (increase by 0, 20% and

40%), and truck load capacity (50%, 40% and 30%) on the overall cost was evaluated. Eight

scenarios were constructed and compared.

In a number of studies, Monte Carlo simulation was used in optimization of forest and

agricultural bioenergy supply chains. Rozakis & Sourie (2005) developed an optimization

model for eco omic a alysis of a biofuel supply chai i Fra ce. They de ermi ed he efficie

tax exemption policies in presence of uncertainty in petroleum prices and feedstock prices

32

using Monte Carlo simulation. A mixed integer programming model was developed by

Schmidt et al. (2009) to determine optimum locations of bioenergy gasification plants in

Austria. The model considered the spatial distribution of biomass supply and biomass

transportation costs. The authors used Monte Carlo simulation to incorporate uncertainties in 9

parameters: annualized district heating costs, biomass supply, biomass costs, plant setup costs,

transportation costs, price local heat, carbon price, connection rate and power price. They used

literature and expert opinion to assign ranges to the uncertain parameters. An optimization

model of a biofuel supply chain was presented in the study done by Marvin et al. (2012). It

determined the location and capacity of biorefineries, and the amount of harvested biomass to

ship to biorefineries while maximizing the Net Present Value (NPV) of the entire supply chain.

The authors performed sensitivity analysis and Monte Carlo simulation to evaluate the impact

of uncertainty in costs, harvestable biomass and conversion factor on the robustness of the

supply chain. The output of the Monte Carlo model was a range of internal rate of return (IRR)

which was used to make conclusions about the probability of having IRR less than a certain

level (10%) and consequently having a negative NPV. It was also concluded from Monte Carlo

simulation results that it was not economical to construct any biorefineries in 21.5% of the

trials. Sharmaa et al. (2013) studied the weather uncertainty in biomass supply chain through

developing a scenario optimization model. The model had a one year planning horizon with

monthly time steps and the objective function of minimizing the cost of biomass supply to

biorefineries. Uncertainty in other parameters, e.g. yield, land rent and storage dry matter loss

was analyzed by sensitivity analysis.

33

Table 2-5 summarizes the previous studies which incorporated uncertainty in the forest

biomass supply chain and biofuel supply chain management and design through sensitivity

analysis, scenario analysis and Monte Carlo simulation model.

Table 2-5: Summary of studies on sensitivity analysis, scenario analysis and Monte Carlo

simulation applied to bioenergy supply chain with uncertainty

Author/ Year Uncertain parameter Method Case Study

Kim et al. (2011a) Demand Sensitivity analysis Biofuel plant in

Southeastern US

Alam et al. (2012b)

Harvesting levels,

processing and felling

costs, conversion

efficiency, moisture

content, energy density

and equivalency of energy

Sensitivity analysis

Forest biomass power

plant in Ontario,

Canada

Rauch & Gronalt

(2011)

Forest fuel supply,

transport modes, energy

price, and truck load

capacity

Scenario analysis CHP in Austria

Rozakis & Sourie

(2005)

Petroleum price and

feedstock prices

Monte Carlo

simulation Biofuel plant in France

Schmidt et al. (2009)

Annualized district heating

costs, biomass supply,

biomass costs, plant setup

costs, transportation costs,

price local heat, carbon

price, connection rate and

power price

Monte Carlo

simulation

Bioenergy gasification

plants in Austria

Marvin et al. (2012) Costs, harvestable biomass

and conversion factor

Sensitivity analysis

and Monte Carlo

simulation

Biorefinery plant in

Midwestern US

Sharmaa et al. (2013)

Weather uncertainty

Yield, land rent and

storage dry matter loss

Scenario analysis

Sensitivity analysis Biorefinery in the US

34

2.3.3 Stochastic programming

Some previous studies incorporated uncertainty in supply chain optimization of other industries

such as chemical and lumber industries (e. g. Gupta & Maranas (2003), You et al. (2009),

Kazemi Zanjani et al. (2010b)). The results of these studies mainly demonstrated that

incorporating uncertainties in the decision making of real case scenarios using stochastic

models provide more robust solutions compared to deterministic models. There are few studies

that included uncertainty in bioenergy supply chain models. A number of them are reviewed

and presented by Awudu & Zhang (2012).

Some of the studies used stochastic programming in modeling bioenergy and biofuel supply

chains. Kim et al. (2011b) performed a global sensitivity analysis and two-stage stochastic

programming on a biofuel supply chain model and evaluated the effect of uncertainty in

different parameters on the final result using Monte Carlo simulation. The authors concluded

that the most important uncertain parameters affecting the profit were the price of the final

product, the conversion yield ratios of the two conversion processes, maximum demand and

biomass availability. They then generated 33 scenarios from changing these five most

important uncertain parameters by ± 20% and developed an optimal model to maximize the

expected profit from all these scenarios plus the expected value scenario. The first stage

decision variables were the size and location of the processing plants and the second stage

recourse decision variables were flows of biomass and product within plants and markets. They

implemented the robustness analysis and Monte Carlo global sensitivity analysis to compare

the performance of the multiple scenario design with the single scenario design. It was

concluded that the impact of variations in stochastic parameters on the optimal solution was

mitigated in the model that optimized multiple scenarios. In this paper, uncertain parameters

35

were only changed within some range rather than using probability distributions. Moreover, it

only modeled the supply chain in one time-step and ignored the possibility of having

correlation between random parameters.

Strategic decisions regarding choosing investment options for heat savings and decreasing

energy imports or increasing energy exports in pulp mills under market uncertainty was studied

in Svensson et al. (2011). The objective function of the developed model was to maximize the

expected net present value of the investments. The decision variables were related to

investment in heat saving and energy conversion technologies as well as distribution of the

obtained heat from different energy conversion technologies. The uncertainty was considered

in future energy prices and policy instruments through a scenario tree of five different

combinations of several emission reduction policies, and electricity, lignin and bark prices. The

original optimization model was a mixed-binary linear programming model and uncertainty

was included using a multistage stochastic programming model. Each stage contained five

years and the total planning horizon was assumed to be 30 years. Svensson & Berntsson (2011)

introduced a methodology for making decisions about investment in new energy technologies

in industrial plants. The main focus of the paper was to include uncertainty in the energy

market. The authors used stochastic programming and scenario analysis to take into account

the uncertainty in CO2 charge, and crude oil and electricity prices. They then applied the

methodology to a pulp mill and presented examples of possible future investments to show the

usefulness of the proposed approach.

Chen & Fan (2012) developed a mixed integer stochastic programming model to incorporate

uncertainty in strategic planning of bioenergy supply chain systems. They considered

bioethanol production, feedstock procurement, and fuel delivery in an integrated model to

36

minimize costs. The raw material included both agricultural and forest biomass. A two-stage

stochastic programming model was developed and applied to a case study in California.

Uncertainty was considered in available feedstock supply and fuel demand. A Lagrangian

relaxation based decomposition solution algorithm, called the progressive hedging method, was

used to reduce the computational effort needed to solve the stochastic model. This method

decomposed a stochastic problem across scenarios by partitioning the original problem into

manageable sub-problems. The first stage decision variables were refinery and terminal

locations and sizes and the second stage decision variables were feedstock procurement and

transportation, ethanol production and transportation. In the baseline scenario, four discrete

demand scenarios with equal probabilities were considered and the results showed that the

stochastic programming provided lower expected with lower variation total cost than

deterministic solutions. In the second scenario, uncertainty in biomass supply was considered

using ten scenarios of feedstock with equal probabilities and a fixed demand. The authors

stated that the optimal solution of the bioethanol supply chain was not sensitive to the

uncertainty in supply. This study did not integrate different sources of uncertainty within a

single framework. It also ignored uncertainty in production yields and prices. In addition, the

model had single annual time-step and did not consider multiple shorter time variations.

A two-stage stochastic MILP modeling approach was proposed by Kostina et al. (2012) to

include uncertainty in the demand of an integrated ethanol-sugar supply chain. Different risk

measures were studied in their model including value at risk, opportunity value and risk area

ratio. Another MILP model with demand uncertainty was developed by Gebreslassie et al.

(2012) which minimized the design and planning cost of biofuels network and risk

simultaneously. Awudu & Zhang (2013) developed a stochastic production planning model for

37

a biofuel supply chain which included biomass suppliers, biofuel refinery plants and

distribution centers. They incorporated uncertainty in the demand and price in a single period

planning framework. Decision variables were the amount of raw materials purchased and

consumed and the amount of products produced. Kazemzadeh & Hu (2013) determined the

optimal design of biofuel supply chain with uncertainties in fuel market price, feedstock yield

and logistics costs. They developed two stage stochastic programming models with two

different objective functions (profit and conditional value at risk) and compared the results of

the two models. In Giarolaa et al. (2013), an MIP model was developed to provide optimum

design and planning decisions for a multi-period multi-echelon ethanol supply chain. The

model had multi-objectives, considering both the economic and environmental performances as

well as risk mitigation preferences. Moreover, uncertainty in feedstock cost and carbon cost

was captured through developing a two-stage stochastic model. For controlling risks, two risk

indices were considered: expected downside risk and value at risk.

Table 2-6 summarizes the previous studies, which incorporated uncertainty in the forest and

bioenergy supply chains using the stochastic programming method.

38

Table 2-6: Summary of studies on stochastic programming of forest and bioenergy supply

chains

Author/ Year Uncertain parameter Method Case Study

Kim et al. (2011b)

Price of the final

product, the conversion

yield ratios of the two

conversion processes,

maximum demand and

biomass availability

Two-stage stochastic

programming

Biofuel industry in

Southeastern United

States

Svensson et al. (2011)

Emission reduction

policies, electricity,

lignin and bark prices

Multi-stage stochastic

programming Pulp industry in Sweden

Svensson & Berntsson

(2011)

CO2 charge, fuel and

electricity prices

Multi-stage stochastic

programming Pulp industry in Sweden

Chen & Fan (2012) Feedstock supply and

demand

Two-stage stochastic

programming with

decomposition

Biofuel plants in

California

Kostina et al. (2012) Demand Stochastic programming Ethanol-sugar plants in

Argentina

Gebreslassie et al.

(2012) Demand Stochastic programming Biofuel plants in the US

Awudu & Zhang (2013) Demand and price Stochastic programming Biofuel plants

Kazemzadeh & Hu

(2013)

Fuel market price,

feedstock yield and

logistics costs

Two-stage stochastic

programming models Biofuel plants

Giarolaa et al. (2013) Feedstock cost and

carbon cost

Two-stage stochastic

model Ethanol plants

2.3.4 Robust optimization model

Robust optimization has been used and applied in several fields of study, such as facility

location and inventory management (Gülpı ar e al. (2013), Solyali et al. (2012)), resource

allocation and project management (Wiesemann et al. 2012), and in specific supply chain

39

optimization problems such as the refinery industry (Leiras et al. 2010). This method has also

been used and applied in forest industry problems. Palma & Nelson (2009) used the robust

optimization approach to incorporate uncertainty in volume and demand of two products over

the entire planning horizon in a harvest scheduling decision making model. In their model, they

assumed that the uncertain parameters were uniform and independently distributed within a

symmetrical range of values. Their results showed that the robust optimization model with

different protection levels provided lower objective function and infeasibility rates than the

deterministic model. They then used robust optimization in a bi-objective planning model with

random objective weights for forest planning problem (Palma & Nelson, 2010). The two

objective functions were the amount of employment and the proportion of old forest through

the planning horizon. Using the robust optimization method, no large changes in the weighted

sum of the objectives were expected even when the weights changed over time. In both studies,

they used Bertsimas formulation (Bertsimas & Sim (2003), (2004)). Kazemi Zanjani et al.

(2010a) developed a robust optimization model based on Mulvey et al. (1995) formulation for

sawmill production planning. Robust optimization was combined with a two-stage stochastic

programming model and the uncertainty in raw material quality was included in the model.

They used two variability measures: 1) solution robustness, which measured the variability of

the recourse cost in the stochastic model for any of the scenarios, and 2) model robustness,

which measured the feasibility of all scenarios. The result demonstrated the trade-off between

variability in backorder/inve ory cos (pla ’s robus ess) a d raw ma erial co sump io a d

expected backorder/inventory cost. Alvarez & Vera (2011) applied the robust optimization

methodology to a sawmill planning problem with uncertainty in the yield coefficients

associated with the cutting patterns. The authors employed Bertsimas formulation (Bertsimas

40

& Sim (2003), (2004)) and concluded that the solution provided by robust optimization is

feasible for a large proportion of randomly generated scenarios with moderate reduction in the

objective function value.

One of the by-products of the Kraft process in the pulp and paper industry is black liquor,

which is used for energy production. Tay et al. (2013) considered uncertainties in raw material

supply and product demand in integrated biorefineries using the robust optimization method

based on Bertsimas formulation (Bertsimas et al., 2011). The model was mixed integer

nonlinear programming with one time step. Different scenarios with associated probabilities

were considered for supply and product demand. The results identified the optimum capacity of

each process technology and its corresponding amount of biomass, intermediate and final

products. Bredström et al. (2013) developed a formulation for robust optimization to be applied

in the production planning problem with rolling time horizon. The method was based on the

decomposition approach and recourse decision making framework so that some decisions were

made after realization of uncertain parameters. A heuristic algorithm was proposed to solve this

model iteratively. Their case study was a biofuel heating plant with uncertain energy demand

over time. Carlsson et al. (2014) used the same approach in the pulp and paper industry

considering uncertainty in customer demand. They concluded that using the robust

optimization approach, building safety stock was not needed since the variation in the demand

would be considered in the decision making process.

Table 2-7 summarizes the previous studies which incorporated uncertainty in the forest and

bioenergy supply chains.

41

Table 2-7: Summary of studies on robust optimization of forest and bioenergy supply chain

Author/ Year Uncertain parameter Case Study

Palma & Nelson (2009) Volume and demand of products Harvest scheduling in Canada

Palma & Nelson (2010) Weight of objectives in a bi-objective

optimization problem Forest planning in Canada

Kazemi Zanjani et al.

(2010a) Raw material quality Sawmill industry in Canada

Alvarez & Vera (2011) Yield coefficients Sawmill planning in Canada

Tay et al. (2013) Raw material supply and product

demand Kraft pulp and paper industry

Bredström et al. (2013) Energy demand Biofuel heating plant

Carlsson et al. (2014) Demand Pulp and paper industry in Sweden

2.4 Discussion and conclusions

Mathematical programming and optimization techniques have been used in the design and

management of forest biomass supply chains. These models provided the optimum solution for

decisions related to the network design including technology choices, plant size and location,

storage location, mix of products and raw materials, logistics options, supply areas, and

material flows. The objective functions included profit/cost, CO2 emissions, travel time, etc.

Both deterministic and stochastic mathematical programming models were developed for forest

biomass supply chains.

Deterministic models are necessary and helpful but not sufficient for capturing all aspects of

forest biomass supply chains. Despite the interest and much effort in extending deterministic

models, only a number of previous studies considered uncertainty in optimizing forest

bioenergy supply chains. Understanding the problem characteristics, gathering sufficient data

and choosing the appropriate methodology are important in developing stochastic models. The

42

appropriate methods to be implemented depend on the characteristics of the problem, data

availability, form and type of uncertainties.

In the literature, sensitivity analysis and stochastic programming approach were used to deal

with uncertainties in supply, demand, prices, conversion yields, carbon tax and emission

reduction policies in forest biomass supply chains. None of the previous studies dealt with

uncertainty in biomass quality, such as moisture content and heating value, and its effect on the

produced energy and its costs. Moreover, most of the published papers did not consider supply

chain planning in different time steps, and instead modeled the supply chain system in a single

time-step optimization framework. The only study that provides a dynamic model is (Dal-Mas

et al., 2011), which is related to a corn-ethanol supply chain. Therefore, the temporal

uncertainty or seasonality of uncertain parameters in forest biomass power plant has not been

studied in any of the previous studies.

43

Chapter 3 Deterministic model

3.1 Synopsis

In this chapter, an optimization model is developed to maximize the overall value of a forest

biomass power plant supply chain and is applied to a real case study. The model has a one-year

planning horizon with monthly time steps. It includes all parts of the supply chain from

procurement, to storage, production and ash management. The effects of the quality (moisture

content, energy value, ash content) of different types of biomass purchased and mixed in the

storage area on the amount of generated ash, cost of ash handling, total production cost and

total amount of generated electricity are considered in the model. The results of the model are

analyzed and compared to the power plant situation in 2011. Moreover, scenario and sensitivity

analyses are performed to evaluate the impact of variations on the model solution.

3.2 The power plant supply chain

The supply chain of a power plant, including procurement of different raw materials from

different suppliers, storage of biomass, electricity production and ash management, is

considered in this research. Figure 3-1 shows the supply chain components of the considered

power plant.

Figure 3-1: Schematic of supply chain configuration of a forest biomass power plant

Wood Residues

Bark

Sawdust

Shavings

Roadside logging debris

Storage Production

Ash

Electricity

Supplier 1

Supplier n

......

44

Details of each component are explained below.

Raw materials: The forest biomass can be supplied from forest residues, by-products of forest

product mills, or fast growing crops grown specifically for energy purposes (Demirbaş, 2001).

Forest residues include branches and tops left in the harvest areas after the logs have been

transported to wood manufacturing facilities, as well as small diameter and infected trees not

suitable for lumber production. By-products of forest product mills include wood chips,

sawdust, bark and shavings (Demirbaş, 2001). Poplar and willow are examples of fast growing

tree-crops grown specifically for energy purposes (Rockwood et al., 2004).

In addition to the long term availability of biomass, its quality is an important factor in

economic feasibility of bioenergy projects and the amount of energy generated from it. The

quality of biomass depends on a variety of factors such as Higher Heating Value (HHV),

moisture content, physical, chemical and thermal properties. HHV is the amount of heat

released from complete combustion of dry biomass under standard conditions. Different types

of biomass (e.g. bark, sawdust, shavings, etc.) and different species have different HHV’s. The

moisture content affects the biomass heat content since energy has to be used to evaporate

water at the beginning of the combustion process (Saidur et al., 2011). Density, porosity, size

and shape of biomass are other important physical properties of biomass that impact its

utilization as fuel (Saidur et al., 2011). Different energy conversion technologies require

different particle size ranges. For combustion, the particle size of biomass should be between

0.6 cm (Demirbaş, 2005) and 10 cm (Personal communication with power plant managers).

Size and shape can be modified and improved through pre-processing operations such as

chipping which can take place in the forest, a sort yard or at the power plant. Bark, sawdust and

shavings usually need minor screening and chipping, while larger sized raw materials, such as

45

roadside logging debris, need to be chipped before they can be used in direct combustion.

Biomass chemical properties include the ratio of chemical elements, such as carbon, hydrogen,

oxygen, nitrogen, in biomass and its structural components, which is the amount of cellulose,

lignin and hemicelluloses. These properties are different in different species and impact the

HHV of the biomass as mentioned earlier. However, the impact of chemical composition on

combustion is less significant compared to moisture content (Demirbaş, 2007). Important

thermal properties are specific heat, thermal conductivity, and emissivity and vary with the

moisture content and temperature (Saidur et al. (2011), Lehtikangas (2001), Demirbaş (2003),

Demirbaş (2005)).

Transportation: Different transportation modes, e.g. trucks, railcars, vessels and barge, can be

used for transporting biomass to the power plant. However, forest biomass power plants are

usually supplied from local suppliers and therefore road transport is more likely to be used

among other possible methods. As forest biomass density is relatively low (400 and 900 kg/m3

(Demirbaş, 2001)), its transportation to the power plant requires a large number of trucks

which increases the delivered cost of biomass and the complexity of its logistics system.

Biomass can be transported directly to the power plants, or stored at a satellite storage facility

and used later. The transportation cost of biomass depends on the power plant size, raw

material availability, average transportation distance, biomass density, carrying capacity, and

the travelling speed. Transportation and handling costs usually represent a significant

proportion of the total biomass delivered cost (as high as 50% in some cases (Allen et al.,

1998)).

Storage: Storage is an important issue in a forest biomass supply chain. The storage site can be

located either in the forest, at the power station or at an intermediate point. Usually, when

46

forest biomass is kept in a pile, it generates internal heat over time which is the result of

respiration of living cells in wood (Fuller, 1985). The internal heat generated during storage

makes the biomass more homogenous and warmer which makes it burns more easily. Thus, to

improve the quality of biomass, it is better to keep it in storage for a period of time (1-2

months) (Personal communication with power plant managers) and not to let the amount of

biomass in storage drop below a certain level. The storage level has to be kept within certain

limits, which depend on the size of power plant. However, if the storage amount is higher than

a certain level, its handling cost increases incrementally since there is a need for an extra

operator and material handling equipment. The risk of fire and biomass deterioration are also

higher when it is kept in large piles (Fuller, 1985).

Combustion: Direct combustion is a way to convert the energy embedded in biomass and

derived from the sun and stored in biomass through photosynthesis into other forms of energy,

mainly heat and then electricity (Demirbaş, 2001). Combustion is defined as “a series of

chemical reactions resulting in carbon and hydrogen deoxidization” (Demirbaş, 2005).

Biomass elements, moisture content, and air are critical components of wood combustion. The

products of these reactions include CO, hydrocarbons, oxides of nitrogen, sulfur and inorganic

species such as the alkali chlorides, sulfates, carbonates and silicates (Saidur et al., 2011).

When wood is combusted, mass losses occur which is related to the combustion temperature.

This can be depicted in a plot, called the burning profile, which shows the rate of weight loss

against temperature. The first peak in the burning profile is related to the release of moisture

content and a small amount of other absorbed gases. Then, as the temperature increases to

175°-225°, other volatiles chemicals start to be released and ignite. Then, the rate of mass

losses increases drastically up to temperatures between 325° and 425°, when the mass losses

47

start to decrease. This decline will continue and eventually the weight will be almost constant.

These temperatures are different for different types of biomass (Saidur et al., 2011).

Production: The electricity production process depends on the technology used and the layout

of the power plant (Rentizelas et al., 2009b). There are different biomass combustion

technologies for energy generation such as fixed bed combustion, fluidized bed combustion

and pulverized bed combustion (Saidur et al., 2011). The scale of forest biomass energy

conversion plants can vary from very small scale (for domestic heating) up to a scale in the

range of 100 MWe1. The main restriction on the power plant size is the availability of the local

feedstock, which makes it difficult to have biomass power plants larger than 25 MW. If

dedicated feedstock supplies are available, larger power plants can be built producing 50–75

MW. The net electrical efficiency depends on the scale of the power plant and ranges from

20% to 40%. Small sized biomass power plants have low efficiencies, while the large sized

biomass power plants can be as efficient as fossil fuel systems, however, access to high volume

of biomass throughout the year and high cost of transporting low density material are the issues

with these types of plants (Demirbaş, 2001).

In this study, the overall configuration of the forest biomass power plant is based on a

conventional power cycle used in typical thermal utility plants. The power plant uses forest

biomass which is transferred by conveyors to a boiler where it is burned to generate heat for

steam production. The steam is then transferred into a turbine, where the thermal energy is

converted to electrical power. The exhaust steam is converted to water by a condenser and used

in the system again. The water is re-used for almost seven cycles and then it is discharged to

the sewer network. Other equipment pieces might include a high voltage step up transformer, a

1 Megawatt Electrical

48

solid fuel handling system, an ash removal/handling system and other steam cycle auxiliary

pieces of equipment, multiple cyclones and an electrostatic precipitator.

Ash management: Ash management is an important task in a direct combustion power plant.

Generally, wood ash properties are related to several factors such as: species of tree or shrub,

part of the tree or shrub (bark, wood, and leaves), type of waste (wood, pulp, or paper residue),

combination with other fuel sources, type and quality of soil, weather conditions and

combustion process (Saidur et al., 2011). Forest biomass ash generally contains calcium (Ca)

and potassium (K) (Saidur et al., 2011). Two kinds of ash are generated in the production

process: one is derived from soil and rock contamination (called bottom ash) and the other one

is derived from the minerals in the foliage or wood (called fly ash). Ash disposal is a challenge

for most of the power plants and has economic, environmental and social costs for the power

plant. High ash content biomass is less desirable as a fuel (Demirbaş, 2005). For example,

sawdust has lower ash content than bark and logging residues (Lehtikangas, 2001).

In order to design and manage an efficient supply chain for biomass power plants, all of the

above mentioned components of the supply chain should be considered. It is important that the

power plant receives the required amount of biomass at the right time with a competitive price

to meet the electricity demand and maximize the profit. This can be determined by an

optimization model. An optimum configuration of all processes in the supply chain can help

decision makers run a better operation. The optimality is usually related to cost/profit, however

other factors, such as fuel consumption, greenhouse gas emissions and customer satisfaction,

are also important in the supply chain management. Usually there is a trade-off between these

different objectives, which adds its own challenges to the model.

49

3.3 The optimization model

The model presented here considers multiple procurement sources, several storage options,

different types of forest fuel, and several time periods. The overall objective of the

optimization model developed in this study is to provide estimates of the amount of biomass to

be purchased, stored and consumed in each month during a one-year planning horizon. Here,

the main component of the model, decision variables, constraints and the objective function are

defined. Table 3-1 shows the notation and definition of different sets and decision variables

used in the developed model.

Table 3-1: List of indices and decision variables used in the optimization model

Indices

Product type p∈{Bark, Sawdust, Shavings, Roadside Logging Debris (RLD)}

Supplier s ∈ { Supplier1, ..., Supplier8}

Time steps t ∈ {Jan, Feb, ..., Dec}

Decision Variables

Average energy value of the mix of biomass in storage in month t (MW/green tonnes)

Amount of mixed biomass consumed in month t to produce electricity (green tonnes)

Amount of electricity generated in month t (MWh)

Amount of biomass purchased from supplier s in month t (green tonnes)

Amount of mixed biomass stored in month t (green tonnes)

Binary variable, 1 if storage is higher than the storage upper limit (SUL) in month t

Binary variable, 1 if storage is less than the storage lower limit (SLL) in month t

Table 3-2 shows the notation and definition of different sets and decision variables and

parameters used in the developed model.

50

Table 3-2: List of parameters used in the optimization model

Model Parameters

Average ash content of mix of biomass (%)

Cost of ash handling ($/green tonnes)

Unit cost of biomass type p produced at supplier s ($/green tonnes)

Incremental cost of chemical used for power production ($/MWh)

Electricity demand from client in month t (MWh)

ys em’s overall efficie cy (%)

Electricity price in month t ($/MWh)

Energy value of biomass type p purchased from supplier s in month t (MWh/green

tonnes)

Higher heating value of biomass type p purchased from supplier s in month t (MWh/dry

tonnes)

Maximum biomass available from supplier s in month t (green tonnes)

Maximum absolute storage capacity (green tonnes)

Moisture content of biomass type p purchased from supplier in month t (%)

Penalty cost if storage is above the storage upper limit ($)

The percentage of reduction in biomass quality if storage level is below SLL

The ratio of biomass type p produced at supplier s in month t (%)

Incremental cost of sewer used for power production ($/MWh)

Storage lower limit. If the storage level is below this level, the biomass quality decreases

since there is not enough time for it to be mixed and produce internal heat (green tonnes)

Storage upper limit (green tonnes)

Target storage for the last month (green tonnes)

Biomass transportation cost for supplier s in month ($/green tonnes)

Incremental cost of water used for power production ($/MWh)

The objective function is to maximize the profit (Equation 3.1).

(3.1)

51

Revenue from selling electricity to customer(s) is calculated by multiplying the electricity unit

price by the amount of electricity produced by the power plant (Equation 3.2).

∑ (3.2)

Biomass procurement costs include the cost of purchasing biomass and its transportation costs

(Equation 3.3). The biomass purchase cost is calculated by multiplying the biomass price by

the ratio of each biomass type produced in each supplier and the amount of biomass purchased.

The transportation costs depend on the distance between suppliers and the power plant, and the

amount of purchased biomass. It is assumed that no chipping is done in this power plant. If this

is not the case, the model can easily be modified to include that.

∑ ∑ (3.3)

Ash handling cost is equal to the average ash content of all biomass multiplied by the amount

of biomass consumed for power production multiplied by the cost of handling ash (Equation

3.4). It should be noted that the average ash content is considered in this equation which can be

substituted by more detailed parameter such as ash content of each type of biomass if enough

data were available. Usually different types of wood are burnt together; therefore, the ash

content of the mix of biomass is available.

∑ (3.4)

The penalty storage cost is calculated by multiplying a fixed penalty cost by a binary variable if

the decision variable related to storage is more than a certain level (SUL) (Equation ‎3.5).

∑ (3.5)

52

The production cost contains water, sewer and chemical costs multiplied by the amount of

power produced by the power plant (Equation 3.6).

∑ (3.6)

Constraints of the model are listed below:

In month t, biomass purchased from supplier s has to be less than or equal to the maximum

biomass produced by that supplier.

[For all s, t] (3.7)

Storage level in each month has to be less than or equal to the absolute maximum storage

levels.

[For all t] (3.8)

Equation 3.9 implies the storage in the final month of the year has to be equal to a target

storage level. This constraint guarantees a desired initial storage condition for the next year that

can be set by managers.

(3.9)

Mass balance equation is considered in Equation ‎3.10, which indicates that storage in month

is equal to storage in month (t-1) plus the total biomass purchased from all suppliers minus

biomass consumption in month t.

∑ [For all t] (3.10)

53

Equation 3.11 shows that the power generated by the power plant in each month has to be

equal o he cus omer’s mo hly dema d.

[For all t] (3.11)

Equation 3.12 relates the amount of electricity produced in each month to the energy value of

biomass utilized in that month, efficiency of the system and the reduction in biomass quality if

the storage level is less than the SLL (1-QRF×Zt).

[For all t] (3.12)

The average energy value of biomass in storage in month t is calculated based on the weighted

average of energy values of biomass purchased from suppliers in that month and the average

energy value of stored biomass in month (t-1) as shown in Equation 3.13.

∑ ∑

[For all t] (3.13)

AveEVt is a decision variable since it is calculated based on variables Fs,t and St-1. EVs,p,t is

calculated based on higher heating value HHVs,p,t and the corresponding moisture content

MCs,p,t of biomass as implied in Equation ‎3.14 (Bowyer et al., 2007).

[For all s, p, t] (3.14)

And finally, all the continuous variables are non-negative as shown in Equation 3.15.

[for all t] (3.15)

54

When the model is formulated, additional constraints have to be added for definition of binary

variables Yt and Zt. Yt is defined as:

{

(3.16)

The following constraints (Equations 3.17 and 3.18) are equivalent to the above definition and

replaced it in the model formulation, where M is a sufficiently large number.

[For all t] (3.17)

[For all t] (3.18)

The same constraints have to be added to the model for Zt. Zt is defined as:

{

(3.19)

It can be seen that the model is a nonlinear mixed integer programming (MINLP) model since

Equations 3.12 and 3.13 contain non-linear terms and the model contains both continuous and

binary variables.

MINLP is a combination of two theoretically difficult to solve categories of problems

particularly for large scale problems: mixed integer programs (MIP) and nonlinear programs

(NLP). Therefore, it is not straightforward to solve MINLP problems and get precise results

(Bussieck & Pruessner, 2003). However, several methods have been developed and improved

in the past few years which make it possible to solve MINLP problems more precisely under

special circumstances (convexity, etc.), i.e. terminate with a guaranteed optimal solution or

55

prove that no such solution exists (Bonami & Gonçalves, 2012). The solution methods include

branch-and-bound method (Dakin, 1965), Benders decomposition (Geoffrion, 1972) and Outer

Approximation (OA) algorithm (Duran & Grossmann, 1986).

In this study, the Outer Approximation (OA) algorithm was used since it was provided in the

AIMMS software package (AIMMS. Paragon Decision Technology). The idea in this

algorithm is to switch between solving the nonlinear programming sub-problems and the

relaxed versions of a mixed integer linear master program for a finite sequence. The main

assumption in the Outer Approximation algorithm is the convexity of the nonlinear sub-model.

3.4 Case study

The above model was applied to a case study located in Canada to help maximize profit and

manage supply chain efficiently. The supply chain is the same as what was described in section

3.3. Most of the biomass used by the power plant was supplied from residues from local

sawmills at a low cost. These inexpensive sources of raw material started to decline in recent

years due to the economic downturn that resulted in the closure of some of the mills in the area,

and the competition for biomass from pellet mills. Therefore, biomass prices have increased

and the power plant was forced to use other sources of biomass such as logging debris.

Biomass from different sources has different quality and cost which affect the costs and

efficiency of the operation.

Table 3-3 shows some of the characteristics of the power plant.

56

Table 3-3: Characteristics of the case study

Number of suppliers with a fixed (long term) contract 4

Number of suppliers without a fixed contract More than 50 suppliers

Average ash content 8%

Average moisture content (MC) 32.4% (range: 10.2-46.7)

Average higher heating value (HHV) 5.53 MWh/tonnes (range: 4.06-

5.89)

8565 BTU/lb (range: 6275-9110)

Efficiency 30%

Range of available biomass from suppliers with a fixed contract 0-12,861 (green tonnes per month)

Range of available biomass from suppliers without a fixed

contract

0-18,900 (green tonnes per month)

Some of the specifications of the case study supply chain are presented below.

Raw material: The feedstock used in the power plant includes bark, sawdust, shavings and

roadside logging debris (RLD):

All types of biomass are transported to the power plant by trucks and then kept in storage and

mixed together before combustion. There is a long list of possible suppliers for the power plant

with different contracts, terms, conditions and prices. For some suppliers, if they are in the

operation, the power plant has to buy their biomass. These mills are located close to the power

plant and their biomass is relatively inexpensive because of the short transportation distance

and long term contracts. However, if these mills decide not to operate, there will be a raw

material shortage for the power plant. Therefore, the amount of biomass that can be purchased

from these suppliers is not exactly known in advance. There are other suppliers which can be

considered as ad-hoc suppliers which have no contract with the power plant and usually

57

provide more expensive biomass. Therefore, Equation 3.7 can be broken into two sets of

constraints:

[For all s in suppliers with a fixed contract, t] (3.20)

[For all s in suppliers without a fixed contract, t] (3.21)

Storage: The capacity of the storage for biomass is limited and known. Storing more than a

certain level of biomass forces the power plant to hire an additional operator and use additional

material handling equipment. In addition, there is another upper storage limit, above which

additional operator and additional handling material equipment are needed, and the risk of fire

in the storage facility increases. Therefore, in the model if storage is more than the first upper

level in month t, a binary variable called Xt becomes 1 and if it is more than the second upper

storage level, another binary variable called Yt becomes also 1. Then, the storage penalty cost

will be:

∑ (3.22)

If the level of the storage decreases below a certain level called the minimum storage level, the

quality of biomass is decreased by a certain amount (6% in this case) and there is a risk of

biomass shortage (Equation 3.12).

Demand: The power plant has a long term contract with a customer to provide a fixed amount

of electricity per year, called the firm load. The rest of the production, named the surplus load,

can be produced and sold to the same customer whenever it is profitable. The firm load demand

has to be met at all times, while the power plant has the option to not produce the surplus load.

58

There is also a total amount for the firm load demand in a year. Usually, the power plant

decides whether or not to produce the surplus load in the beginning of the year and informs the

customer about its decision. If the power plant decides to produce the surplus load, the first

fixed amount of production in each hour is considered as the firm load and the rest is the

surplus load. When the production covers the firm demand, the rest of the production is

considered as the surplus demand. If the power plant decides to not produce the surplus load,

the total production in each hour is considered as the firm load and the power plant will be shut

down for the rest of the year after meeting the total firm demand.

To model different demand types, several parameters and decision variables have to be added

explained in Table 3-4.

Table 3-4: Variables and parameters of the case study

Symbol Type Definition

Continuous Variable Total electricity generated in each month (MWh)

Binary Variable 1 if surplus electricity is produced in a year, 0 otherwise

Parameter Energy price for firm load ($/MWh)

Parameter Total firm demand in one year (MWh)

Parameter Amount of firm electricity generated in month t if the surplus is

not being produced (MWh)

Parameter Total surplus demand in one year (MWh)

Parameter Energy price for surplus load ($/MWh)

Parameter Number of working hours in month t

Then, the following constraint (Equation 3.23) needs to be added to the model. This constraint

implies that the electricity produced in each month has two forms. If the surplus load is being

produced, it is equal to the total firm and surplus demand in one year distributed monthly based

on the ratio of working hours of that month to the total working hours in one year. If the

59

surplus load is not being produced, the electricity demand in each month is equal to the firm

load generated in each month.

∑ (3.23)

The objective function also needs to be modified since the revenue from selling electricity to

the customer is equal to the firm load demand multiplied by the power price for the firm load

plus the surplus load demand multiplied by the surplus energy price multiplied by the binary

variable if the surplus power is being produced.

(3.24)

The optimization model for this case study includes Equations 3.8, 3.9, 3.10, 3.11, 3.12, 3.13,

3.15, 3.17, 3.18, 3.20, 3.21 and 3.23 (and corresponding constraints for definition of Zt) and the

objective function is to maximize (3.24 – 3.3 – 3.4 – 3.6 – 3.22).

The MINLP model was solved using the AIMMS software (AIMMS. Paragon Decision

Technology) and the Outer Approximation algorithm. The model had 260 variables (73 binary

variables and 187 continuous variables) and 333 constraints. It took less than a minute to solve

the problem using a 2.80 GHz CPU. The model was implemented with real data and validated

by the power plant managers. It is notable that this model can be used for monthly decision

making. It means that the model can be used at the beginning of the planning horizon (Jan) to

indicate the optimal decision variables. However, if some parameters change later, the model

can be updated using the new information, the previous month’s variables have to be fixed

based on the real values and the model can be run again for the rest of the months.

60

3.5 Results

Table 3-5 shows the results of the optimization model for the power plant using the 2011 data.

The profit from the optimization model is higher than the actual profit of the power plant in

2011. The actual total biomass procurement cost of the power plant in 2011 was $11M which

was 15% higher than the optimum total biomass procurement cost resulted from the model.

Table 3-5 also shows different costs components of the power plant. The biomass purchase cost

accounts for more than 63% of the total cost. Transportation cost encompass 33% of the total

cost which is relatively low compared to other studies (Allen et al., 1998) due to the power

pla ’s access o biomass from local suppliers.

Table 3-5: Results of cost, revenue and profit for the optimization model (in $M)

Biomass

Purchase

Cost

Transportation

Cost

Ash

Handling

Cost

Firm

Revenue

Surplus

Revenue

Total

Profit

Model

results

6.08 3.18 0.32 21.5 4.72 15.03

Figure 3-2 shows the firm and surplus production profile in different months. The production in

May is lower than that in other months due to scheduled maintenance.

Figure 3-3 shows the optimum decision variables for the amount of biomass purchased, stored

and consumed in each month based on the 2011 data. The initial storage level was 72,500

green tonnes in 2011. The optimum storage profile declines until April, then increases in May

due to the maintenance shut down in the power plant.

61

Figure 3-2: The amount of firm and surplus electricity production in each month

Figure 3-3: Optimum amount of biomass stored, purchased and consumed in each month

based on the 2011 data

3.5.1 Scenario analysis

In addition to the base case scenario considered in the previous section, three other scenarios

are analyzed here. These scenarios are important for the power plant managers.

Scenario 1: One of the suppliers with a fixed contract shuts down its operation which is

possible due to the economic situation (pessimistic scenario).

0

10,000

20,000

30,000

40,000

50,000

60,000

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

MW

h

Firm

Surplus

0

10,000

20,000

30,000

40,000

50,000

60,000

70,000

80,000

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Gre

en T

on

nes

Storage

Purchase

Consumption

62

Scenario 2: One of the previous suppliers with a fixed contract which was closed in the

past couple of years resumes its operation (optimistic scenario).

Scenario 3: The power plant makes an investment in a new piece of equipment for ash

recovery. Having a new ash recovery system with a capital cost of $2M, the ash

generated in the power plant is screened and the unburnt parts are separated and sent

back to the operation process. It has been estimated that this system will collect 20% of

the ash and use it as a biomass source. It is assumed that the energy value of the unburnt

part of the ash is the same as the average energy value of the biomass consumed in that

month. Therefore, it has the potential benefit of providing more biomass and less ash

for the power plant which reduces the ash disposal costs.

Table 3-6 shows the total profit and the total purchase cost of biomass for different scenarios.

The results show that Scenario 1 has a lower profit (24.48%) than the base case scenario.

Scenario 2 has a higher profit than the base case scenario since the power plant has access to

cheaper biomass from a supplier with a fixed contract. The amount of increase in profit is

$3.27M (21.76%) which is the result of reduction in the biomass purchase cost. Scenario 3,

investment in a new ash recovery system, has a higher profit than the base case scenario

($160,000, 1.06%). This means that investing in a new ash recovery system has economic and

environmental benefits for the power plant. The increase in profit is partly from the reduction

in the biomass purchase cost (3.46%), and partly from reduction in the ash handling cost

(20.75%). The capital cost of the ash recovery equipment was converted to the yearly cost

using a 2% interest rate and 10 years’ service life, and then the yearly cost was subtracted from

the profit.

63

Table 3-6: Total profit and biomass procurement cost for four different scenarios

Base Case Scenario Scenario 1 Scenario 2 Scenario 3

Total Profit ($M) 15.03 11.35 18.30 15.19

Biomass Procurement Cost ($M) 9.26 12.94 6.00 8.94

Profit difference compared to the base

case scenario

0 -24.48% +21.76% +1.06%

3.5.2 Sensitivity analysis

To evaluate the effect of variations in model parameters on the profit, sensitivity analysis was

performed. The results of variation in profit are depicted in Figure 3-4. It can be seen that the

impact of variation in some of the parameters, such as electricity price, higher heating value

and maximum available biomass from suppliers with a fixed contract have a large impact on

the profit. When these parameters increase by 20%, the increase in profit is between 20-35%.

On the other hand, when they are reduced by 20%, the loss is 23-45%. Other parameters are:

moisture content, biomass cost and transportation costs which have the opposite impact on the

profit. Increasing these parameters by 20%, decreases the profit by about 4-15%. The effects

of variation in other parameters on the overall profit are low and can be considered negligible.

The sensitivity of the results to changes in initial storage level was also determined. The initial

storage level was changed from 0 to 118,000 green tonnes, while the target storage level was

set to the initial storage level. As shown in Figure 3-5, an initial storage level of 109,000 green

tonnes generates the highest profit.

64

Figure 3-4: Variations in profit with 20% change in different parameters

Figure 3-5: Variations in profit for different initial storage levels

-50%

-40%

-30%

-20%

-10%

0%

10%

20%

30%

40%

-30% -20% -10% 0% 10% 20% 30%

Pro

fit

Dif

fere

nce

Change in Parameter

Max biomass from supplier

with fixed contract

Max biomass from supplier

without fixed contract

Ash content

Electricity price

Quality reduction factor

HHV

MC

Transportation cost

Wood cost

0

2

4

6

8

10

12

14

16

0 20,000 40,000 60,000 80,000 100,000 120,000 140,000

Pro

fit

($M

)

Initial Storage (Green Tonnes)

65

3.6 Discussion and conclusions

In this chapter, a mathematical optimization model was developed to determine the amount of

biomass to purchase from each supplier in each month, the amount of mixed biomass to burn

and store in each month, and whether or not to produce extra electricity to maximize the total

profit. The developed model was a mixed integer non-linear model, which was solved by the

Outer Approximation algorithm using the AIMMS software package. The results for the

specific case study presented here indicated that the profit from the optimization model was

higher than that of the actual profit generated by the power plant in 2011 when he compa y’s

managers made decisions based on their experience. For instance, the optimum biomass

procurement cost was $9.26M, which was 15% lower than the actual purchase and

transportation costs of the power plant in 2011 ($11M). The model provided the optimum

profile for biomass storage, purchase and consumption in each month to achieve maximum

profit over a year. The model could be re-optimized and updated if input data varied from what

had been used here.

The power plant situation in 2011 was considered as the base case scenario and three other

scenarios were evaluated in this study. The first one was a pessimistic scenario in which the

production shutdown of a supplier with a fixed contract was investigated. The second scenario,

an optimistic scenario, was related to the resumption of production at one of the previous

suppliers with a fixed contract. And finally, the third scenario investigated the investment in a

new ash recovery system. This system collects and reuses the unburnt parts of the produced

ash. Based on the results, the profit of the first scenario was 24% lower than that of the base

case scenario, while the profit of second and third scenarios were 22% and 1% higher than that

66

of the base case scenario, respectively. The reason for having lower profit in the first scenario

is that the power plant had no longer access to one of its supplier (located in the area and

providing inexpensive biomass). Therefore, it needed to purchase more expensive biomass

from ad-hoc suppliers. For the second scenario, it was other way around, meaning that an extra

inexpensive source of biomass with a fixed long term contract was available for the power

plant. Therefore, it resulted in having a higher profit than the base case scenario. In the third

scenario, the ash handling cost reduced by 21%. Therefore, investing in a new ash recovery

system could have economic benefit as well as environmental advantage for the power plant.

This chapter addressed the first, second and part of the third objectives mentioned in 1.2. It

provided an integrated framework for optimizing the supply chain of a forest biomass power

plant. The developed model was then applied to a real case study. The model and its results

were validated by the power plant managers. It was a mixed integer non-linear programming

model with multiple time steps and captured seasonal variations in parameters particularity in

biomass availability and quality. The quality of the mix of material in the storage was

calculated, and the impact of any change in the quality of each biomass type from each supplier

in each month on the solution could be examined. Therefore, it included the issue of quality in

the modeling and provided optimum decisions on biomass purchase, storage and consumption

in order to meet electricity demand in each month.

The impact of changes in input parameters on the profit was determined using sensitivity

analysis. Parameters that had higher impact on the model solution were also identified.

However, uniform changes in input parameters were considered in sensitivity analysis. In other

words, it was assumed that uncertainties in all parameters had the same nature and they could

67

have any value within a defined range with the same probability. This is not a realistic

assumption since there are different sources of uncertainty in the system and some of the

uncertain parameters are correlated with one another. Moreover, not all the possible

realizations of uncertain parameters occur with the same probability. To capture these issues,

Monte Carlo simulation is used to model the uncertainties in input parameters more

realistically, as can be seen in the next chapter.

68

Chapter 4 Monte Carlo simulation

4.1 Synopsis

In this chapter, Monte Carlo simulation is used in combination with the optimization model

developed in Chapter 3 to evaluate the impact of uncertainty in different parameters including

biomass quality, availability and cost and electricity prices based on historical data. Different

types of uncertain parameters are incorporated in the model. Historical data on biomass quality

(moisture content and higher heating value) obtained from the power plant is used to generate

fitted probability distributions. The Monte Carlo simulation and the optimization model

provide a probability distribution for possible optimal solution that can be used for

u ders a di g he sys em’s behaviour i he prese ce of u cer ai y, ide ifyi g he risks a d

opportunities in the supply chain, and making informed decisions.

4.2 Uncertainty and Monte Carlo simulation

Vose (2008) divided the uncertainty based on its origin into two categories: 1) uncertainty

resulted from incomplete understanding of a system due to its complexity, and 2) uncertainty

related to the random nature of a parameter, which was called “variabili y” by he au hor.

Approaches to be used for dealing with uncertainties in the model and actions to be taken in

response to them vary given the source of uncertainty in stochastic parameters. While

collecting more information is effective in expanding the knowledge about a random event,

understanding a system or trying to modify it might be more useful in managing uncertainties

resulting from system complexity. Thus, it is important to evaluate the impacts of these two

types of uncertainty separately to realise the impact of each on the solution. This helps decision

69

makers identify the proper steps to take to mitigate the risks and take advantage of possible

opportunities associated with variations in the system (Vose, 2008).

The results of the sensitivity analysis performed and presented in Chapter 3 showed that

parameters with higher impact on the optimum solution were the monthly available biomass

from each supplier (MaxFs,t), biomass MC and HHV, biomass cost and electricity prices.

Availability and cost of biomass, as well as electricity price can be grouped together because

they are affected by economic and market conditions. Moreover, due to interdependency

between different forest sectors and diverge production structure of industries, variations in one

part of this supply chain can propagate into the other parts. Therefore, availability and cost of

biomass are impacted by variations in other forest product sectors such as lumber and pulp and

paper sectors. On the other hand, uncertainty in biomass quality is related to the heterogeneous

nature of wood resulting from the variations in its physical and chemical characteristics from

different species and different parts of a tree (Bowyer et al., 2012). Biomass quality is also

affected by external factors, such as growth conditions and climate, and can change during

storage, production, and transportation. It impacts collection, transportation, handling, storage

and the efficiency of conversion technology (Rentizelas et al., 2009b).

AIMMS software package (AIMMS. Paragon Decision Technology) was used to model Monte

Carlo simulation along with the optimization model. Monte Carlo simulation is first performed

for each group of the uncertain parameters separately to study their specific impact on the

power pla ’s profi . The , a overall u cer ai y a alysis is co duc ed by allowi g he i pu

parameters to change over their respective probability distributions/ ranges simultaneously. The

70

probability distributions of the outputs, including total profit, monthly storage level, biomass

consumption and biomass purchase, are recorded and analyzed.

4.2.1 Uncertainty in biomass quality

The feedstock for the power plant is provided by different suppliers. Suppliers either have a

fixed contract or no contact with the power plant. Suppliers 1 to 4, which include 3 sawmills

(Suppliers 1, 2 and 4) and a sawmill and a plywood mill (Supplier 3), have fixed contracts with

the power plant. Suppliers 5 to 8 do not have fixed contracts with the power plan. Suppliers

with no contract include a sawmill (Supplier 5) which produces bark and sawdust, a harvesting

company (Supplier 6) that only provides roadside logging debris (RLD), a number of small

suppliers that occasionally provide biomass to the power plant (Supplier 7) and some suppliers

that are considered as ad-hoc suppliers and occasionally sell biomass to the power plant

(Supplier 8). The biomass cost provided by Suppliers 7 and 8 are different from each other and

the types and the amount of biomass from them are not known. Table 4-1 shows the product

mix (based on weight percentage) and type of contract for the suppliers.

Table 4-1: Product type of suppliers and their contract type

Product Mix (%) Type of contract

Supplier 1 Bark (29%), Sawdust (71%) Fixed

Supplier 2 Bark (100%) Fixed

Supplier 3 Bark (41%), Sawdust (29%), Shavings (30%) Fixed

Supplier 4 Bark (63%), Sawdust (37%) Fixed

Supplier 5 Bark (70%), Sawdust (30%) Flexible

Supplier 6 Roadside logging debris (100%) Flexible

Supplier 7 Uncertain Flexible

Supplier 8 Uncertain Flexible

71

Based on the historical data received from the power plant, the best fitted probability

distributions are determined for MC and HHV of biomass using the Stat::Fit software.

Moisture Content

Suppliers 1 to 5 produce bark, sawdust and shavings. The operation process of these suppliers

takes place in closed areas year-round. Therefore, it is assumed that the seasonality does not

impact the distribution of MC of their products, but the MC of different products varies. In

contrast, RLD provided by Supplier 6, is collected and stored at roadside before being

delivered to the power plant and is affected by seasonality. For Suppliers 7 and 8, the product

types are unknown, so, only the seasonality effect is considered for their biomass MC.

Table 4-2 shows average and standard deviation as well as the fitted probability distribution

and its parameters for MC of bark, sawdust and shavings from Suppliers 1 to 5.

Table 4-2: Average and standard deviation of bark, sawdust and shavings MC for Suppliers 1

to 5

Average MC

(%)

Standard

Deviation (%)

Fitted Probability Distributions

Bark 30.0 8.7 Weibull (1.8,13,18.6)

Sawdust 34.8 8.4 Beta (4.2,5.9,16.0,61.1)

Shavings 11.5 11.3 Triangular (0.01,2.3,45.2)

Figure 4-1 illustrates the histogram and probability distribution fitted to the MC of bark,

sawdust and shavings.

72

Figure 4-1: Histogram and probability distribution of MC of bark (a), sawdust (b), shavings (c)

Table 4-3 presents the average and standard deviation as well as the fitted probability

distribution and its parameters for biomass in different months for suppliers 6 (RLD only). It

should be noted that harvesting operation is not done in the area in May so there is no RLD

available for that month.

Figure 4-2 shows the histogram and probability distribution fitted to the MC of RLD in

different months.

a b

c

73

Table 4-3: Average and standard deviation of biomass MC for Supplier 6

Supplier 6 (RLD) Fitted Probability Distributions

Ave MC (%) SD (%)

January 26.6 5.6 Triangular (0.9,13.8,33.5)

February 36.2 8.4 Gamma (2.8,21,5.41)

March 33.7 9.1 Weibull (2.6,10.0,26.4)

April 45.5 16.6 Beta (0.9,0.8,15.0,69.7)

May - - -

June 25.5 10.7 Triangular (0.3,7.3,51.5)

July 23.7 5.6 Weibull (1.3,15.0,9.32)

August 21.1 4.9 Beta (0.8,0.7,12.0,27.4)

September 24.0 5.6 Triangular (0.5,10.3,38.7)

October 27.1 10.1 Weibull (1.5,12.0,16.8)

November 28.5 6.1 Weibull (1.5,19.0,10.4)

December 35.6 7.6 Triangular (0.1,25.3,55.7)

a b

74

c d

e f

g h

i j

75

Figure 4-2: Histogram and probability distribution for MC of RLD in January (a), February (b),

March (c), April (d), June (e), July (f), August (g), September (h), October (i), November (j),

and December (k)

Higher Heating Value

Table 4-4 shows the average, standard deviation, the fitted probability distribution and its

parameters for HHV of different types of biomass received by the power plant from different

suppliers. It was not possible to fit a distribution for the HHV of bark and shavings because of

the low number of data points, therefore, a range was considered for their variations.

Table 4-4: Average and standard deviation of HHV for different biomass types

HHV Average Standard Deviation Fitted Probability Distributions

Bark 8534 BTU/lb2

5.51 MWh/tonnes

670 BTU/lb

0.43 MWh/tonnes

Uniform (7570.0,9110.0) (BTU/lb)

Uniform (4.89,5.89) (MWh/tonnes)

Sawdust 8537 BTU/lb 5.51

MWh/tonnes

325 BTU/lb

0.21 MWh/tonnes

Uniform (8050.0,8960.0) (BTU/lb)

Uniform (5.20,5.79) (MWh/tonnes)

Shavings 8517 BTU/lb 5.50

MWh/tonnes

254 BTU/lb

0.16 MWh/tonnes

Uniform (8360.0,8810.0) (BTU/lb)

Uniform (5.40,5.69) (MWh/tonnes)

RLD 8392 BTU/lb 5.42

MWh/tonnes

347 BTU/lb

0.22 MWh/tonnes

Weibull (6.1,6960.0,1580.0) (BTU/lb)

Weibull (6.19, 4.41,1.089)

(MWh/tonnes)

2 BTU/lb= 0.0006461MWh/tonnes

k

76

Figure 4-3 show the histogram and probability distribution fitted for HHV of bark and RLD.

Figure 4-3: Histogram and probability distribution of HHV of sawdust (a) and RLD (b)

4.2.2 Uncertainty in biomass availability and cost and electricity price

The amount of by-products generated by he power pla ’s suppliers depends on the production

of their main products, which are mainly used in housing construction. In the past few years,

many sawmills in the area were closed due to the real estate meltdown in the US. Thus, the

power plant faced a shortage of raw material. Conversely, when the market is promising, the

sawmills produce a large amount of biomass that the power plant is obliged to buy. The

situation with a supplier without a fixed contract is less critical since there is no obligation for

the power plant to buy their biomass. However, there still exists uncertainty in the available

biomass that they can provide. Moreover, biomass procurement and transportation costs vary

due to the competition for biomass from other sectors, such as the wood pellets sector, and as a

result of variations in fuel cost. Electricity prices are also uncertain and have to be forecasted

ahead of the time so that the power plant can decide about whether or not to produce surplus

demand in the upcoming year.

a b

77

Uncertainty in electricity prices, and biomass availability and cost are considered together

because they are all related to the economic situation and considered to be correlated. Based on

the historical data, three electricity price scenarios (high, medium and low) with their

corresponding annual available biomass from different suppliers were selected. The annual

available biomass was then distributed over a year based on the monthly production profile of

2011. The Monte Carlo simulation model was run for each scenario with ±20% variation in

monthly biomass availability and ±10% variation in biomass cost.

4.3 Results

The number of iterations for Monte Carlo simulation depends on the desired accuracy and

confidence levels, and is calculated using Equation 4.1 (You et al., 2009):

where N is the number of iterations, Z is the confidence interval of a two tailed normal

distribution, σ is he model’s ou pu s a dard devia io a d εμ is the desired marginal

error. Based on the marginal error of 6%, confidence level of 95% (Z=1.96), and standard

deviation of model results for 100 runs, the required number of iterations was estimated to be

1000.

Table 4-5 shows mean, minimum, maximum and standard deviation of profit and the feasibility

rate obtained from the output of the Monte Carlo simulation-optimization model when MC and

HHV were generated separately from their probability distributions. It can be seen that

incorporating the variability in MC resulted in higher average and standard deviation for the

profit compared to that of HHV. Variability in HHV changes the profit within a $280,000

78

interval with an average of $15.1M, while for MC, the variation in the profit is $700,000 with

an average of $15.3M as shown in Figure 4-4 and Figure 4-5.

Table 4-5: Minimum, maximum, average and standard deviation of profit for considering

uncertainty in different parameters

Uncertain

Parameter

Mean, (Min, Max) of

Profit ($M)

Standard Deviation

($M)

Feasibility

Rate

MC 15.3 (14.1-16.5) 0.35 (2%) 99.9%

HHV 15.1 (14.6-15.6) 0.14 (1%) 99.9%

Figure 4-4: Histogram of profit when MC varies

Figure 4-5: Histogram of profit when HHV varies

0%

5%

10%

15%

20%

25%

30%

14.1

0-1

4.1

9

14.2

0-1

4.2

9

14.3

0-1

4.3

9

14.4

0-1

4.4

9

14.5

0-1

4.5

9

14.6

0-1

4.6

9

14.7

0-1

4.7

9

14.8

0-1

4.8

9

14.9

0-1

4.9

9

15.0

0-1

5.0

9

15.1

0-1

5.1

9

15.2

0-1

5.2

9

15.3

0-1

5.3

9

15.4

0-1

5.4

9

15.5

0-1

5.5

9

15.6

0-1

5.6

9

15.7

0-1

5.7

9

15.8

0-1

5.8

9

15.9

0-1

5.9

9

16.0

0-1

6.0

9

16.1

0-1

6.1

9

16.2

0-1

6.2

9

16.3

0-1

6.3

9

16.4

0-1

6.4

9

Fre

quen

cy

Profit ($M)

0%

5%

10%

15%

20%

25%

30%

14.6

0-1

4.6

9

14.7

0-1

4.7

9

14.8

0-1

4.8

9

14.9

0-1

4.9

9

15.0

0-1

5.0

9

15.1

0-1

5.1

9

15.2

0-1

5.2

9

15.3

0-1

5.3

9

15.4

0-1

5.4

9

15.5

0-1

5.5

9

Fre

quen

cy

Profit ($M)

79

As explained before, three scenarios were considered for electricity price and biomass

availability and cost. The results of the Monte Carlo simulation-optimization model for the

three scenarios are shown in Table 4-6.

Table 4-6: Results of Monte Carlo simulation-optimization model for scenarios of electricity

price and biomass availability and cost

Scenario Surplus Electricity Price

($/MWh)

Mean Profit ($M) SD ($M) Feasibility rate

(%)

Low 35.00 14.4 0.1 100

Average 42.50 15.1 0.5 99.9

High 103.87 17.6 0.2 99.3

Figure 4-6 represents the profit histogram when availability and prices vary together for low,

average and high electricity price scenarios. It can be seen that despite the low standard

deviation of low electricity price scenario, its profit histogram is more bell-shaped, while the

histogram of profit for average and high electricity price scenario are more skewed to higher

profit ranges.

0%

10%

20%

30%

40%

50%

60%

70%

14.3

0-1

4.3

9

14.4

0-1

4.4

9

14.5

0-1

4.5

9

14.6

0-1

4.6

9

14.7

0-1

4.7

9

14.8

0-1

4.8

9

14.9

0-1

4.9

9

15.0

0-1

5.0

9

Fre

quen

cy

Profit ($M) a

0%

10%

20%

30%

40%

50%

60%

70%

13.8

0-1

3.8

9

13.9

0-1

3.9

9

14.0

0-1

4.0

9

14.1

0-1

4.1

9

14.2

0-1

4.2

9

14.3

0-1

4.3

9

14.4

0-1

4.4

9

14.5

0-1

4.5

9

14.6

0-1

4.6

9

14.7

0-1

4.7

9

14.8

0-1

4.8

9

14.9

0-1

4.9

9

15.0

0-1

5.0

9

15.1

0-1

5.1

9

15.2

0-1

5.2

9

15.3

0-1

5.3

9

15.4

0-1

5.4

9

Fre

quen

cy

Profit ($M) b

80

Figure 4-6: Histogram of profit when electricity price and biomass availability and cost vary

for a) low, b) average and c) high scenarios

Table 4-7 provides the probability of having profit within a certain range obtained from the

Monte Carlo simulation-optimization model. The last row represents the results for the case

that uncertainty is included in all parameters. The risk of having profit less than $14M for the

low electricity price scenario is 99%. For the average electricity price scenario, it is more likely

to have profit between $14M and $15M, while in the high electricity price scenario the profit is

always more than $17M. When variation and uncertainty in all parameters are considered, the

risk of having a profit less than $14M is 6%, while there is 32% chance of having a profit

higher than $17M. Most of the time (62%), the compa y’s profi is be wee $14M a d $16M.

Because variations in biomass cost and availability and electricity prices affect the profit of the

power plant significantly and increase the risk of having low profits when electricity prices are

low, the power plant managers should pay special attention to the contract details with

suppliers and forecast of electricity prices.

The outputs of the Monte Carlo simulation-optimization model show that in order to have a

profit higher than $15.5M, the average MC should be less than 28% for bark, 34% for sawdust,

10% for shavings and 25% for RLD. If the power plant desires a profit higher than $15.0M, the

0%

10%

20%

30%

40%

50%

60%

70%

17.3

0-1

7.3

9

17.4

0-1

7.4

9

17.5

0-1

7.5

9

17.6

0-1

7.6

9

Fre

quen

cy

Profit ($M)

c

81

MC level will be 31% for bark, 41% for sawdust, 24% for shavings and 37% for RLD. This

indicates that drying of biomass in order to reduce its moisture content level could increase the

annual profit by $0.5M.

Table 4-7: Probability of having profit within different ranges when considering uncertainty in

different model parameters

Uncertainty in Pr.(Profit<

$14M )

Pr.($14M<Profit

<$15M)

Pr.($15M<Profit

<$16M)

Pr.($16M<Profit

<$17M)

Pr.(Profit

>$17M)

MC 0 0.17 0.81 0.02 0

HHV 0 0.27 0.73 0 0

Availability and

Prices (Low)

0.99 0.1 0 0 0

Availability and

Prices (Average)

0.01 0.99 0 0 0

Availability and

Prices (High)

0 0 0 0 1

All parameters 0.06 0.51 0.11 0 0.32

Table 4-8 represents ranges of annual biomass purchase from suppliers without a fixed

contract, biomass consumption and monthly storage levels from the Monte Carlo simulation-

optimization model. It can be seen that when variability in HHV is included, total consumption

and total biomass purchase from suppliers without a fixed contract have a higher average and a

lower standard deviation compared to the case when variability in MC is considered. It is also

indicated that the maximum monthly storage never goes beyond the storage upper level

(109,000 green tonnes), therefore, the penalty storage cost is not active in any cases. For the

model with uncertainty in availability and prices and also all parameters, there is a risk of

having monthly storage lower than the minimum storage level (45,000 green tonnes), and

consequently lower biomass quality.

82

Table 4-8: Ranges of biomass purchase from suppliers without contract, biomass consumption

and storage levels when considering uncertainty in different parameters (1000 green tonnes)

Uncertainty in Annual

Consumption

Annual purchase from suppliers

without a fixed contract

Monthly storage

Average SD Average SD Average Min Max

MC 483 5 118 5 56 45 89

HHV 488 3 122 3 56 45 94

Availability and

prices (Low)

478 1 76 1 57 38 68

Availability and

prices (Average)

486 1 119 1 51 33 63

Availability and

prices (High)

485 1 140 1 59 45 74

All parameters 489 6 117 25 56 34 102

4.4 Discussion and conclusions

In this chapter, uncertainty is studied in the mixed integer non-linear programming model of a

forest biomass power plant supply chain through developing a Monte Carlo simulation-

optimization model in order to determine the range of variations in profit due to variations in

input parameters.

The results showed a lower standard deviation for profit when variation in higher heating value

(HHV) was included compared to the case when variation in moisture content (MC) was

included. Therefore, it could be concluded that the impact of variations in MC on the optimum

profit was more severe than the variation in HHV. Moreover, variations in MC of biomass

resulted in the highest variation in the annual consumption and annual purchase of biomass

from suppliers without a fixed contract. This was in contrary with the result obtained from

sensitivity analysis which showed that variations in HHV change the profit more than

83

variations in MC. The reason for that was that in sensitivity analysis the HHV and MC changed

within a same range (±20%), while variations in the historical data showed different ranges

with different probability distributions. These results showed that the Monte Carlo simulation

provided more realistic results that might differ from those ones obtained from sensitivity

analysis.

It was concluded that to have a profit of more than $15.0 M, the MC had to be less than 31%

for bark, 41% for sawdust, 24% for shavings and 37% for road logging debris (RLD).

Reducing MC to less than 28% for bark, 34% for sawdust, 10% for shavings and 25% for RLD

would increase the profit to $15.5 M, if all other parameters of the model remained the same.

This information would be helpful for the power plant supply manager when making supply

contracts and decisions on the drying of biomass because the amount of profit earned when

biomass was dried and had lower MC was shown.

The results indicated that when the electricity price was high (103.87 $/MWh), the profit

probability density function was more skewed to higher profit, while when the electricity price

was low (35.00 $/MWh), the profit distribution was more bell-shaped. This means that higher

electricity prices mitigate the impact of variations in cost and availability of biomass. The

results also showed that when all the uncertain parameters fluctuated at the same time, the

profit ranged between $13.4M and $17.9M with an average of $15.5M. Therefore, uncertainty

in the real system would resul i $4.5M varia io i he power pla ’s profi .

Uncertainty in biomass availability and cost and electricity prices, not the variability in

biomass quality (MC and HHV), resulted in the risks of having a profit lower than $14M and a

storage level below the minimum storage level, as well as the opportunity of having a profit

84

higher than $17M. Therefore, to mitigate the risks of having low profit and low storage levels,

it is needed to control uncertainties in the availability and prices.

This chapter addressed objective 3 of the thesis mentioned in 1.2. Historical data on important

uncertain parameters, which were identified previously through sensitivity analysis in 3.5.2

were analysed here. Uncertain parameters with different sources were grouped together and the

Monte Carlo simulation was implemented for each group as well as for all of them at the same

time. The profit distribution was determined when uncertain parameters varied based on their

distribution. Therefore, the output of this model represented the model behaviour when

uncertainty occurred and risks associated with uncertainties.

The Monte Carlo simulation method is a post-optimization approach, which means that the

decisions have to be made before this model is implemented and it only provides possible

changes in the decision variables and the profit in response to uncertainty. Taking uncertainty

into account when modeling the supply chain is a challenging task because the model becomes

complex and sometimes it may become computationally intractable especially if uncertainty is

included in different parameters simultaneously. Therefore, in the next chapter, uncertainty in

only one parameter, biomass availability, is included in the optimization model.

85

Chapter 5 Stochastic programming

5.1 Synopsis

So far, the impact of uncertainty on model solution has been evaluated through sensitivity

analysis and Monte Carlo simulation which are post-optimization approaches. In this chapter,

uncertainty is incorporated in the optimization model during the decision making process. This

is done by considering uncertainty in the monthly available biomass from different suppliers

during the planning horizon. The mixed integer non-linear programming model is first

reformulated into a mixed integer programming model. This enables a direct formulation of a

linear stochastic programming model. A two-stage stochastic programming model is developed

to maximize the expected profit of possible scenarios of available biomass. Two risk measures

are also included in the model to make the results more stable. The trade-off between

maximizing the profit and minimizing the risk is then evaluated.

5.2 The mixed integer programming model of the power plant supply chain

To facilitate incorporating uncertainty in the model, a new mixed integer linear programming

(MILP) model is proposed in this chapter. The main difference between the formulation of

MILP model in this chapter and the one developed in Chapter 3 is in the definition of the

storage and consumption variables. The notation of these variables is presented in Table 5-1.

Table 5-1: Decision variables of the linear programming model

Decision Variables

Amount of biomass stored in month from supplier (green tonnes)

Amount of biomass consumed in month from supplier (green tonnes)

86

The mixed integer linear model is as follows:

Objective function:

∑ (∑ )

∑ ∑ ∑ ∑ (5.1)

Here, Profit equals revenues minus costs as shown in Equation 5.1. Ash handling cost has

changed to ∑t AshHC × AshC × ∑s Cs,t.

The changes in the constraints are as follows:

The total storage level of biomass received and stored from all suppliers in month t has to be

less than or equal to the maximum storage level. Therefore, Equation 3.8 is changed to

Equation 5.2.

∑ [For all t] (5.2)

In the last month of the year, the total storage level of biomass from all suppliers has to be

equal to a target storage level. Therefore, Equation 3.9 is changed to Equation 5.3.

∑ (5.3)

The storage level of biomass received from supplier and stored in month t is equal to the

storage level of biomass received from supplier s and stored in previous month (t-1) plus

biomass purchase minus biomass consumption for the same supplier and month. For the first

time step, the initial storage level for each supplier has to be indicated. Therefore, Equation

3.10 is changed to Equation 5.4.

87

[For all s, t] (5.4)

In the MINLP model shown in Chapter 3, the decisions on purchase and consumption amounts

were not separated for each supplier because the total initial storage level was provided in the

obtained data. The initial energy value of the biomass in the storage was not provided and an

initial value for it in the MINLP was assumed. In the MILP model developed here, the initial

storage level for each supplier at the beginning of the planning horizon has to be known in the

mass balance equation for January (Equation 5.4). Therefore, it is assumed that the share of

biomass i he i i ial s orage level for each supplier is he same as each supplier’s share i he

total annual available biomass for the power plant in 2011. The initial average energy value in

the MILP is then calculated using those ratios.

Equation 3.12 is changed to Equation 5.5 to connect the energy demand in month t to the

biomass consumption variable. Et equals to the total biomass consumption for all suppliers in

month t multiplied by the total ratio of biomass type p for supplier s in month t, the energy

value of biomass type p from supplier s in month t, system efficiency, and the ratio of quality

reduction factor, if the storage level in month t is lower than SLL.

∑ ∑ [For all t] (5.5)

There is no need for variable AveEVt and Equation 3.13. It can be seen that constraint 5.5

contains the multiplication of a binary variable Zt and a continuous variable Cs,t. This non-

linear term can be converted to a linear constraint through replacing the term Zt×∑sCs,t by an

additional continuous variable, i.e. Lt≥0 and the following constraints 5.6, 5.7 and 5.8, where M

is a sufficiently large number.

88

[For all t] (5.6)

∑ [For all t] (5.7)

∑ [For all t] (5.8)

Having the above changes in the constraints and the objective function of the model presented

in Chapter 3 forms a mixed integer linear programming (MILP) model which was solved using

the AIMMS software package and the CPLEX solver.

5.3 The stochastic mixed integer programming model of the power plant supply chain

A two-stage stochastic programming approach is used here to deal with uncertainty in available

monthly biomass and incorporate it into the above multi-period planning model. In this

approach, it is assumed that at the beginning of the planning horizon, the available biomass

from each supplier for the first quarter of a year is known and it is uncertain for the rest of the

year. Therefore, the first stage includes the first three months (each period is a month) of a year

and the next nine months are considered as the second stage. The decision variables, including

the amount of biomass to be purchased from each supplier, the amount of biomass from each

supplier to be stored and the amount of biomass from each supplier to be consumed depend on

the available biomass, hence, these decision variables are considered as the first stage decision

variables in the first three months and the second stage decision variables in the next nine

months. The scenarios are made by varying the monthly available biomass from each supplier

by ±20% of the average scenario together with the expected volume, i.e. 3 different values. It is

assumed that the change in available biomass is stationary for three months since uncertainty in

available biomass is the result of fluctuation in lumber and other wood products market and the

89

market situation is assumed to be constant for that period. Consequently, the total number of

scenarios is 3 multiplied by the number of aggregated periods in the second stage, i.e. 33=27.

The general formulation for a two-stage stochastic programming is shown in the literature (e.g.

Birge & Louveaux (1997)) and is not repeated here. In this study, an implicit formulation of the

stochastic model for the forest biomass power plant supply chain is presented. This means that

there is one set of decision variables and the first stage decision variables are distinguished

from the second stage decision variables using non-anticipatory constraints.

There is a new set and a new parameter used in the developed two-stage stochastic model

defined as below:

Scenarios: List of scenarios (index i) {Scenario1, …, Scenario27}

: Maximum available biomass from supplier s in month t for scenario i.

Table 5-2 represents the definition of the stochastic model decision variables.

The objective function is to maximize the expected profit of all scenarios. To calculate the

expected value of profit, profit of scenario i (Profiti) is multiplied by Pri, the probability of

occurrence of scenario i (∑iPri =1), and the objective function is the sum of expected values as

shown in Equation 5.9.

(5.9)

Here, Profiti equal revenues minus costs for scenario i as shown in Equation 5.10.

90

Table 5-2: Stochastic model decision variables

Amount of biomass purchased from supplier s in month t for scenario i (green tonnes), first

stage decision variable for t=1-3, and second stage decision variable for t=4-12.

Amount of biomass from supplier s stored in month t for scenario i (green tonnes), first stage

decision variable for t=1-3, and second stage decision variable for t=4-12.

Amount of biomass from supplier s consumed in month t for scenario i (green tonnes), first

stage decision variable for t=1-3, and second stage decision variable for t=4-12.

Amount of electricity generated in month t for scenario i (MWh), first stage decision variable

for t=1-3, and second stage decision variable for t=4-12.

Binary variable, 1 if surplus electricity is produced in a year, 0 otherwise for scenario i.

Binary variable, 1 if Ss,t,i is higher than the desired storage level (SDL) in month for scenario

i, first stage decision variable for t=1-3, second stage decision variable for t=4-12.

Binary variable, 1if Ss,t,i is higher than the upper storage limit (SUL) in month for scenario i,

first stage decision variable for t=1-3, and second stage decision variable for t=4-12.

Binary variable, 1 if Ss,t,i is less than lower storage limit (SLL) in month for scenario i, first

stage decision variable for t=1-3, and second stage decision variable for t=4-12.

∑ (∑ )

∑ ∑ ∑ ( ) ∑ (5.10)

Subject to:

Monthly available biomass for each supplier s with fixed contract in each month t and each

scenario i:

[For s∈{Supplier1, …, Supplier4}, all t and i] (5.11)

91

Monthly available biomass for each supplier s without fixed contract in each month t and each

scenario i:

[For s∈{Supplier5, …, Supplier8}, all t and i] (5.12)

Maximum storage level in each month t and each scenario i:

∑ [For all t and i] (5.13)

Last time period target storage level for each scenario i:

∑ [For t=Dec and all i] (5.14)

Mass balance equation for storage, consumption and purchase for each supplier s, each month t

and each scenario i:

[For all s, t and i] (5.15)

Energy demand in each month t and each scenario i:

[For all t, i] (5.16)

Converting biomass to electricity in each month t and each scenario i:

∑ (∑ ) ( )

[For all t, i] (5.17)

92

And non-negativity of all variables for each supplier s, each month t and each scenario i:

, , , [for all s, t, i] (5.18)

In the stochastic programming model, it is assumed that at the start of each stage, the decision

makers have sufficient information on the amount of available biomass in that stage and can

adjust the production plan for that scenario. A decision xt is called implementable if it cannot

distinguish between different scenarios that are indistinguishable at a stage (Kazemi Zanjani et

al., 2010b). The constraints that are implying this are called non-anticipatively constraints as

follows.

, ,

[for t ∈{Jan,…,Mar}, all i and i' (i≠i') and s] (5.19)

5.4 Managing the risk

The output of a stochastic programming model is the total expected profit of all scenarios. This

model is risk-neutral because only the expected profit is considered in the formulation and

profit variations are ignored. The desired solution from a management perspective may include

decisions that generate higher than the expected profit or a desired target profit for a scenario.

Usually, decision makers are interested in knowing the potential risk and managing it while

optimizing the economic objective. To do this, different risk metrics can be defined and

included in the optimization model. In this paper, variability index and downside risk are

considered as two risk measures (You et al., 2009) and a weighted bi-objective model is

developed for the first risk measure. The objective functions are: to maximize the total

expected profit and minimize the risk measures.

93

5.4.1 Variability index

Uncertainty in input parameters of an optimization model results in uncertainty in the optimal

solution. Therefore, the solution, which in our case is profit, has mean and variance.

Considering the variance of the optimum solution in addition to its expected value is one of the

formula io s ha have bee called “robus op imi a io ” i he li era ure (Mulvey et al., 1995).

In this method, the variance of the total profit is minimized while the expected total profit is

maximized. However, using variance adds a quadratic term to the objective function which

makes the optimization model non-linear. Moreover, it treats the profits higher and lower than

the expected profit in the same way. Therefore, instead of variance in this study the negative

deviation between the scenario profit (Profiti) and the expected profit (Profit) is considered as

the risk measure as defined by Ahmed & Sahinidis (1998). The variability index is defined as a

non-negative continuous variable (∆i) for each scenario and the following constraint is added to

the model:

∑ [ , for all i] (5.20)

According to Equation 27, if the profit of scenario i (Profiti) is less than the expected profit

(Profit), ∆i is equal to their positive difference; otherwise it is 0. The objective function then is

to maximize the weighted sum of the total expected profit minus the expected variability index

as follows:

∑ ∑ ) (5.21)

94

Where ρ is the weight associated with the second objective function. The output of this bi-

objective model provides the trade-off between the expected profit and the profit variability

index.

5.4.2 Downside risk

When considering risk, the extremes of the optimum solutions are also important in addition to

their variability. It is desirable to have a lower probability of having low profit or a higher

probability of having high profit. To address this, another risk metric is used here called

downside risk, which is defined as the probability that the real profit is less than a certain

threshold or target Ω (You et al., 2009). The difference between the variability index defined

earlier and the downside risk is that in the latter the deviation is evaluated by comparing to a

target level while in the former it is compared to the expected profit. Here, the variable ψi is

defined as the positive deviation between the target Ω and the profit of scenario i (Profiti). If

the scenario profit (Profiti) is less than the target Ω, ψi is equal to their difference, otherwise it

is 0, as shown by the following constraints:

[ψi≥0, for all i] (5.22)

The following equation defines the downside risk metric associated with target Ω:

∑ (5.23)

In this case, Risk(Ω) is set to be equal to an appropriate predefined level and constraint 5.22 is

added to the model.

95

5.5 Results

5.5.1 Result of deterministic models

Based on the assumption for the initial storage explained in 5.3, the optimum profit for the

MILP model was $16.2M, 5% higher than that of the MINLP mode, which was $15.4M. It has

to be noted that there was no guarantee for the MINLP model to provide the exact optimal

solution since the model was not convex.

5.5.2 Results of the stochastic model

Table 5-3 shows the expected value of profit for three cases. Column 2 indicates the results if

perfect information on the scenario that is going to happen was available. In this case, the

deterministic model is solved for each scenario one by one. The average of profit for all

scenarios are calculated and shown in column 2 (scenario analysis model). However, in reality,

such perfect information is not available, and this is the motivation for developing a stochastic

model. Column 3 shows the expected profit for the stochastic model. The deterministic model

is usually solved using the average values for uncertain parameters. If the first stage decisions

based on average values are implemented, while other scenarios than the average scenario

occur in the second stage, the optimization model can be solved with implemented first stage

decisions to determine second stage decisions. Column 4 of Table 5-3 shows the expected

profit when the deterministic model is solved for each scenario using the first stage decisions of

the average scenario.

96

Table 5-3: Expected value of profit for scenario analysis, stochastic and average scenario

models ($M)

Deterministic

model –scenario

analysis

Stochastic

model

Deterministic model -

first stage decisions based

on average scenario

Expected value of profit 16.2 16.0 15.8

It can be seen that the expected value of profit for the scenario analysis is $16.2M which is

$0.2M higher than the expected value of profit for the stochastic model ($16.0M). This

difference is also called “Expected Value of Perfect Information” (EPVI) meaning that this is

the cost that the power plant is paying due to the existence of uncertainty in available biomass.

On the other hand, the expected value of profit of the stochastic model is $0.2M higher than the

expected value of profit for the case that the deterministic model result for the average scenario

is implemented but other scenarios occur. This difference is named “Value of Stochastic

Solution” (VSS) and represents the profit that is associated with inclusion of uncertainty in

available biomass and justifies using more sophisticated methods such as stochastic

programming in production planning of the power plant (Birge & Louveaux, 1997).

Figure 5-1 shows the histogram of profit for using average scenario results and stochastic

model.

It can be seen that the standard deviation of the average scenario is lower than that of the

stochastic model. It also indicates that the frequency of having very high profit ($17-18M) and

very low profit ($12-13M) is higher in the case of the stochastic model while there is a higher

probability of having profit in the range of $16-17M in the deterministic model using average

scenario results for the first stage decisions.

97

Figure 5-1: Histogram of profit distribution for the deterministic model with first stage

decisions based on average scenario and the stochastic model

Table 5-4 presents the biomass procurement cost, including purchase and transportation costs,

from the deterministic model with average scenario and the stochastic model and their

difference for each scenario. It can be seen that while the difference between the average

procurement costs over all scenarios is $0.1M, this difference could be as high as $500,000 for

some scenarios. It means that employing the stochastic programming model results has $0.5M

lower costs for the company for some of the scenarios whereas the maximum loss of using

stochastic model is only $0.2M for one scenario. This also demonstrates that developing and

implementing the stochastic model reduces the biomass purchase cost for the power plant.

If the deterministic model is solved for each scenario one by one, then the first stage decision

variables are fixed for each scenario, and the model is solved to determine the second stage

decisions of each scenario, 9 scenarios will have infeasible solutions as shown in Table 5-5.

Therefore, ignoring uncertainty and using the deterministic model results for different

scenarios, result in a 33% risk of having an infeasible solution. Moreover, as shown in Table

0%

10%

20%

30%

40%

50%

60%

70%

12.0-12.9 13.0-13.9 14.0-14.9 15.0-15.9 16.0-16.9 17.0-17.9

Fre

quen

cy

Profit ($M)

Stochastic Model Deterministic Model with Average Scenario

98

5-5 the average profit for those scenarios with feasible results are always less than that of the

stochastic solution, which is $16.0M.

Table 5-6 indicates the standard deviation of the amount of biomass consumption, which is one

of the decision variables of the model, of 27 scenarios for each month from all suppliers, for

both the deterministic scenario analysis and the stochastic model. It can be seen that for the

stochastic model the standard deviation is zero for the first stage decisions, as expected from

definition of non-anticipatory constraints. The decision variables are therefore independent of

the scenario occurring in the future which makes them implementable. The same table can be

presented for other decision variables, such as the amount of biomass purchased in each month.

Other remarks from the results are as follows. In the scenario analysis model, the optimization

results indicate that the surplus load should not be produced for three scenarios, while in the

stochastic model the surplus load is always produced. Moreover, the storage level never go

beyond the upper and lower limits ( , and ) neither in stochastic nor in

deterministic models. Therefore, there is no risk of having very low or very high storage levels

related to the uncertainty in available biomass.

It should be noted that the model and the above comparisons are related to a process where

decisions on the first stage are used and implemented. It is possible that after implementing the

first three months, the system is re-planned with a new stochastic model over the rest of the

year, i.e. 3+6 months as the new first and second stages.

99

Table 5-4: Biomass procurement cost for each scenario of stochastic and deterministic models

($M)

Scenario Stochastic Model

(SM)

Deterministic Model -

Average Scenario (DM)

Difference

(SM-DM)

S1 6.7 7.0 -0.5

S2 6.9 7.4 -0.5

S3 7.2 7.6 -0.4

S4 7.0 7.5 -0.5

S5 7.2 7.7 -0.5

S6 8.0 8.0 0.0

S7 7.4 7.8 -0.4

S8 8.2 8.1 0.1

S9 9.1 9.0 0.1

S10 7.0 7.5 -0.5

S11 7.2 7.7 -0.5

S12 7.9 7.9 0.0

S13 7.4 7.8 -0.4

S14 8.1 8.1 0.0

S15 9.0 8.9 0.1

S16 8.4 8.3 0.1

S17 9.2 9.1 0.1

S18 10.1 10.0 0.1

S19 7.4 7.8 -0.4

S20 8.1 8.1 0.0

S21 9.0 8.9 0.1

S22 8.4 8.3 0.1

S23 9.2 9.1 0.1

S24 10.1 10.0 0.1

S25 9.4 9.3 0.1

S26 10.3 10.2 0.1

S27 11.3 11.1 0.2

Average 8.3 8.4 -0.1

100

Table 5-5: Average profit if the first stage decision variables of each scenario is implemented

and other scenarios happen ($M)

Scenario which its first stage

decisions are implemented

Expected value if other 26 scenarios

happen

S1 Infeasible

S2 Infeasible

S3 Infeasible

S4 Infeasible

S5 Infeasible

S6 15.8

S7 15.9

S8 15.8

S9 15.7

S10 Infeasible

S11 Infeasible

S12 15.9

S13 15.9

S14 15.8

S15 15.5

S16 15.8

S17 15.4

S18 Infeasible

S19 15.9

S20 15.8

S21 15.5

S22 15.8

S23 15.4

S24 15.3

S25 15.3

S26 15.3

S27 Infeasible

101

Table 5-6: Average and standard deviation of the monthly biomass consumption for

deterministic model with scenario analysis and stochastic models (1000 green tonnes)

Deterministic model - Scenario

analysis

Stochastic model

Average STD Average STD

Jan 41 0.3 42 0

Feb 41 0.4 41 0

Mar 40 0.3 40 0

Apr 41 1.3 41 0.4

May 26 0.3 25 0.3

Jun 35 0.4 35 0.7

Jul 43 1.4 43 1.2

Aug 41 1.4 41 0.7

Sep 40 0.8 39 0.3

Oct 42 2.7 43 0.2

Nov 42 0.3 42 0.4

Dec 43 0.2 43 0.2

5.5.3 Results for the variability index

Figure 5-2 shows the mean and standard deviation of the total profit for the bi-objective model

containing variability index mentioned in 5.4.1. It can be seen that as coefficient ρ increases,

lower standard deviation is obtained with lower average profit e.g. to reduce standard deviation

from $1.2M to 0, the profit reduces from $16M to $13.5M too.

Figure 5-3 presents the histogram of profit distribution before and after managing the

variability index. It is observed that after managing the variability index (ρ=2) the expected

profit is less spread, while it has more variation when risk is not managed (ρ=0). However,

assigning a high weight to the risk in the objective function (e.g. ρ=5) results in lower

probability of having higher profit without reducing the probability of having lower profit. It

102

means that although the standard deviation of profits of different scenarios has been reduced,

they are distributed around a lower expected value. Therefore, it is not reasonable to assign

high weight to the risk in the objective function.

Figure 5-2: Profit mean and standard deviation for different weights (ρ)

Figure 5-3: Histogram of profit distribution for different weights (ρ) associated with variability

index

0

0.2

0.4

0.6

0.8

1

1.2

1.4

12

12.5

13

13.5

14

14.5

15

15.5

16

16.5

0 1 2 5 7 10 20 27

Sta

nd

ard

Dev

iati

on (

$M

)

Pro

fit

($M

)

ρ

Profit STDV

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

12.0-12.9 13.0-13.9 14.0-14.9 15.0-15.9 16.0-16.9 17.0-17.9

Fre

quen

cy

Profit ($ M)

ρ=0 ρ=2 ρ=5

103

5.5.4 Results for the downside risk

Figure 5-4 shows the result after managing the downside risk as mentioned in 5.4.2 with the

target profit set at $14M (Ω=$14M). Before managing the downside risk, the expected value of

profit was $16.0M and Risk($14M)=$40,669, calculated based on Equations 5.22 and 5.23. In

order to reduce the downside risk to $30,000 (Risk($14M)=$30,000), the total the expected

profit decreases to $15.8M. It can also be observed that after managing downside risk, the

probability of having high profit in the range of $17.0-18.0M is zero.

Figure 5-4: Histogram of profit distribution for before and after managing the downside risk

(Ω=$14M)

5.6 Discussion and conclusions

In this chapter, a mixed integer linear programming model is developed to optimize the supply

chain of a forest biomass power plant at the tactical level, based on a previously developed

mixed integer non-linear programming model. The developed model was then extended to a

two-stage stochastic linear programming model to include uncertainty in available biomass

0%

10%

20%

30%

40%

50%

60%

70%

12.0-12.9 13.0-13.9 14.0-14.9 15.0-15.9 16.0-16.9 17.0-17.9

Fre

quen

cy

Profit ($M)

Risk=$40669 [max E(profit)] Risk=$30000

104

from different suppliers. The results of the case study showed that having the perfect

information about monthly available biomass from suppliers, he power pla ’s a ual profi

would be $16.2M. However, in reality, the amount of available biomass from supplier varies

due to a number of factors such as market fluctuations resulting in an average profit of $15.8M,

$0.4M lower than $16.2M. Part of this loss could be compensated using a stochastic

programming model. The profit of the stochastic model was $0.2M higher than the solution

when the average scenario was implemented while other scenarios occurred and $0.2M lower

than the solution of the deterministic model with the perfect information. The procurement cost

of the stochastic model was $0.5M lower than that of the average scenario. The reason that the

stochastic model had lower profit than the deterministic model was that in the stochastic

optimization, the expected value over all scenarios of monthly available biomass, including the

scenarios with 20% reduction in supply, were optimized. In addition, ignoring uncertainty and

implementing deterministic model results for each scenario resulted in a 33% chance of having

an infeasible solution. Moreover, the stochastic model decision variables were implementable

since they did not depend on future uncertainties. This was obtained, however, with a more

sophisticated model, higher amount of computational effort and the need for data and

assumptions about uncertain parameters.

Two risk measures were also considered in this study and two models were developed and

solved in order to explicitly consider the risks included in the supply chain planning process.

The considered risk measures were the variability index and downside risk. The results of risk

management models indicated that the total expected profit of the power plant would reduce

after risk management. Lower profit variance was obtained by managing the variability index,

but it resulted in lower probability of having higher profit. When the downside risk is

105

controlled, the probability of having very low and very high profit decreased. It is important to

manage risk measures properly in order to prevent low expected values for profit. Basically,

this is up to the power plant managers to decide on the risk measures as well as the risk

threshold that are acceptable to them, and then implement the optimum solution of using those

risk measures and thresholds in the models.

This chapter addressed the thesis fourth objective partially by including uncertainty in biomass

availability in a two-stage stochastic optimization model. It is important to incorporate the

uncertainty in biomass quality into the modeling to have a comprehensive robust model in

response to uncertainty. This is done in the next chapter.

106

Chapter 6 Hybrid stochastic programming-robust optimization model

6.1 Synopsis

In the previous chapter, uncertainty in one parameter, biomass availability, was included in the

optimization model. In this chapter, uncertainty in several parameters of the power plant supply

chain is incorporated simultaneously. Uncertain parameters incorporated here are biomass

quality, namely moisture content and higher heating value, in addition to monthly available

biomass from suppliers. Uncertainties in electricity prices and biomass cost are not included

because the power plant has long term contracts with the client and most of the suppliers. This

means the prices and costs do not change during the planning horizon. i.e. a year. There is a

small variation in biomass cost over a year for suppliers without a fixed contract, therefore,

uncertainty in biomass cost for those suppliers can be ignored as well.

Moisture content and higher heating value change within a range and are considered in a

developed robust optimization model. Uncertainty in monthly available biomass is included

through a two-stage stochastic programming model as explained in Chapter 5. A hybrid

stochastic programming-robust optimization model is used to incorporate uncertainty in all of

the mentioned parameters simultaneously, and to have an appropriate balance between

computational complexity associated with the stochastic programming formulation and

conservatism of the optimal solution (lower profit) associated with the robust optimization

approach.

6.2 Robust optimization formulation

Two indicators of biomass quality are considered here: moisture content (MC) in % and higher

heating value (HHV) in BTU/lb. These parameters are not correlated since HHV is calculated

107

based on dry biomass. Energy value depends on MC and HHV as was shown in Equation 3.14.

Based on the historical data provided by the power plant, the range of variation is considered to

be 25-35% for MC with an average of 30%, and 8000-9000 BTU/lb (5.17-5.81 MWh/tonnes)

for HHV with an average of 8500 BTU/lb (5.49 MWh/tonnes). To be able to include

uncertainty in MC and HHV in a stochastic programming model, different scenarios can be

obtained for them by discretizing their ranges. However, if these scenarios are combined with

the 27 scenarios of monthly available biomass, the total number of scenarios is going to be

huge even with a small number of discretized scenarios. For instance, considering only 3

scenarios for each of MC and HHV, within the scenario structure described above for monthly

available biomass, the number of scenarios will be 93×27=19683. This is even without

considering the combination of different MC and HHV for different suppliers. Therefore,

robust optimization is used to include uncertainty in biomass quality into the decision making

process.

The formulation for the robust optimization is derived from Ben-Tal et al. (2009) which is

available in the AIMMS software package. It helps to come up with a tractable model

providing a solution that is feasible for all MC and HHV ranges with a price of having lower

profit since the worst case is optimized. Here, a brief explanation of the idea of robust

optimization is provided.

Consider a general linear programming (LP) model as shown in Equation 6.1:

(6.1)

108

Where x represent model decision variables, c is the objective function coefficients vector, b is

the right hand side values vector and A is the constraints coefficient matrix. If the parameters of

the model, c, A and b, are uncertain and range within a convex set of U, then the robust

counterpart of this uncertain LP model can be defined as shown in Equation 6.2:

∈ ∈ ∈ (6.2)

It should be noted that in robust optimization, the objective is to optimize the worst case.

Equation 6.2 can be reformulated as follows:

∈ ∈ ∈ (6.3)

The robust counterpart model shown in Equation 6.3 has a single objective function and

continuous constraints. Any solution for Equation 6.3 satisfies the constraints of

Equation 6.2 for c, A and b∈U. These solutions are robust feasible and among such robust

feasible solutions, the robust optimal value, namely , is the one providing the best possible

objective function for Equation 6.3. Ben-Tal & Nemirovski (2002) proved that problem shown

in Equation 6.3 is computationally tractable (or in the other words polynomially solvable) for a

wide choice of the uncertainty sets U. Particularly, problem in Equation 6.3 is an LP model if

U is a box or polyhedral set (Li & Floudas). Without loss of generality, only uncertainty in

parameters of matrix A, named as is considered. If belongs to a box set, the uncertainty in

is defined as:

(6.4)

109

where is the nominal value of parameters and is the positive constant perturbation and is

independent random variables which are subject to uncertainty. Then, constraint –

can be written as:

∑ ∑ (6.5)

The solution has to be feasible for any in a given uncertainty set U or alternatively:

∑ ∈ ∑ (6.6)

In particular for the box uncertainty set the random vector is defined as:

(6.7)

Where is an adjustable parameter controlling the size of uncertainty set. Then the

corresponding robust counterpart in Equation 6.3 is equivalent to the following constraints:

∑ ∑ – (6.8)

(6.9)

Where u is a decision variable (u≥0). Equation 6.9 is still linear (Li & Floudas). In the present

model, uncertain parameters belong to certain ranges and therefore can be considered as box.

This formulation is being used to include uncertainty in biomass quality in the optimization

model presented in this thesis. The MILP model presented in Chapter 5 is first extended to a

robust optimization model. The formulation for robust optimization model, in case of

uncertainty only in MCs,p,t is included is as follows:

(6.10)

110

∑ (∑ ) ∑

∑ ∑ ∑ (6.11)

Subject to:

[For s∈{Supplier1, …, Supplier4} and all t] (6.12)

[For ∈{Supplier5, …, upplier8} and all t] (6.13)

∑ [For all t] (6.14)

∑ (6.15)

[For all s and t] (6.16)

[For all t] (6.17)

∑ ∑ –

[For ∈[25,30] and all t] (6.18)

All continuous variables have to be non-negative as shown in Equation 6.19.

, , , [for all s, t] (6.19)

Notice that constraint (6.18) is converted to inequality in order to make sure it can be met for

all realization of .

111

The robust counterpart of Equation 6.18 is as follows (according to Equations 6.8 and 6.9):

∑ (∑

( )) –

∑ (∑

) –

[for all t] (6.20)

, (6.21)

Where s,p,t is the positive constant perturbation in MCs,p,t and ψ is the adjustable parameter

controlling the size of uncertainty set. The same formulation can be written for uncertainty in

HHVs,p,t. When uncertainty in both parameters is included, the term HHVs,p,t × (1 – MCs,p,t) is

replaced by EVs,p,t and the formulation is written for this parameter with the range derived from

ranges in both HHVs,p,t and MCs,p,t.

6.3 Hybrid stochastic programming-robust optimization model

Here, the formulation for hybrid stochastic programming-robust optimization model, in case of

uncertainty in MCs,p,t only, is presented:

(6.22)

∑ (∑ ) ∑

∑ ∑ ( ) ∑ (6.23)

112

Subject to:

[For s∈{Supplier1, …,Supplier4}, all t and i] (6.24)

[For s∈{Supplier5, …,Supplier8}, all t and i] (6.25)

∑ [For all t and i] (6.26)

∑ [For t=Dec and all i] (6.27)

[For all s, t and i] (6.28)

[For all t, i] (6.29)

∑ ∑ –

[For ∈[25,30] and all t and i] (6.30)

All continuous variables have to be non-negative as shown in Equation 6.19.

, , , [for all s, t, i] (6.31)

Again, constraint (6.30) is converted to inequality in order to make sure it can be met for all

realization of .

The robust counterpart of Equation 6.30 is as follows:

113

∑ (∑ ) ( – )

∑ (∑ ) ( – )

(6.32)

(6.33)

The same formulation can be written for uncertainty in HHVs,p,t. When uncertainty in both

parameters is included, the term HHVs,p,t×(1-MCs,p,t) is replaced by EVs,p,t and the formulation

is written for this parameter with the range derived from ranges in both and .

6.4 Results

First, the results of the robust optimization model are presented where variations in biomass

quality are considered. A solution that is feasible for all possible instances of uncertain

parameters will likely be a conservative solution. The model is solved with different ranges for

MCs,p,t and HHVs,p,t values to assess how the solution changes when the uncertainty set is

widened, the results can be seen in Table 6-1. The profit of the robust optimization model using

the average MCs,p,t (30%) and average HHVs,p,t (8500 BTU/lb) is $15.64M. Using the average

HHVs,p,t in the model, as the uncertainty set of MC widens from 2% to 10%, the optimum profit

drops from $15.07M to $13.13M, which is also shown in Figure 6-1. Alternatively, using the

average MCs,p,t of 30% in the model, as the uncertainty set of HHVs,p,t is expanded from 200

BTU/lb (0.13 MWh/tonnes) to 2000 BTU/lb (1.29 MWh/tonnes), the optimum profit reduces

from $15.15M to $13.58M, as shown in Figure 6-2. When both parameters are considered

uncertain at the same time, the result becomes more conservative with a more severe reduction

in profit to $11.97M for HHVs,p,t∈[8100,8900], MCs,p,t∈[26,34]. In the extreme ranges, the

114

model is infeasible. Figure 6-3 shows the results when uncertainty is included in both MCs,p,t

and HHVs,p,t.

Table 6-1: Profit ($M) for different ranges of MCs,p,t and HHVs,p,t used in the robust

optimization model

MCs,p,t (%)

30 (Average) 29-31 28-32 27-33 26-34 25-35

HH

Vs,

p,t

8500 (BTU/lb) (Average)

5.49 (MWh/tonnes)

15.64 15.07 14.61 14.13 13.64 13.13

8400-8600 (BTU/lb)

5.42-5.56 (MWh/tonnes)

15.15 14.72

8300-8700 (BTU/lb)

5.36-5.62 (MWh/tonnes)

14.78 13.76

8200-8800 (BTU/lb)

5.30-5.69 (MWh/tonnes)

14.39 12.94

8100-8900 (BTU/lb)

5.23-5.75 (MWh/tonnes)

13.99 11.97

8000-9000 (BTU/lb)

5.17-5.81 (MWh/tonnes)

13.58 Infeasible

Figure 6-1: Solution of the robust optimization model for different ranges of moisture content

11

12

13

14

15

16

29-31% 28-32% 27-33% 26-34% 25-35%

Pro

fit

($M

)

Moisture Content Range

115

Figure 6-2: Solution of the robust optimization model for different ranges of higher heating

value

Figure 6-3: Solution of the robust optimization model for different ranges of energy value

Despite having a more conservative solution with a lower profit, the decisions provided by

robust optimization are feasible for ranges of MC and HHV shown in Figure 6-3. The model

was run for the case where the decision variables of biomass purchase, storage and

consumption in the first three months are made and implemented based on the results of the

model using average MC and HHV (MCs,p,t=30% and HHVs,p,t=8500 BTU/lb). Then, it was

11

12

13

14

15

16

8400-8600 8300-8700 8200-8800 8100-8900 8000-9000

Pro

fit

($M

)

Higher Heating Value Range (BTU/lb)

11

12

13

14

15

16

29-31%

8400-8600

28-32%

8300-8700

27-33%

8200-8800

26-34%

8100-8900

Pro

fit

($M

)

Moisture Content Range

Higher Heating Value Range (BTU/lb)

116

assumed that for the rest of the months, MC and HHV take other values (any of the ranges

showed in Table 6-1, for instance HHVs,p,t∈[8100,8900] and MCs,p,t∈[26,34]). It was observed

that the model becomes infeasible when MCs,p,t and HHVs,p,t vary from their average values,

which in reality is very probable.

Figure 6-4 and Figure 6-5 show the optimum biomass storage and consumption level in

different months based on the results of the robust optimization model with

HHVs,p,t∈[8100,8900], MCs,p,t∈[26,34] and the deterministic model. It can be seen that robust

optimization provides more storage levels compared to the deterministic model for most of the

months and more consumption levels in all months. The consumption level is higher due to a

lower than average energy value of biomass used in the robust optimization model.

Figure 6-4: The optimum storage level in different months from the robust optimization model

with HHVs,p,t∈[8100,8900] and MCs,p,t∈[26,34] and the deterministic model

0

10

20

30

40

50

60

70

80

90

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Sto

rag

e L

evel

(1

00

0 G

reen

To

nn

es)

Robust Optimization

Deterministic

117

Figure 6-5: The optimum biomass consumption level in different months from the robust

optimization model with HHVs,p,t∈[8100,8900] and MCs,p,t∈[26,34] and the deterministic

model

The result of the hybrid stochastic programming-robust optimization model is provided in

Table 6-2. The first two columns show MC and HHV ranges. The next columns show the

robust optimization solutions, the hybrid stochastic programming-robust optimization solutions

and their differences, respectively. It can be seen that the hybrid stochastic programming-

robust optimization model provides a more conservative solution compared to the robust

optimization model. However, the reduction in profit is small (0.2-0.7%). The reduction in

profit increases as the range of uncertainty decreased. For the average values of HHV and MC,

the profit for the deterministic model is $15.64M and it is $15.45M for the stochastic

programming model.

Despite having a lower profit from the hybrid stochastic programming-robust optimization

model, it provides feasible solutions for all scenarios of monthly available biomass and all

ranges of variation in MC and HHV. Figure 6-6 shows the optimum storage level over 27

scenarios for the robust optimization model and the hybrid stochastic programming-robust

optimization model when HHVs,p,t∈[8100,8900] and MCs,p,t∈[26,34]. To make the graphs

0

10

20

30

40

50

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Bio

ma

ss C

on

sum

pti

on

(1

00

0 G

reen

To

nn

es)

Robust Optimization

Deterministic

118

readable, only the storage levels from January to May are depicted. It can be seen that the first

stage decision variables are unique and independent when stochastic programming is used.

Table 6-2: Profit for different ranges of MCs,p,t and HHVs,p,t used in the robust optimization and

hybrid models

MCs,p,t

range (%)

HHVs,p,t range

(BTU/lb)

Robust optimization

solution ($M)

Hybrid model

solution ($M)

Difference

($M)

26-34 8100-8900 11.97 11.93 0.03

27-33 8200-8800 12.94 12.87 0.07

28-32 8300-8700 13.76 13.66 0.10

29-31 8400-8600 14.72 14.61 0.11

Figure 6-6: Optimum storage level of 27 scenarios for the first five months for a) robust

optimization model, and b) hybrid stochastic programming-robust optimization model when

HHVs,p,t∈[8100,8900] and MCs,p,t∈[26,34]

6.5 Discussion and conclusions

In this chapter, first uncertainty in biomass quality is included in the optimization model of a

forest bioenergy supply chain through developing a robust optimization model. This model

provides a solution that is feasible for any value of uncertain parameters within a defined

range. Different profit can be generated depending on the variation range for MC and HHV.

For the average MC (30%) and average HHV (8500 BTU/lb), the profit was $15.64M. For

40

50

60

70

80

90

100

Jan Feb Mar Apr May

Sto

rag

e (1

00

0 G

reen

To

nn

es)

40

50

60

70

80

90

100

110

120

Jan Feb Mar Apr May

Sto

rag

e (1

00

0 G

reen

To

nn

es)

119

average HHV and MC range of 29-31%, the profit decreased to $15.07M, while it dropped to

$13.13M for MC range of 25-35%. When HHV changed within 200 BTU/lb range (8400-8600

BTU/lb) with the average MC, the optimum profit reduced to $15.15M. It also decreased to

$13.58M, as the HHV range was widened to (8000-9000 BTU/lb). When MC and HHV

changed simultaneously the solution became more conservative. For instance, the profit for the

widest range with a feasible solution, which was (26-34%) for MC and (8100-8900 BTU/lb)

for HHV, was $11.97M. From these results, it was concluded that as the range of uncertain

parameters widened, the profit obtained from the robust optimization model decreased. The

reason was that in the robust optimization model, the worst case scenario was optimized.

Expanding the range of the uncertain parameters made the worst case scenario worse.

A hybrid stochastic programming-robust optimization model was then proposed to incorporate

uncertainties in different parameters simultaneously. Stochastic programming was used to

include uncertainty in monthly available biomass using 27 scenarios over a one-year time

horizon. Robust optimization was used to model the uncertainty in biomass quality within a

polyhedral set. Comparing the results of robust optimization and stochastic programming

models, it can be concluded that the conservatism in the robust optimization model was more

severe because it optimized the worst cases while the stochastic programming model optimized

the expected value of all scenarios. The degree of conservatism can be controlled by selecting

the appropriate range of uncertain parameters. The conservative solution was compensated by

having more stable solution which was free of infeasibility risk. No matter what happens in

terms of uncertain parameter values, one can be sure that the model remains feasible and the

decisions can be adjusted if any perturbation occurs within the defined sets or scenarios. To

balance the conservatism of optimizing the worst case, stochastic programming was used to

120

include uncertainty in available biomass. When the hybrid model was implemented, the profit

reduced even more. For instance, for the case with (26-34%) range for MC and (8100-8900

BTU/lb) range for HHV, the profit was $11.93M, $0.03M lower than that of the robust

optimization model with the same ranges, $11.97M. However, this model has the advantage of

providing a solution which is implementable, e.g. unique decision variables for the first stage.

The stochastic programming results have the advantage of being unique for the first stage, or in

other words, being implementable.

This chapter completed on the fourth objective of the thesis. Making decisions in supply chain

planning at the tactical level considering uncertainty in biomass quality was unavailable before.

The hybrid model is a unique approach to include uncertainty in several parameters

simultaneously and balance the advantages and disadvantages of both stochastic programming

and robust optimization approaches.

121

Chapter 7 Conclusions, strength points, limitations and future research

7.1 Conclusions

The optimization models developed and presented in Chapters 3-6 achieved all the objectives

of this study. The deterministic model presented in Chapter 3 achieved objectives 1 and 2 and

part of objective 3, the Monte Carlo simulation model presented in Chapter 4 fully achieved

objective 3 and the stochastic programming, robust optimization and the hybrid model

developed and presented in Chapters 5 and 6 achieved the 4th

objective.

From the deterministic model presented in Chapter 3 it was concluded that optimization models

were useful tools that could help managers of forest biomass power plants to make better

decisions on how much biomass to purchase, store and consume in each month to meet the

electricity demand and maximize their profit. Different parts of the supply chain from biomass

supply to storage and electricity production, were integrated in the deterministic model,

therefore, the optimum solution prescribed by the model was based on the impact of parameters

in all parts of the supply chain and the interactions between different supply chain parts. This

would be helpful for decision makers because it would not be possible for them (without using

a decision support tool) to consider all these factors and interactions simultaneously when

making decisions. In fact, the decisions in the power plant are made by different people (fibre

manager, production manager, and some by finance manager). The managers make their own

decisions and then have quarterly meetings to discuss important issues without being able to

know how each decision would affect other decisions or the profit as a whole. The model

developed here is useful since it integrates all the supply chain parts.

The impact of uncertainty in input parameters on the profit was shown from sensitivity analysis

results. These results showed that variations in electricity price, higher heating value,

maximum available biomass from suppliers with fixed contracts, moisture content and biomass

122

cost had high impacts on the optimum solution. The results of the Monte Carlo simulation

model showed that variation in moisture content had a higher impact on profit than variation in

higher heating value.

Including uncertainty in the decision making process may seem to provide lower profit for the

power plant. However, it provided a robust solution against perturbations in input parameters.

The solution was feasible for all considered instances of uncertain parameters and was

implementable. It was also concluded that there was a trade-off between expected value of

profit and probability of having low profit. Moreover, it was shown that developing a hybrid

model that was a combination of stochastic programming and robust optimization models

provided a flexible framework for considering uncertainty in different parameters at the same

time. The hybrid model took advantage of both stochastic programming and robust

optimization models; it is computationally tractable and not too conservative. If the hybrid

model is implemented, the power plant managers would make sure even with the changes in

the parameters (biomass availability and quality), that are probable in reality, the power plant

still would be able to generate electricity and meet the demand, have biomass for production

and do not have shortage or an excess storage.

7.2 Strengths points

The main strength and contribution of this thesis is that the issue of uncertainty which is a

critical factor in the forest biomass supply chain is incorporated in the modeling and analysis

comprehensively. This can be used as a guideline for dealing with uncertainty in forest

biomass/ bioenergy supply chains, or even other supply chains that have a similar structure.

The case study presented here, is a real forest biomass power plant. Having access to real data

and being able to communicate with people working directly in this industry is a strength point

123

of this research. The model structure and solutions were validated by the power plant

managers. Moreover, data analysis, particularly probability distribution of biomass quality for

different biomass types in different months, provided useful information that could be used in

other relevant studies.

The only study similar to this work was conducted by Alam et al. (2012b). They developed an

optimization model with monthly time steps for supply chain management of a forest biomass

power plant. The main difference between this deterministic optimization model and the one

developed by Alam et al. (2012b) is that in the former, biomass quality for each biomass types,

from each suppliers and in each months is considered to be different. This allows variations in

biomass availability and quality during the year to be included in the optimization model. Alam

et al. (2012b) used an average value for energy value and moisture content in all months.

Moreover, they focused more on the harvesting side of the supply chain while the model

developed here included the whole supply chain from procurement to storage, production and

ash management. It also included the effects of mixing biomass and storage on biomass quality

and production efficiency which were not done before, even in studies of other forest bioenergy

plants such as district heating systems or biofuel plants.

Most of the time, uncertainty is ignored due to the lack of data, complexity of models with

uncertainty, and difficulty to solve them. All the studies in forest biomass power plants supply

chains ignored uncertainty. Alam et al. (2012b) only evaluated the impact of uncertainty in the

parameters of an optimization model of a forest biomass power plant only through sensitivity

analysis. They did not include uncertainty in the modeling and decision making process. In this

thesis, historical data were used for identifying and quantifying uncertainty in different

parameters. Ignoring uncertainty in parameters gave infeasible solutions (33% of the cases),

124

which means the constraints of the model (supply, demand, storage, etc.) were not met. The

approaches for dealing with uncertainties were selected based on the model structure,

characteristics of uncertain parameter, the quality of the solution they provided, and the

computational effort needed to solve them. In the literature, there were studies that developed

stochastic programming models to include uncertainty in similar industries, such as biofuel

supply chains. However, all of them were at the strategic level with one time step. In the

present study, the stochastic programming model was developed and solved for a multiple time

step problem, which is more complex. Developing a hybrid stochastic programming–robust

optimization model to incorporate uncertainty in different parameters of the supply chain was

also a novel idea in field of supply chain management and helped to balance the model

complication and conservatism. Those who work in this field should use models that can

incorporate uncertainties in the decision making such as stochastic modeling, robust

optimization or the proposed hybrid model to make sure the solutions are feasible for all

realizations of uncertain parameters.

7.3 Limitations

One limitation of this study is related to the calculation of the mix of biomass, where it is

assumed that all the biomass in the storage yard is mixed together and an average energy value

is obtained. In reality a portion of biomass may be mixed completely. Moreover, assumptions

were made about the initial storage level and initial quality of biomass provided by each

supplier in storage at the beginning of the planning horizon because such information was not

available.

The structure of the stochastic model was based on the assumption that biomass availability

remains the same for three months and changes by 20% in the following quarters which may

125

not be the case in reality. In the robust optimization model and the hybrid model, the same

ranges were taken for MC and HHV of all biomass types, from each supplier in each month.

This is contrary to the fact that they could be different for each of the indices as was discussed

in Chapter 4. However, in order to address these two limitations a much more complex model

is needed that may still be challenging to solve.

Another limitation of this study is that it does not consider any environmental or social

objectives of the system. Instead, it only focuses on the economic objective and cost/profit in

the system. It should be noted that the main incentive in building and operating a forest

biomass power plant is to provide sustainable energy and local jobs in remote communities, as

explained in the introduction chapter. However, these objectives should be considered more at

the strategic level, e.g. when a forest biomass power plant is being designed.

7.4 Future research

As environmental and social impacts of products and processes are becoming more important

than the pure economic value, future studies using deterministic modeling should incorporate

environmental and social objectives in addition to the economic objective into the models and

perform multi-objective optimization. Considering non-economic objectives into the models

would help deal with important issues, such as emissions, land use, communities' interests, job

creation, governmental policies, ecological impacts of removing residues from forest areas, and

recreational aspects of forests, and provide possible trade-offs between different objectives.

Moreover, to make the model usable by people in the industry, designing a user interface will

be helpful.

It is worthy to integrate different forest supply chains due to their dependency and model the

whole value chain considering harvest areas, sawmills, pulp mills, and bioenergy conversion

126

facilities, and evaluate the effect of uncertainty in one chain on the supply, production and

demand of the other chains. Moreover, integration of planning decisions at strategic, tactical

and operational levels could also be done. These, however, would result in more complex

models that may not be solvable for large case problems.

127

References

Ahmed, S., & Sahinidis, N. V. (1998). Robust process planning under uncertainty. Industrial &

Engineering Chemistry Research, 37(5), 1883–1892.

Ahtikoski, A., Heikkilä, J., Alenius, V., & Siren, M. (2008). Economic viability of utilizing biomass

energy from young stands—The case of Finland. Biomass and Bioenergy, 32(11), 988–996.

AIMMS. Paragon Decision Technology. (Available at:www.aimms.com) Date of access: Jan. 28, 2014.

Alam, M. B., Shahi, C., & Pulkki, R. (2009). Wood biomass supply model for bioenergy production in

Northwestern Ontario. Presented at the 1st International Conference on the Developments in

Renewable Energy Technology (ICDRET).

Alam, M. B., Pulkki, R., & Shahi, C. (2012a). Road network optimization model for supplying woody

biomass feedstock for energy production in Northwestern Ontario. The Open Forest Science

Journal, 5, 1–14.

Alam, M., Pulkki, R., Shahi, C., & Upadhyay, T. (2012b). Modeling woody biomass procurement for

bioenergy production at the Atikokan Generating Station in Northwestern Ontario, Canada.

Energies, 5, 5065–5085.

Alfonso, D., Perpiñá, C., Pérez-Navarro, A., Peñalvo, E., Vargas, C., & Cárdenas, R. (2009).

Methodology for optimization of distributed biomass resources evaluation, management and

final energy use. Biomass and Bioenergy, 33(8), 1070–1079.

Allen, J., Browne, M., Hunter, A., Boyd, J., & Palmer, H. (1998). Logistics management and costs of

biomass fuel supply. International Journal of Physical Distribution & Logistics Management,

28(6), 463–477.

Alvarez, P. P., & Vera, J. R. (2011). Application of robust optimization to the sawmill planning

problem. Annals of Operations Research, Advance online publication (doi: 10.1007/s10479-

011-1002-4).

128

Awudu, I., & Zhang, J. (2012). Uncertainties and sustainability concepts in biofuel supply chain

management: A review. Renewable & Sustainable Energy Reviews, 16(2), 1359–1368.

Awudu, I., & Zhang, J. (2013). Stochastic production planning for a biofuel supply chain under demand

and price uncertainties. Applied Energy, 103, 189–196.

Azar, C., Lindgren, K., & Andersson, B. A. (2003). Global energy scenarios meeting stringent CO2

constraints—cost-effective fuel choices in the transportation sector. Energy Policy, 31(10),

961–976.

Ben-Tal, A., Margalit, T., & Nemirovski, A. (2000). Robust modeling of multi-stage portfolio

problems. In High Performance Optimization (pp. 303–328). Kluwer Academic Publishers.

Retrieved from http://www2.isye.gatech.edu/~nemirovs/rob_portf.pdf. Date of access: Jan. 28,

2014.

Ben-Tal, A., & Nemirovski, A. (2000). Robust solutions of Linear Programming problems

contaminated with uncertain data. Mathematical Programming, Series A, 88(3), 411–424.

Ben-Tal, A., & Nemirovski, A. (2002). Robust optimization-methodology and applications.

Mathematical Programming Series B, 92, 453–480.

Ben-Tal, A., Boaz, G., & Shimrit, S. (2009). Robust multi-echelon multi-period inventory control.

European Journal of Operational Research, 199(3), 922–935.

Benders, J. F. (1962). Partitioning procedures for solving mixed-variables programming problems.

Numerische Mathematik, 4(1), 238–252.

Bertsimas, D., & Sim, M. (2003). Robust discrete optimization and network flows. Mathematical

Programming, 98(1), 49–71.

Bertsimas, D., & Sim, M. (2004). The price of robustness. Operations Research, 52(1), 35–53.

Bertsimas, D., Brown, D. B., & Caramanis, C. (2011). Theory and applications of robust optimization.

SIAM Review, 53(3), 464–501.

BIOCAP Canada. (2008). An information guide on pursuing biomass energy opportunities and

technologies in British Columbia for first nations, small communities, municipalities and

129

industry. Prepared for BC Ministry of Energy, Mines and Petroleum Resources and BC

Ministry of Forests and Range.

Birge, J. R., & Louveaux, F. (1997). Introduction to stochastic programming. Springer Verlag.

Bonami, P. & Gonçalves. J. P. M. (2012). Heuristics for convex mixed integer nonlinear programs.

Computational Optimization & Applications, 51(2), 729–747.

Börjesson, M., & Ahlgren, E. O. (2010). Biomass gasification in cost-optimized district heating

systems—A regional modelling analysis. Energy Policy, 38(1), 168–180.

Bowyer, C., Baldock, D., Kretschmer, B., & Polakova, J. (2012). The GHG emissions intensity of

bioenergy: Does bioenergy have a role to play i reduci g GHG emissio s of Europe’s

economy? London: Institute for European Environmental Policy (IEEP).

URL:http://www.ieep.eu/assets/ 1008/IEEP_-_The_GHG_Emissions_Intensity_of_Bioenergy_-

_October_2012.pdf. Date of access: Jan. 28, 2014.

Bowyer, J. L., Shmulsky, R., & Haygreen, J. G. (2007). Forest products and wood science: an

introduction (5th Edition). Blackwell publishing.

Bradley, D. (2010). Canada report on bioenergy. Natural Resources Canada. Retrieved from

http://www.bioenergytrade.org/downloads/canadareportonbioenergy2010sept152010.pdf. Date

of access: Jan. 28, 2014.

Bredström, D., Flisberg, P., & Rönnqvist, M. (2013). A new method for robustness in rolling horizon

planning. International Journal of Production Economics, 143(1), 41–52.

Bussieck. M. R. & Pruessner A. (2003). Mixed-integer non-linear programming. SIAG/OPT

Newsletter: Views & News, 14(1), 19-22.

Carlsso , D., D’Amours, ., Mar el, A., & Rö qvis , M. (2009). Supply chain planning in the pulp and

paper industry. INFOR, 47(3), 167–183.

Carlsson, D., Flisberg, P., & Rönnqvist, M. (2014). Using robust optimization for distribution and

inventory planning for a large pulp producer. Computers & Operations Research, 44, 214–225.

130

Chen, C.-W., & Fan, Y. (2012). Bioethanol supply chain system planning under supply and demand

uncertainties. Transportation Research Part E: Logistics & Transportation Review, 48(1), 150–

164.

Chinese, D., & Meneghetti, A. (2005). Optimisation models for decision support in the development of

biomass-based industrial district-heating networks in Italy. Applied Energy, 82(3), 228–254.

Chinese, D., & Meneghetti, A. (2009). Design of forest biofuel supply chains. International Journal of

Logistics Systems and Management, 5(5), 525–550.

Chrwan-JYH, H. (1989). Evaluating the impact of operating environments on MRP system

nervousness. The International Journal of Production Research, 27(7), 1115–1135.

Dakin, R. J. (1965). A tree-search algorithm for mixed integer programming problems. The Computer

Journal, 8(3), 250–255.

Dal-Mas, M., Giarola, S., Zamboni, A., & Bezzo, F. (2011). Strategic design and investment capacity

planning of the ethanol supply chain under price uncertainty. Biomass & Bioenergy, 35(5),

2059–2071.

Demirbaş, A., Gullu D., Çaglar A., Akdeniz F. (1997). Estimation of calorific values of fuels from

lignocellulosics. Energy Sources, 19(8), 765–770.

Demirbaş, A. (2001). Biomass resource facilities and biomass conversion processing for fuels and

chemicals. Energy Conversion and Management, 42(11), 1357–1378.

Demirbaş, A. (2003). Fuelwood characteristics of six indigenous wood species from the Eastern Black

Sea region. Energy Sources, 25(4), 309–316.

Demirbaş, A. (2005). Potential applications of renewable energy sources, biomass combustion problems

in boiler power systems and combustion related environmental issues. Progress in Energy &

Combustion Science, 31(2), 171–192.

Difs, K., Wetterlund, E., Trygg, L., & Söderström, M. (2010). Biomass gasification opportunities in a

district heating system. Biomass & Bioenergy, 34(5), 637–651.

131

Duran, M. A., & Grossmann, I. E. (1986). An outer-approximation algorithm for a class of mixed-

integer non-linear programs. Mathematical Programming, 36(3), 307–339.

Ekşioğlu, . D., Acharya, A., Leigh ley, L. E., & Arora, . (2009). A aly i g he desig a d

management of biomass-to-biorefinery supply chain. Computers & Industrial Engineering,

57(4), 1342–1352.

Ells, A., Bulte, E., & van Kooten, G. C. (1997). Uncertainty and forest land use allocation in British

Columbia: vague priorities and imprecise coefficients. Forest Science, 43(4), 509–520.

Eriksson, L. O., & Björheden, R. (1989). Optimal storing, transport and processing for a forest-fuel

supplier. European Journal of Operational Research, 43(1), 26–33.

Fe g, Y., D’Amours, ., Lebel, L., & Nourelfa h, M. (2010). I egra ed bio-refinery and forest products

supply chain network design using mathematical programming approach. Tech. Rep.

CIRRLET-2010-50 Universite de Montreal, Universite Laval Interuniversity Research Centre

on Enterprise Networks, Logistics and Transportation-CIRRLET, Montreal and Laval, Que.

Flisberg, P., Frisk, M., & Ronnqvist, M. (2012). FuelOpt: a decision support system for forest fuel

logistics. Journal of the Operational Research Society, 63, 1600–1612

Freppaz, D., Minciardi, R., Robba, M., Rovatti, M., Sacile, R., & Taramasso, A. (2004). Optimizing

forest biomass exploitation for energy supply at a regional level. Biomass & Bioenergy, 26(1),

15–25.

Frombo, F., Minciardi, R., Robba, M., Rosso, F., & Sacile, R. (2009a). Planning woody biomass

logistics for energy production: A strategic decision model. Biomass & Bioenergy, 33(3), 372–

383.

Frombo, F., Minciardi, R., Robba, M., & Sacile, R. (2009b). A decision support system for planning

biomass-based energy production. WESC 2006 6th World Energy System Conference

Advances in Energy Studies 5th Workshop on Advances, Innovation and Visions in Energy and

Energy-Related Environmental and Socio-Economic Issues, 34(3), 362–369.

132

Fuller, W. S. (1985). Chip pile storage—a review of practices to avoid detoriation and economic losses.

Tappi Journal, 68(8), 48–52.

Gabrel, V., Murat, C., & Thiele, A. (2013). Recent advances in robust optimization: an overview. 2012,

http://www.optimization-online.org/DB_FILE/2012/07/3537.pdf. Date of access: Jan. 28, 2014.

Gebreslassie, B. H., Yao, Y., & You, F. (2012). Design under uncertainty of hydrocarbon biorefinery

supply chains: multiobjective stochastic programming models, decomposition algorithm, and a

comparison between CVaR and downside risk. AIChE Journal, 58(7), 2155–2179.

Geoffrion, A. M. (1972). Generalized Benders decomposition. Journal of Optimization Theory &

Applications, 10(4), 237–260.

Giarolaa, S., Bezzoa, F., & Shah, N. (2013). A risk management approach to the economic and

environmental strategic design of ethanol supply chains. Biomass & Bioenergy, 58, 31–51.

Gilmour, B., & Warren, J. (2007). Advancing district energy development in Canada: a process for site

selection. Review & Community Participation. Available at: http://www.districtenergy.org/

assets/CDEA/Best-Practice/Site-Selection-community-review-UES-Sept-07.pdf. Date of

access: Jan. 28, 2014.

Gülpı ar, N., Pachama ova, D., & Ça akoğlu, E. (2013). Robus s ra egies for facili y loca io u der

uncertainty. European Journal of Operational Research, 225(1), 21–35.

Gunn, E. A. (2009). Some perspectives on strategic forest management models and the forest products

supply chain. International Journal of Physical Distribution & Logistics Management, 47(3),

261–272.

Gunnarsson, H., Rönnqvist, M., & Lundgren, J. T. (2004). Supply chain modelling of forest fuel.

European Journal of Operational Research, 158(1), 103–123.

Gupta, A., & Maranas, C. D. (2003). Managing demand uncertainty in supply chain planning.

Computers & Chemical Engineering, 27(8-9), 1219–1227.

Hall, D. O., & Scrase, J. I. (1998). Will biomass be the environmentally friendly fuel of the future?

Biomass & Bioenergy, 15(4–5), 357–367.

133

Hall, J. P. (2002). Sustainable production of forest biomass for energy. The Forestry Chronicle, 78(3),

391–396.

Han, S.-K., & Murphy, G. E. (2012). Solving a woody biomass truck scheduling problem for a transport

company in Western Oregon, USA. Biomass & Bioenergy, 44, 47–55.

International Energy Agency (IEA). (2012). CO2 emission from fuel combustion. Available at:

http://www.iea.org/co2highlights/co2highlights.pdf. Date of access: Jan. 28, 2014.

Kangas, A. S., & Kangas, J. (2004). Probability, possibility and evidence: approaches to consider risk

and uncertainty in forestry decision analysis. Forest Policy and Economics, 6(2), 169–188.

Kanzian, C., Holzleitner, F., Stampfer, K., & Ashton, S. (2009). Regional energy wood logistics–

optimizing local fuel supply. Silva Fennica, 43(1), 113–128.

Kazemi Zanjani, M., Ait-Kadi, D., & Nourelfath, M. (2010a). Robust production planning in a

manufacturing environment with random yield: A case in sawmill production planning.

European Journal of Operational Research, 201(3), 882–891.

Kazemi Zanjani, M., Nourelfath, M., & Ait-Kadi, D. (2010b). A multi-stage stochastic programming

approach for production planning with uncertainty in the quality of raw materials and demand.

International Journal of Production Research, 48(16), 4701–4723.

Kazemzadeh, N., & Hu, G. (2013). Optimization models for biorefinery supply chain network design

under uncertainty. Journal of Renewable & Sustainable Energy, 5(053125).

Keirstead, J., Samsatli, N., Pantaleo, A. M., & Shah, N. (2012). Evaluating biomass energy strategies

for a UK eco-town with an MILP optimization model. Biorefinery, 39, 306–316.

Kim, J., Realff, M. J., Lee, J. H., Whittaker, C., & Furtner, L. (2011a). Design of biomass processing

network for biofuel production using an MILP model. Biomass & Bioenergy, 35(2), 853–871.

Kim, J., Realff, M. J., & Lee, J. H. (2011b). Optimal design and global sensitivity analysis of biomass

supply chain networks for biofuels under uncertainty. Computers & Chemical Engineering,

35(9), 1738–1751.

134

Kostina, A. M., Guillén-Gosálbeza, G., Meleb, F. D., Bagajewiczc, M. J., & Jiménez, L. (2012). Design

and planning of infrastructures for bioethanol and sugar production under demand uncertainty.

Chemical Engineering Research & Design, 90(3), 359–376.

Labadie, J. W. (1997). Reservoir system optimization models. Water Resources Update, 108,

Universities Counsil on Water Resources. Available at:

http://ucowr.org/files/Achieved_Journal_Issues/

V108_A7Reservoir%20System%20Optimization%20Models.pdf. Date of access: Jan. 28,

2014.

Lehtikangas, P. (2001). Quality properties of pelletised sawdust, logging residues and bark. Biomass &

Bioenergy, 20(5), 351–360.

Leiras, A., Elkamel, A., & Hamacher, S. (2010). Strategic planning of integrated multirefinery

networks: A robust optimization approach based on the degree of conservatism. Industrial &

Engineering Chemistry Research, 49(20), 9970–9977.

Li, Z., & Floudas, C. A. Robust counterpart optimization: uncertainty sets, formulations and

probabilistic guarantees, http://focapo.cheme.cmu.edu/2012/proceedings/data/papers/ 030.pdf.

Date of access: Jan. 28, 2014.

Limayem, A., & Ricke, S. C. (2012). Lignocellulosic biomass for bioethanol production: Current

perspectives, potential issues and future prospects. Progress in Energy & Combustion Science,

38(4), 449–467.

Marvin, W. A., Schmidt, L. D., Benjaafar, S., Tiffany, D. G., & Daoutidis, P. (2012). Economic

optimization of a lignocellulosic biomass-to-ethanol supply chain. Chemical Engineering

Science, 67(1), 68–79.

Mccormic, K. (2011). Bioenergy systems and supply chains in Europe: conditions, coordination and

capacity. In T. M. Choi & T. C. E. Cheng (Eds.), Supply chain coordination under uncertainty

(Vols. 1-Book, 1-Section, pp. 545–562). Springer-Verlag New York Inc.

135

Mula, J., Poler, R., García-Sabater, J. P., & Lario, F. C. (2006). Models for production planning under

uncertainty: A review. International Journal of Production Economics, 103(1), 271–285.

Mulvey, J. M., Vanderbei, R. J., & Zenios, S. A. (1995). Robust optimization of large-scale systems.

Operations Research, 43(2), 264–281.

Nagel, J. (2000). Determination of an economic energy supply structure based on biomass using a

mixed-integer linear optimization model. Ecological Engineering, 16, Supplement 1, 91–102.

Palma, C. D., & Nelson, J. D. (2009). A robust optimization approach protected harvest scheduling

decisions against uncertainty. Canadian Journal of Forest Research, 39(2), 342–355.

Palma, C. D., & Nelson, J. D. (2010). Bi-objective multi-period planning with uncertain weights: a

robust optimization approach. European Journal of Forest Research, 129(6), 1081–1091.

Pannell, D. J. (1997). Sensitivity analysis of normative economic models: theoretical framework and

practical strategies. Agricultural Economics, 16(2), 139–152.

Petrovic, D. (2001). Simulation of supply chain behaviour and performance in an uncertain

environment. International Journal of Production Economics, 71(1-3), 429–438.

Petrovic, D., Roy, R., & Petrovic, R. (1999). Supply chain modelling using fuzzy sets. International

Journal of Production Economics, 59(1-3), 443–453.

Pérez-Fortes, M., Laínez-Aguirre J. M., Bojarski A. D., & Puigjaner L. (2014). Optimization of pre-

treatment selection for the use of woody waste in co-combustion plants. Chemical Engineering

Research and Design. In press. doi: http://dx.doi.org/10.1016/j.cherd.2014 .01.004. Date of

access: Jan. 28, 2014.

Ragauskas, A. J., Williams, C. K., Davison, B. H., Britovsek, G., Cairney, J., Eckert, C. A.,

Tschaplinski, T. (2006). The path forward for biofuels and biomaterials. Science 27, 311(5760),

484–489.

Ralevic, P., K. J., Smith C.T., Richardson J. (2008). IEA bioenergy task 31, country reports: Canada.

Canadian Forest Service of Natural Resources Canada.

136

Rauch, P., & Gronalt, M. (2011). The effects of rising energy costs and transportation mode mix on

forest fuel procurement costs. Biomass & Bioenergy, 35(1), 690–699.

Reche López, P., Jurado, F., Ruiz Reyes, N., García Galán, S., & Gómez, M. (2008). Particle swarm

optimization for biomass-fuelled systems with technical constraints. Engineering Applications

of Artificial Intelligence, 21(8), 1389–1396.

Rentizelas, A. A., Tatsiopoulos, I. P., & Tolis, A. (2009a). An optimization model for multi-biomass tri-

generation energy supply. Biomass & Bioenergy, 33(2), 223–233.

Rentizelas, A. A., Tolis, A. J., & Tatsiopoulos, I. P. (2009b). Logistics issues of biomass: The storage

problem and the multi-biomass supply chain. Renewable & Sustainable Energy Reviews, 13(4),

887–894.

Rockwood, D. L., Naidu, C. V., Carter, D. R., Rahmani, M., Spriggs, T. A., Lin, C. Segrest, S. A.

(2004). Short-rotation woody crops and phytoremediation: opportunities for agroforestry?

Agroforestry Systems, 61-62(1-3), 51–63.

Rozakis, S., & Sourie, J. C. (2005). Micro-economic modelling of biofuel system in France to

determine tax exemption policy under uncertainty. Energy Policy, 33(2), 171–182.

Saidur, R., Abdelaziz, E. A., Demirbaş, A., Hossain, M. S., & Mekhilef, S. (2011). A review on

biomass as a fuel for boilers. Renewable & Sustainable Energy Reviews, 15(5), 2262–2289.

Saltelli, A., Ratto, M., Andres, T., Campolongo, F., Cariboni, J., Gatelli, D. Tarantola, S. (2008). Global

sensitivity analysis. The primer. England: John Wiley & Sons.

a iba e -Aguilar, . E., Go le -Campos, . ., Po ce- r ega, . M., er a-Go le , M., & El-

Halwagi, M. M. (2011). Optimal planning of a biomass conversion system considering

economic and environmental aspects. Industrial & Engineering Chemistry Research, 50(14),

8558–8570.

Schmidt, J., Leduc, S., Dotzauer, E., Kindermann, G., & Schmid, E. (2010). Cost-effective CO2

emission reduction through heat, power and biofuel production from woody biomass: a

spatially explicit comparison of conversion technologies. Applied Energy, 87(7), 2128–2141.

137

Schmidt, J., Leduc, S., Dotzauer, E., Kindermann, G., & Schmid, S. E. (2009). Using Monte Carlo

simulation to account for uncertainties in the spatial explicit modeling of biomass fired

combined heat and power potentials in Austria,

ftp://ftp.boku.ac.at/pub/repecftpg/repecftp/RePEc/sed/ wpaper/432009.pdf. Date of access: Jan.

28, 2014.

Shapiro, J. F. (2004). Challenges of strategic supply chain planning and modeling. FOCAPO 2003

Special Issue, 28(6-7), 855–861.

Sharmaa, B., Ingallsb, R. G., Jonesa, C. L., Huhnkea, R. L., & Khanchi, A. (2013). Scenario

optimization modeling approach for design and management of biomass-to-biorefinery supply

chain system. Bioresource Technology, 150, 163–171.

Solyali, O., Cordeau, J.-F., & Laporte, G. (2012). Robust inventory routing under demand uncertainty.

Transportation Science, 46(3), 327–340.

Soyster, A. L. (1973). Convex programming with set-inclusive constraintsand applications to inexact

linear programming. Operations Research, 21(5), 1154–1157.

Stat::Fit®. Distribution fitting software. Retrieved from http://www.geerms.com. Date of access: Jan.

28, 2014.

Svensson, E., & Berntsson, T. (2011). Planning future investments in emerging energy technologies for

pulp mills considering different scenarios for their investment cost development. Energy,

36(11), 6508–6519.

Svensson, E., Strömberg, A.-B., & Patriksson, M. (2011). A model for optimization of process

integration investments under uncertainty. Energy, 36(5), 2733–2746.

Tay, D. H. S., Ng, D. K. S., & Tan, R. R. (2013). Robust optimization approach for synthesis of

integrated biorefineries with supply and demand uncertainties. Enviromental Progress and

Sustainable Energy. 32(2), 384–389.

Van Dyken, S., Bakken, B. H., & Skjelbred, H. I. (2010). Linear mixed-integer models for biomass

supply chains with transport, storage and processing. Energy, 35(3), 1338–1350.

138

Vera, D., Carabias, J., Jurado, F., & Ruiz-Reyes, N. (2010). A honey bee foraging approach for optimal

location of a biomass power plant. Applied Energy, 87(7), 2119–2127.

Vose, D. (2008). Risk analysis: a quantitative guide (Vol. 3rd). England: Wiley. com.

Wahlund, B., Yan, J., & Westermark, M. (2004). Increasing biomass utilisation in energy systems: A

comparative study of CO2 reduction and cost for different bioenergy processing options.

Biomass & Bioenergy, 26(6), 531–544.

Wetterlund, E., & Söderström, M. (2010). Biomass gasification in district heating systems – The effect

of economic energy policies. Special Issue of the IGEC-IV, the 4th International Green Energy

Conference (IGEC-IV), Beijing, China, October 20–22, 2008 Special Issue of the First

I er a io al Co fere ce o Applied E ergy, ICAE’09, Hong Kong, January 5–7, 2009, 87(9),

2914–2922.

Wiesemann, W., Kuhn, D., & Rustem, B. (2012). Robust resource allocations in temporal networks.

Mathematical Programming, Series A, 135(1-2), 437–471.

You, F., Wassick, J. M., & Grossmann, I. E. (2009). Risk management for a global supply chain

planning under uncertainty: models and algorithms. AIChE, 55(4), 931–946.

Zhu, J. Y., & Pan, X. J. (2010). Woody biomass pretreatment for cellulosic ethanol production:

Technology and energy consumption evaluation. Special Issue on Lignocellulosic Bioethanol:

Current Status and Perspectives, 101(13), 4992–5002.