You are on page 1of 156

UNIVERSITY OF CALGARY

Upgrading of HydrofactionTM Biocrude Through Catalytic Hydrotreating

By

Anderson Anibal Montanez Rincon

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF MASTER OF SCIENCE

GRADUATE PROGRAM IN CHEMICAL ENGINEERING

CALGARY, ALBERTA

SEPTEMBER, 2019

© Anderson Anibal Montanez Rincon 2019


Abstract

Alternative transportation fuels are critically required nowadays due to environmental concerns
related to the rising of greenhouse gas emissions and the depletion of petroleum reserves. In this
regard, advanced biofuels, commonly known as a second-generation biofuel, are gaining attention
as promising candidates to replace the high polluting conventional transportation fuels or to fulfill
drop-in fuel specifications. Abundant biomass coming from forest residues has been recently
processed applying HydrofactionTM, which is a hydrothermal liquefaction technology to produce
renewable biocrude. Even though it is a high-quality biocrude compared with other second-
generation biofuels, an upgrading step is required, and the catalytic hydrotreating technology was
selected for this investigation. Reduction of oxygen content, acidity and viscosity of
HydrofactionTM biocrude are the primary goals in this study.
Two independent hydrotreating campaigns were accomplished to study the effect of operational
conditions during the upgrading of renewable HydrofactionTM biocrudes in the presence of
commercial hydrotreating catalysts. In both campaigns, the performance of the catalysts over time
was studied as well as the thermal effects on the upgrading of biocrudes; it was accomplished by
performing experiments in the bench hydrotreater pilot plant without catalysts. After performing
experiments using two downflow reactors packed with commercial NiMo catalysts, promising
results were achieved. The oxygen content of the upgraded sample was totally removed, the total
acid number was reduced to 0.06 mgKOH/g oil (99.9%), the viscosity was decreased to 6 cP
(99.5%), the H/C molar ratio was highly improved to 1.71 as well as the heating value up to 43.7
MJ/kg. The baseline for comparison was the initial properties of the biocrude. After upgrading, a
highly improved distillation curve (obtained from gas chromatography, high temperature
simulated distillation) showed low vacuum residue fraction (>550 o C) around 5%, and improved
diesel fraction around 50% in the range 190 to 343 °C were obtained positioning this upgraded
biocrude as a potential pathway for transportation fuels production. Finally, the insights of this
research will allow the establishment of optimal operative conditions for upgrading renewable
biocrudes generated from woody residues.

II
Acknowledgements

I would like to acknowledge the input and invaluable contributions of a number of people who in
one way or another contributed to make this investigation possible:

My supervisor Dr. Matthew A. Clarke from the University of Calgary and Co-supervisor Dr.
William C. McCaffrey from the University of Alberta for all the support, guidance and time spent
in every step of this research. Thank you for trusting me and giving me the opportunity to fulfill
my MSc.

Special thanks to Steeper Energy for the funding, and for the guidance and technical advice
provided for a successful culmination of the research project. Also, thanks to Mitacs Accelerate
for the funding provided.

Dr. Gonzalo Rocha for teaching me many things about upgrading and for the very good job
performed at the beginning of the project.

Furthermore, I express my sincere gratitude to Dr. Parsa Haghighat for being my friend and an
excellent colleague during my stay in Edmonton. Without his knowledge and help, this project
would not be successful.

My old friends in Colombia (Fernando, Diego, Jonathan, Oscar, Javier O., Javier G., Jorge, David,
Daniel Ricardo, etc.) and the new ones in Calgary and Edmonton, who always were there to
encourage me in the difficult times.

Finally, from the deepest of my being, I express the sincerest gratitude to all my family specially
to my mother Clara, my sister Geraldine, my nephew Samuel, my niece Salome, and to the most
important person in my life, my beautiful wife Sandra Liliana for being my strength, inspiration
and unconditional support. I thank God for placing in my path, all the wonderful people mentioned,
and for giving me the strength to conquer this goal.

III
Dedication

Dedico este sueño hecho realidad a la mujer con quien felizmente pasare el resto de mis dias, a
mi hermosa Sandra Liliana. A quien admiro por su inteligencia, constancia, perseverancia y
valentia. Con ella comparto mis tristezas y alegrias, ademas es por ella que trato cada dia ser una
mejor persona.

Tambien dedico este importamte logro en mi vida a quien le debo todo, a mi madre Clara Ibeth.
De de ella he aprendido que los sacrificios siempre son recompenzados, y que las cosas mas
importantes de un ser humano son la sencillez, y la humildad.

Quiero dedicar este logro a mi hermana y a mis sobrinos, quienes siempre estan en mi mente y en
mi Corazon.

Finalmente dedico este logro a Dios y la Virgen Maria.

IV
Table of Contents

Abstract .......................................................................................................................................... II

Acknowledgements ..................................................................................................................... III

Dedication .................................................................................................................................... IV

Table of Contents ......................................................................................................................... V

List of Tables ............................................................................................................................ VIII

List of Figures .............................................................................................................................. IX

Nomenclature ............................................................................................................................. XII

Chapter 1: Research Background ............................................................................................. 15

1.1. Research Objectives ..................................................................................................... 18

1.2. Thesis Outline ................................................................................................................... 19

Chapter 2: Literature Review .................................................................................................... 21

2.1. Biofuel Definition ............................................................................................................. 21

2.2. Biofuels and Their Energy Potential .............................................................................. 22

2.3. Biofuel Feedstocks ........................................................................................................... 24

2.3.1. First-generation feedstocks. ..................................................................................... 24

2.3.2. Second-generation feedstocks. ................................................................................. 27

2.3.3. Third-generation feedstocks. ................................................................................... 30

2.3.4. Fourth-generation feedstocks .................................................................................. 32

2.4. Production Methods for Biofuels .................................................................................... 33

2.4.1. Pretreatment .............................................................................................................. 33

2.4.2. Reaction ..................................................................................................................... 35

2.4.4. Upgrading .................................................................................................................. 49

V
Chapter 3: Experimental Methods ............................................................................................ 57

3.1. Materials .......................................................................................................................... 57

3.2. Apparatus ......................................................................................................................... 59

3.2.1. Description of the bench hydrotreater pilot plant ................................................. 59

3.2.2. Analytical equipment used in the characterization of biocrudes and upgraded


samples. ................................................................................................................................ 64

3.3. Experimental Procedure ................................................................................................. 67

3.3.1. Hydrotreating unit operation................................................................................... 67

3.3.2. Reactor assembly. ..................................................................................................... 70

3.3.3. Catalyst activation. ................................................................................................... 73

3.4. Experimental Design ........................................................................................................ 74

3.4.1. Campaign #1: upgrading of biocrude-1 in the presence of pre-sulfided CoMo and
NiMo catalysts. .................................................................................................................... 74

3.3.2. Campaign #2: upgrading of HydrofactionTM biocrude-2 using two down-flow


reactors in series in the presence of NiMo Catalysts. ...................................................... 77

Chapter 4: Upgrading of Biocrude-1 through catalytic hydrotreating in the presence of


commercial CoMo and NiMo catalysts (campaign #1) ............................................................ 79

4.1. Commissioning and equipment verification .................................................................. 79

4.2. Characterization of HydrofactionTM biocrude-1. ......................................................... 81

4.3. Upgrading of Biocrude #1 in the Presence of CoMo Catalyst ..................................... 81

4.3.1. Effect of temperature on catalytic hydrotreating of HydrofactionTM biocrude-1.


............................................................................................................................................... 81

4.3.2. Space velocity screening over the catalytic hydrotreating of Hydrofaction TM


biocrude-1. ........................................................................................................................... 88

4.3.3. Pressure and H2/Oil ratio screening on catalytic hydrotreating of HydrofactionTM


biocrude-1. ........................................................................................................................... 91

4.3.4. Study of catalyst activity. ......................................................................................... 93


VI
4.4. Upgrading of Partially Upgraded Biocrude in the Presence of NiMo Catalyst ......... 95

4.5. Thermal Experiments Performed Without Catalysts .................................................. 98

Chapter 5: Two-Stage Catalytic Hydrotreating of HydrofactionTM Biocrude-2 in the


presence of commercial NiMo Catalysts (campaign #2) ....................................................... 101

5.1. Characterization of HydrofactionTM biocrude-2 ........................................................ 101

5.2. Effect of temperature in catalytic hydrotreating of HydrofactionTM biocrude-2 (two


reactors in series process) ..................................................................................................... 102

5.3. Effect of Space Velocity on Catalytic Hydrotreating of HydrofactionTM Biocrude-2


(two reactors in series process) ............................................................................................ 109

5.4. Effect of H2/oil Ratio on Catalytic Hydrotreating of HydrofactionTM Biocrude-2


(two reactors in series process) ............................................................................................ 111

5.5. Effect of Hydrogen Flow Pattern over the Catalytic Hydrotreating of


HydrofactionTM Biocrude-2 (two reactors in series process) ............................................ 113

5.6. Evaluation of Catalyst Deactivation ............................................................................. 115

Chapter 6: Conclusions and Recommendations .................................................................... 117

6.1. Conclusions ..................................................................................................................... 117

6.2. Recommendations .......................................................................................................... 121

REFERENCES .......................................................................................................................... 122

Appendix A: Mass Balance Calculation ................................................................................. 138

Appendix B: Standard Operational Procedure ..................................................................... 140

Hazzard Assessment: ................................................................................................................ 150

Appendix C: Copyright ............................................................................................................ 155

VII
List of Tables

Table 2.1. First-generation feedstocks for bioethanol and biodiesel production. ........................ 25
Table 2.2. Properties of typical biocrudes and a Canadian bitumen ............................................ 48

Table 3.3. Catalysts tested for the upgrading of HydrofactionTM biocrudes. ............................... 58
Table 3.4. Different chemicals used in different activities during the investigation ................... 58
Table 3. 5. Gases used during the investigation ........................................................................... 59
Table 3.6. Experimental plan followed for the upgrading of the biocrude-1 in the presence of pre-
sulfided CoMo catalyst. ................................................................................................................ 76
Table 3.7. Experimental protocol followed for the upgrading of the biocrude-1 in the presence of
pre-sulfided NiMo catalyst. .......................................................................................................... 76

Table 4.1. Characterization of HydrofactionTM Biocrude-1 ......................................................... 81


Table 4.2. Effect of temperature on the upgrading of Biocrude-1. Fixed conditions: P=95 bar,
WHSV=0.3 h-1, H2/oil=900 Nm3/m3oil ........................................................................................ 82
Table 4.3. Effect of space velocity on properties of upgraded products. Fixed conditions: T=350oC,
P=95 bar and H2/oil =1300 Nm3/m3oil. ........................................................................................ 88
Table 4.4. Effect of pressure over the upgrading of HydrofactionTM biocrude-1. Fixed parameters:
WHSV=0.3 h-1, H2/oil = 900 Nm3/m3oil. ..................................................................................... 92
Table 4.5. Effect of H2/oil ratio over the upgrading of HydrofactionTM biocrude-1. Fixed
parameters: WHSV=0.3 h-1, H2/oil = 900 Nm3/m3oil. .................................................................. 93
Table 4.6. The comparison of different mass balances taken at different times during experiment
# 3 of the experimental protocol. Fixed conditions: T= 350 oC, P= 95 bar, WHSV= 0.3 h-1 and
H2/Oil= 900 Nm3 H2/m3 oil. .......................................................................................................... 94
Table 4.7. Reproducibility results obtained from step 5 and step 7 of upgrading in the presence of
CoMo catalyst. Fixed conditions: T= 350 oC, P= 80 bar, WHSV=0.3 h-1 and H2/oil= 900 Nm3
H2/m3 oil........................................................................................................................................ 95
Table 4.8. Properties of biocrude-1 and upgraded samples after first and second upgrading steps.
Fixed conditions: WHSV = 0.3 h-1, P = 95 bar and H2/oil = 900 Nm3/m3 oil .............................. 96

VIII
Table 4.9. Temperature effect over the upgrading of biocrude-1. Analysis performed without
catalysts. Fixed conditions: P=95 bars, H2/Oil =900 Nm3/m3, and HWVS= 0.3 h-1 .................... 99

Table 5.1. Characterization of HydrofactionTM Biocrude-2 ....................................................... 102


Table 5.2. Effect of temperature in reactors 1 and 2 over the upgrading of biocrude-2. Fixed
conditions: P=90 bars, WHSV=0.3 h-1, H2/oil=900 Nm3/m3oil ................................................. 103
Table 5.3. Effect of space velocity on properties of upgraded products. Fixed conditions: TR1
=320 oC P=90 bar, and H2/oil = 900 Nm3/m3oil. ........................................................................ 110
Table 5.4.Effect of H2/oil ratio on properties of upgraded products. Fixed conditions: TR1 =320
o
C P=90 bar, and WHSV= 0.2 h-1. .............................................................................................. 112
Table 5.5. Effect of hydrogen flow patterns over the catalytic hydrotreating of biocrude-2. Fixed
conditions: TR1=320 oC TR2=370 oC P=90 bar, HWSV=0,3 h-1 H2/Oil= 900 Nm3/m3 oil ....... 114
Table 5.6 Evaluation of catalysts deactivation. Fixed conditions: TR1=320 oC TR2=350 oC P=90
bar, HWSV=0,3 h-1 H2/Oil= 900 Nm3/m3 oil .............................................................................. 115

Table B. 1. Hazzard Assessment form used by all University of Alberta Deparments and units.
..................................................................................................................................................... 153

List of Figures

Figure 2.1. Feedstocks and pathways of different biofuels production. Adapted from [36]. ...... 23
Figure 2.2. Chemical wood components adapted from [84]. ....................................................... 28
Figure 2.3. Structure of lignocellulosic biomass, adapted from [89], [90]. ................................ 30
Figure 2.4. Pathway to obtain biofuels through fermentation. ..................................................... 37
Figure 2.5. General scheme of the pyrolysis process adapted from [132]. .................................. 40
Figure 2.6. Main steps in gasification adapted from [138] .......................................................... 41
Figure 2.7. Simplified process diagram about the production of HydrofactionTM Biocrude.
Adapted from [21]......................................................................................................................... 44
Figure 2.8. Main reactions of HydrofactionTM process, adapted from [21] ................................. 47
Figure 2.9. Main reactions in hydrotreating processes of biocrudes, adapted from [183]. .......... 54

IX
Figure 3.1. Simplified schematic diagram of the catalytic hydrotreater (double lines represent the
heated and insulated lines in the setup)......................................................................................... 60
Figure 3.2. Schematic of the packed tubular reactor ................................................................... 62
Figure 3.3. OMEGA multipoint thermocouple probe. Adapted from [191]. ............................... 63
Figure 3.4. Configuration of the packed bed reactors with CoMo and NiMo catalysts used in the
first set of experiments. The First and second stage were performed independently. .................. 71
Figure 3.5. Configuration of the packed reactors R1 and R2 installed in series mode in the bench
hydrotreater pilot plant. (Second set of upgrading experiments). * Ratio of Catalyst D and E and
the bed configuration cannot be disclosed. ................................................................................... 72
Figure 3.6. Operational conditions screened during the catalytic hydrotreating in the presence of
CoMo and NiMo catalysts. First and second stage scenarios were performed independently (stage
one with CoMo and stage two with NiMo). ................................................................................. 75
Figure 3.7. Operational conditions screened during the catalytic hydrotreating performed in series
mode in the presence of activated NiMo catalysts........................................................................ 77
Figure 3.8. Experimental conditions proposed for the upgrading of the biocrude-2 in the presence
of pre-sulfided NiMo catalysts...................................................................................................... 78

Figure 4.1. Distribution of gas compounds at different reaction temperatures. Fixed conditions
P=95 bar, H2/oil =900 Nm3/m3oil, and HWSV= 0.3 h-1 ............................................................... 83
Figure 4.2. TGA Analysis of Hydrofaction biocrude at different temperatures. Left hand-side:
weight loss. Right hand-side: derivative weight loss. Fixed conditions: P=95 bar, H2/oil =900
Nm3/m3oil, and HWSV= 0.3 h-1 ................................................................................................... 84
Figure 4.3. Boiling point distribution (SimDist) results of upgraded samples at different
temperatures. Fixed conditions P=95 bar, H2/oil =900 Nm3/m3oil, and HWSV= 0.3 h-1............. 85
Figure 4.4. FTIR analysis for renewable biocrude-1 and upgraded products at different
temperatures. Fixed conditions: 95 bar, H2/oil=900 Nm3/m3oil, and HWSV= 0.3 h-1 ................. 87
Figure 4.5. Selectivity of gas products at different space velocities. Fixed Conditions: T=350oC,
P=95 bar and H2/oil =1300 Nm3/m3oil. ........................................................................................ 89
Figure 4.6. Hydrogen consumption at different space velocities. Fixed Conditions: T=350oC,
P=95 bar and H2/oil =1300 Nm3/m3oil. ........................................................................................ 90

X
Figure 4.7. Boiling point distribution (Simdis) of biocrude-1 and upgraded samples after two
upgrading stages. Fixed conditions: P=95 bar, H2/oil =900 Nm3/m3oil, and HWSV= 0.3 h-1 ..... 97

Figure 5.1. Selectivity of compounds in the gas phase at different temperature in reactor 2. Left-
hand side: produced gases, right-hand side: remaining hydrogen in the exit gases. Fixed conditions:
TR1=320 °C, P=90 bars, WHSV=0.3 h-1, H2/oil=900 Nm3/m3oil.............................................. 105
Figure 5.2. Products yield and hydrogen consumption by the effect of different temperatures in
reactor 2. Fixed conditions: TR1=320 °C, P=90 bars, WHSV=0.3 h-1, H2/oil=900 Nm3/m3oil. 106
Figure 5.3. Products yield and hydrogen consumption at severity conditions, higher temperatures
in both reactors. Fixed conditions: P=90 bars, WHSV=0.3 h-1, H2/oil=900 Nm3/m3oil ............ 107
Figure 5.4. Boiling point distribution (Simdis) results of upgraded samples at different
temperatures in reactors 1 and 2. Fixed conditions: P=90 bars, WHSV=0.3 h -1, H2/oil=900
Nm3/m3oil ................................................................................................................................... 109
Figure 5.5. Products yield and hydrogen consumption at different space velocities. Left hand-side
figure (A) TR2=350 °C; right-hand side (B) TR2= 370 °C. Fixed conditions: TR1= 320 C, P=90
bars, and H2/oil=900 Nm3/m3oil ................................................................................................. 111
Figure 5.6. Products yield and hydrogen consumption at H 2/oil ratios. Left hand-side figure
TR2=350 °C; right-hand side TR2= 370 °C. Fixed conditions: TR1= 320 C, P=90 bars, and HWSV
=0.2 h-1 ........................................................................................................................................ 113
Figure 5.7. Sketch of the hydrogen flow patterns investigated in the catalytic hydrotreating of
biocrude-2. .................................................................................................................................. 114

XI
Nomenclature

Abbreviations
ADP Adenosine diphosphate
ASTM American Society for Testing and Materials
ATP Adenosine triphosphate
BP Beyond Petroleum (Formerly British Petroleum)
BPV Back pressure regulator
CB Catalytic bed
CoMo Cobalt-Molybdenum
FID Flame ionization detector
FTIR Fourier Transform Infra-Red Spectroscopy Analysis
IPCC Intergovernmental Panel on Climate Change
GC Gas Chromatography
GFM Gas flow meter
GHG Greenhouse Gases Emissions
Gt Gigatonne
H/C Hydrogen to carbon ratio
HTL Hydrothermal liquefaction
HHV Higher heating value
HAD Hydro-dearomatization
HDO Hydro-deoxygenation
HDM Hydro-demetallization
HDN Hydro-denitrogenation
HDS Hydro-desulfurization
HPG High-pressure gauge
HVGO Heavy vacuum gas oil
HYC Hydrocracking
HTK Heavy liquid tank
ID Interior diameter
IR Infrared spectroscopy
g/L Grams per liter

XII
L/ha Liter per hectare
LTK Light liquid tank
LVGO Light vacuum gas oil
Nm3 Normal cubic meters
MJ Mega joules
MFC mass flow controller
Mt Megatonne
Mpa Megapascal
NiMo Nickel-Molybdenum
PI Pressure transducer
P&ID Piping and instrumentation diagram
pH Hydrogen ion concentration
R1 Reactor one
R2 Reactor two
R&D Research and Development
RD Rupture disc
RV Gas regulator
SASOL South Africa Synthetic Oil Liquid
SimDist Simulated Distillation
SDGs Sustainable Development Goals
TAN Total acid number
TCD Thermal Conductivity Detector
TGA Thermogravimetric Analysis
TIC Temperature controller
TK Tank
TRS Total reducing sugars
TWV Three-way valve
1G First generation
2G Second generation
3G Third generation
4G Fourth generation

XIII
mg Milligrams
UNFCCC United Nations Framework Convention on Climate Change
USDA United States Department of Agriculture
WGA Water gas shit
WHSV Weight hour space velocity
wt% Weight percent

Greek Letter

ß Beta (Beta-D-glucose)
γ Gamma (In gamma alumina catalysts support)
µ Mu (micro)

Subscripts
A On dry basis
B Oxygen by difference

XIV
Chapter 1: Research Background

The demand for new energy sources continues to increase due to environmental concerns related
to the rising of greenhouse gases emissions (GHG) [1], [2] the depletion of petroleum reserves [3]
and a continuously growing energy demand [4], [5], [6]. Even though historically significant
advances in science have been prosperous for humanity, they have also led to environmental
concerns. For example, internal combustion engines for transportation and power generation as
well as petrochemical processes are high polluting processes that consume millions of barrels of
fossil fuels every day which in turn, is believed to be accelerating the pace of global warming.

According to the last Intergovernmental Panel on Climate Change report (IPCC), the global
average temperature rise caused by human activities including the burning of fossil fuels has
already reached approximately 1°C above the global warming reported at the pre-industrial period.
Thus, if the current warming trend continues (business as usual) it is predicted that the temperature
levels reach 1.5° C between 2030 and 2050 taking as a baseline the temperature reported at the
preindustrial period (1850-1900). This period, just before the industrial activities started, was the
time since when a drastic and abnormal increase in global temperature was recognized [7]. In 2014
the anthropogenic GHG global emissions caused by various industrial energy sectors were around
49 Gt of CO2 [1]. For example, electricity and heat production sector contribute to 25%,
agriculture, forestry and other land use deforestation sectors with 24%, the transportation sector to
14%. Furthermore, buildings contributed to 6.4%, and other energy sectors to 9.6%. The remaining
emissions were indirect CO2 emissions by around 21%. For the specific case of Canada, according
to the national inventory report submitted to the United Nations Framework Convention on
Climate Change (UNFCCC) in 2016, fuel combustion, the transportation sector, and gas emissions
from leaks or irregular releases from industrial activities in the Canadian energy sector produced
572 Mt of greenhouse gas emissions where 317 Mt were CO2 emissions. In addition, the Canadian
GHG emissions projected by 2030 will be around 742 Mt [8].

15
Several environmental regulations regarding greenhouse gases and CO2 emissions were
implemented by several European Union countries, Canada, and the United States. The objective
is to support the production of renewable crude oil to be used as drop-in biofuels chemically
indistinguishable from conventional petroleum fuels. Also, many European countries are
committed to increasing biofuel blending with transport fuels to 10% by 2020 [9]. India also issued
a policy, aiming to reach the blending of biofuel in transport fuels to 20 vol% by 2017 [10]. Based
on the Canadian regulation, the produced or imported gasoline should contain 5 vol% of renewable
biofuel, which is estimated to reduce the GHG emission of around 1 million tons of CO 2 per year.

Furthermore, in the Pan-Canadian Framework, a specific goal was set to decrease by 30 % in 2030
the CO2 emissions reported in 2005, it will be approximately 51 Mt of CO 2 emissions annually
[11]. To accomplish this target, one option is to replace partially or totally, the use of fossil fuels
with liquid renewable biofuels from the abundant biomass resources [12], [13].

Liquid biofuels can be broadly categorized as first, second, third, and fourth-generation biofuels.
First-generation biofuels are produced from (1G) agricultural crops, which can also be grown for
food. Using these crops as feedstock can lead to the production of biofuels such as bioethanol and
biomethane (biogas) produced through fermentation, or biodiesel (bio-esters) via the
transesterification process [14], [15], [16].

On the other hand, second-generation (2G) biofuels are produced from non-food biomasses or
crops such as waste cooking oil, various kinds of organic wastes, vegetative grasses, seaweed
(algae), lignocellulosic crops or by-products from sugar cane, bagasse, cereal straw, animal fats or
recycled greases, among others. 2G biofuels offer some advantages over 1G biofuels such as the
feedstocks used are non-food biomasses and the biofuel produced is drop-in, which means that it
can be blended in different proportions with fossil fuels or in the best case scenario they can be
used directly without blending [17]. Maybe one of the most important second-generation
feedstocks are the lignocellulosic woody feedstocks [6] which include any wood residue from pulp
and paper industries, tree branches, leaves or sawdust. Animal manure has been used as well for
the production of second-generation biocrudes.

The main routes to produce 2G biofuels are biochemical, thermochemical and hydrothermal
liquefaction (HTL) processes. Biochemical routes include enzymatic hydrolysis and fermentation
processes [18] Hydrothermal liquefaction (HTL), pyrolysis and gasification technologies are

16
thermochemical processes for biofuels production. Principally, the HTL technology converts
abundant/renewable lignocellulosic biomasses into a high-quality crude oil that can be blended
with fossil fuels in higher proportions, unlike first-generation biofuels which are not suitable for
aviation or long distances transportation and cannot be blended with commercial fuels such as
diesel and gasoline in high percentages.

Third-generation biofuels (3G) are produced from algae or microalgae feedstocks. This technology
is an attractive alternative because the 3G feedstocks have higher photosynthetic efficiency,
between 3 to 8% compared to crops which have around 0.5% efficiency [19], [20], and the
percentage of biofuel produced with 3G feedstocks are up to 10 times higher compared with 1G
or 2G biofuels production. Overall, third-generation biofuels are a promising technology in which
research and development are ongoing. Also, 4G are biofuels produced from 3G feedstocks
genetically modified.

Of the 1G, 2G, 3G, and 4G biofuels the 1G fuels technologies have been well-researched. Amongst
the 2G fuels, a promising emerging production process is HTL. In particular, Hydrofaction TM
renewable crude-oil is produced using hydrothermal liquefaction technology (HTL). Steeper
Energy is the proprietor of the HydrofactionTM HTL technology which utilizes supercritical water
and homogeneous catalysts to transform lignocellulosic feedstocks into high heating value liquid
biocrudes [21],[22].

HydrofactionTM biocrude has, in general, some undesirable properties such as a high oxygen
content around 10 wt. %, which is high compared to fossil fuels [23]. Also, a high total acid number
(TAN) around 62 mg KOH/g and high viscosity; the biocrude also contains some thermally
unstable components that can trigger gum formations [21], these drawbacks of the feedstock are
the result of the depolymerization process of the forestry residue through the HydrofactionTM
process. Hence, to convert the renewable HydrofactionTM crude oil to value-added finished fuels,
diesel/jet blendstocks or a drop-in crude oil, an upgrading step must be performed. To transform
the biocrude into a high value-added fuel with improved distillation curve, the catalytic hydrogen
treatment (also known as hydrotreating) is the pathway of choice to upgrade the Hydrofaction TM
biocrude. Hydrotreating is a mature technology widely applied for the upgrading of conventional
fossil fuels that can be used to upgrade unconventional biofuels as well without significant
technical modifications of the hydrotreater units. This technology includes high temperatures,

17
moderate to high H2 pressures and the presence of metal catalysts to upgrade petroleum products
originally by hydrogenating unsaturated hydrocarbons and removing undesirables compounds
such as carboxylic acids, phenols, ketones among others through hydro-deoxygenation (HDO),
and decarboxylation reactions [24].

Studies to date have been performed on HDO of biofuels with NiMo and CoMo catalyst with
alumina support (γ-Al2O3) [25], [26], [27] and upgrading in the presence of different
heterogeneous catalysts [28], [29]. Among these studies, most of the published works have been
focused on the upgrading of pyrolysis bio-oils [30]. However, very few studies have been
performed upgrading of HTL biofuels. Based on the background presented and the lack of
technical data about the upgrading of HTL biofuels, the present research project aims to fill the
knowledge gaps mentioned by following the objectives presented below.

1.1. Research Objectives

In this study, the upgrading of HTL HydrofactionTM renewable biocrudes was performed in a
continuous bench hydrotreater pilot plant provided with two downflow packed-bed reactors in the
presence of commercial NiMo and CoMo catalysts. The unit was designed and constructed at the
University of Alberta to acquire reliable data to scale up the upgrading of HydrofactionTM
renewable biocrude. This study also includes the screening of several operational conditions such
as temperature, pressure, space velocity, H2/Oil ratio and the evaluation of different hydrotreating
commercial catalysts.

The objectives of this research study are to:

1. Build and commission the hydrotreater pilot plant, to understand the dynamics of the unit by
performing cold and hot tests and develop all the safety protocols such as standard operating
procedures and hazard assessments.

2. Investigate and optimize at continuous bench-scale, the upgrading of HydrofactionTM


renewable biocrude produced from challenging feedstocks such as woody residues.

18
3. Achieve the upgrading of Hydrofaction™ renewable biocrude, using sulphide NiMo and
CoMo commercial catalysts; as well as to find the best operational conditions performing the
screening and optimization of temperature, pressure, space velocity and H 2/oil ratio in order
to transform the biocrudes into finished or close as possible improved fuels for transportation
such as diesel, jet fuel, marine or propulsion engines or as drop-in biocrudes.

4. Improve the physio-chemical properties of the biocrude and reduce its high oxygen content,
high viscosity and aromaticity to produce a high upgraded product to be used as blendstocks
for transportation fuels.

5. Perform experiments (without catalysts) in the bench hydrotreater to acquired information


about the thermal effects (cracking reactions). With this data, it will be possible to verify the
performance of the catalysts used in the hydrotreating experiments.

6. Carry out studies at different operational times to track the catalysts deactivation.

7. Perform a comprehensive characterization of the upgraded Hydrofaction™ renewable


biocrudes after each set of experiments carried out through the investigation.

1.2. Thesis Outline

This thesis is divided into 6 chapters, as follows:

Chapter 1 introduces a general overview of biofuels, particularly about second-generation


HydrofactionTM renewable biocrude as an eco-friendly alternative to replace conventional fuels
after performing the upgrading step. Also, it introduces the main problem of the research and the
main objectives.

Chapter 2 presents the literature review about biofuels, renewable feedstocks with the main topics
about 1G, 2G, 3G, and 4G feedstocks and biofuels produced from these feedstocks, particularly
about the second generation HydrofactionTM biocrude. A review of biofuel production methods

19
and the upgrading of biofuels is presented with emphasis about hydrotreating. The problem
statement is included in this chapter.

Chapter 3 is the experimental methods chapter; it includes the description and operation procedure
of the bench hydrotreater pilot plant, the materials used in the investigation, the reactor assembly
procedure as well as a brief mention about the catalyst activation. Also, in this chapter, the
experimental protocols designed for each hydrotreating campaign are included.

Chapter 4 contains de results of the commissioning and equipment verification and the results
obtained from the catalytic hydrotreating campaign #1. In this campaign, two independent sets of
experiments were performed, the first one in the presence of pre-sulfided and reactivated CoMo
catalysts, and the second one in the presence of NiMo catalyst. In both campaigns, the screening
of conditions was accomplished. Additionally, a study of the catalyst activity was included, as well
as the thermal experiments performed without catalysts

Chapter 5 presents the results of the upgrading of biocrude-2 through a process performed with
two reactors in series packed with three different commercial NiMo catalysts. The results of
screening different operational conditions are presented. Finally, the results of the study of
catalysts deactivation are included.

Chapter 6 contains the main conclusions of both campaigns achieved and the recommendations
for future studies.

20
Chapter 2: Literature Review

In this chapter, the fundamentals and the relevant literature regarding biofuels, renewable
feedstocks, conversion pathways, and upgrading technologies are reviewed. Concepts about
hydrothermal liquefaction are discussed including the Steeper Energy HydrofactionTM technology
for production of biocrudes from woody feedstocks. Furthermore, a review about upgrading is
included because the focus of this thesis is the upgrading of HydrofactionTM biocrudes trough
catalytic hydrotreating. However, mention of first, third and fourth generation biofuel feedstocks
and processing technology will also be included in this section. Finally, the problem statement is
presented at the end of this chapter.

2.1. Biofuel Definition

Biofuels are solid, liquid or gaseous energy fuels derived from living, or recently living, plant or
animal matter. Among them, liquid biofuels such as bioethanol or biodiesel are predominantly
used for the transportation sector and produced from different renewable feedstocks [31]. Solid
biofuels such as firewood or pressed pellets can be burned in stoves or furnaces for heat production
or can be used in boilers for thermal energy generation [32]. Gaseous fuels are used for heating
and power generation; the most important gaseous fuels are natural gas, synthesis gas (syngas)
and pure hydrogen [33].
As transportation fuels, biofuels are almost as old as cars; the first car produced by Henry Ford
was powered with ethanol [34]; the first diesel engine invented by Rudolf Diesel worked with
peanut oil in an exhibition in Paris 1900 [13]. Unprocessed biomaterials like wood or manure have
been used for cooking and to warm up houses for centuries [35]. These unprocessed biofuels
played an essential role in manufacturing and transportation because many kinds of machinery
were powered by steam produced from burning dry wood.
Biofuels are classified into two main groups; the first ones are related to biomaterials or
unprocessed feedstocks such as animal wastes, firewood or wood chips, forest crops residues and
landfill gases such as carbon dioxide and methane [36]. The second group, are biofuels produced
from processed feedstocks, which depending on their composition, are classified into three main
subgroups; first-generation biofuels (1G) are those which are derived from food crops and animal

21
wastes, second-generation biofuels (2G) are those derived from lignocellulosic biomass. Third-
generation biofuels (3G) are those derived from microalgae and microbes [37], and fourth-
generation biofuels are those produced with aquatic feedstocks (microalgae) genetically modified
with the purpose to enhance biofuel production.
The production of biofuels is a continuously evolving area of research. The production of liquid
biofuels from alternative fuel sources is the scope of intense research [38], [39], [2], [40] because
they have the potential to reduce demand for conventional fossil fuels as well as to address
environmental concerns and, thereby, they can contribute to poverty reduction in developing
countries [41]. The growing amount of research works about biofuels published in the past
decades, indicates an increasing interest in alternative renewable energies in many countries
around the world [42]. Much of this research interest arises from the perception that biofuels are a
greenhouse gas neutral fuel source. This means that the rate of atmospheric carbon dioxide
absorbed as plants grow, equilibrates the amount generated when they are burned. Thus, the
greenhouse gas emissions’ lifecycle can be reduced up to 80%. Finally, biofuels typically contain
much less harmful environmental contaminants than fossil fuels and are biodegradable; for
example in the case of a spill in the environment [43].
The biofuels’ production is accomplished in a series of steps: first, in most cases, a pre-treatment
step is required, next, a reaction step is performed. One common drawback of biofuels is its high
oxygen content. As a result, an upgrading step of the biofuels is carried out in order to meet
specifications for liquid transportation fuels or drop-in biofuels [44], [45].

2.2. Biofuels and Their Energy Potential

The need for different energy alternatives is the result of growing energy demand, especially in
remote locations and/or emerging countries with limited fossil fuels reserves. Also, the rapid
deppletion of petroleum reserves, severe environmental as a result of the high contamination levels
produced from burning conventional fuels [46]. Thus, liquid biofuels are seen as one promising
alternative to produce transportation fuels such as biodiesel and bioethanol, or chemicals such as
butanol, methanol, and molecular hydrogen (biohydrogen) [47], [48].
The primary raw materials and pathways used to produce different biofuels are shown in Figure
2.1. Between them, oil crops (starches), sugar-containing crops, wet biomasses and residual or
animal fats are used for 1G biofuels production. 2G biofuels and biocrudes are produced from

22
lignocellulosic materials, manure and lipids. Finally, microalgae are the biomass for third-
generation biofuels production and genetically modified microalgae in the case of fourth-
generation biofuels.

Figure 2.1. Feedstocks and pathways of different biofuels production. Adapted from [36].

The time frame 2017- 2040 is contemplated in the policies of the Sustainable Development Goals
(SDGs) of the United Nations (agreed by 193 nations in 2015), and the commitments of the Paris
Agreement. [49], [50], [51]. According to the last BP Energy Outlook, it is foreseeable a
continuous growing demand for energy from renewable feedstocks, including solar, wind
geothermal and biomass [49]. It is foreseeable that the renewables mentioned will penetrate in the
global energy system more quickly than any fuel in history. By 2040, the renewable energies
mentioned, including renewables feedstocks for the production of advanced biofuels are expected
to be the fastest-growing energy sector. Besides, renewable energy sources will enter in the global
23
energy system more quickly than any fuel in history. As a result, the predicted renewable energy
like primary energy wind, solar and wood and second energy sources obtained from the
transformation of renewable feedstocks into biofuels including biomasses for first, second, third
and fourth generation biofuels production will increase be around 30% in 2040 [49]. The
assumptions presented are base on global energy trends.
The transportation sector will increase energy consumption by around 1.1 % per year until 2040.
The good news is that the share of biofuels with the transportation sector is expected to reach up
to 8% in 2040, the current share is 3.5%. As a result, in 2040, a GHG reduction of around 2 Gt of
CO2 is expected to occur [51]. The environmental improvements are predicted to happen not only
by the continuous growing research and development in biofuels production and sharing with
conventional fossil fuels; also, a robust technological innovation is also required [52].

2.3. Biofuel Feedstocks

In this section, the most important feedstocks to produce first, second, third and fourth generation
biofuels are presented. Main feedstocks include lignocellulosic biomasses, sugars, lipids, starches
aquatic biomasses, among others. Biomass produced every year in the world is around 146 billion
metric tons and are considered the fourth most important energy source available, providing 14%
of the total energy [53].

2.3.1. First-generation feedstocks.

First-generation feedstocks are food crops belonging mostly to the human food chain, such as
sucrose-containing, starchy materials, oily feedstocks, aquatic biomasses, industrial and municipal
residues, among others. First-generation feedstocks can also be cultivated not only as a source of
biomass energy but also for the production of fibre, nutritional products, different chemical
derivative products such as pharmaceuticals, polymers, plastics, lubricants among others [54],
[55]. The main first-generation feedstocks to produce bioethanol and biodiesel can be divided into
two main groups tabulated below in Table 1.1.

24
Table 2.1. First-generation feedstocks for bioethanol and biodiesel production.

First Group Second Group (Bioethanol Production)


(Biodiesel Production) Sucrose containing Starches Lignocellulosic
Sugar beet Corn Wood
Oil crops (rapeseed,
Sweet sorghum Wheat Straw
soybean, sunflower,
Sugar cane Potatoe Grasses
palm oil, cottonseed
Palm Juice Barley
etc.)
Molasses (by- Rice
Residual or cooking
product)
oils
Animal fats

The oil containing biomasses from the first group is dedicated mainly for biodiesel production.
After bioethanol, biodiesel is the most important biofuel industrially produced and, Europe is the
main biodiesel producer in the world [56]. Regarding industrial oil crops, McKeon et al. [54]
explored its physical-chemical characteristics and present the most commonly used feedstocks and
the recent progress in the breeding of the main industrial oil crops. The world production of
soybean, rapeseed, cottonseed, peanut and sunflower is presented by Friedt et al. [57]. Biodiesel
is produced through transesterification technology transforming triglycerides from vegetable oils
or fats into a more usable and ecofriendly source of energy for transportation. From the first group
in Table 1, oil crops have very low sulphur content and less nitrogen content than fossil fuels. The
most widely used vegetable oils for biodiesel production are soybean and rapeseed with an oil
yield of 446 and 1146 L/ha, respectively [58]. Some challenges are associated with the growth of
oil crops, the most common problems are related to diseases or pests or even fungus present in the
soil. For example, rapeseed is attacked by diseases such as canker, alternaria, light leaf spot, among
others [59], [60]. Soybean is affected by pests such as bean leaf beetle, aphididae [61], [62]
rhizoctonia solani disease [63], just to mention some. The application of chemical insecticides is
an effective control against most of the pests but, it also can harm other important insects such as
bees or parasitoids; the use of natural biocontrol agents (spiders, ground beetles among others), or
genetically modified seeds are an option to resists diseases or pests.

25
In the United States, soybean is the most important oil crop for biodiesel production and, in
European countries, most of the production of biodiesel is from rapeseed. Furthermore, in tropical
countries, the biodiesel produced is mainly from palm oil and coconut [64]. Europe has been using
biodiesel more extensively than the Unites States and other countries.
Biodiesel is also produced from nonedible oils as well such as castor oil, linseed oil, used
lubricants, paints, tall, jojoba or even pharmaceutical residues [56]. Oilseeds mentioned in Table
2.1, belong to the group of crops or biomasses with economical incentives from the United States
Department of Agriculture (USDA). The incentives are intended to help farmers growing
bioenergy feedstocks; the land must be located within an economically practicable distance from
the biomass conversion facility [65]. The subsidies are $750 US dollars per acre, for underserved
producers and $500 per acre for other producers. Another incentive is for transportation of biomass
residues that are not economically retrievable; this help is 1 dollar per ton transported to an
approved conversion facility.
In the second group of Table 2.1, the feedstocks to produced bioethanol are presented. Capehart et
al. [38], performed a detailed forecast about harvested areas, production, yields and other
important features about corn (mainly for bioethanol production) by the leading producers of this
commodity in the world. As was mentioned in the former, bioethanol is the most important biofuel
produced all over the world, in 2018 the United States with 56% of the global production was the
principal producer. In the second place, Brazil produced around 28% of the global production of
ethanol. Around 90% of the bioethanol produced in the US is from corn [66]. European Union
was the third producer with 5% of the global production in 2018. On the other hand, Brazil is the
second producer of bioethanol, but, is the biggest producer of sugarcane globally [67]. The
feedstock used depends on the climate conditions of the regions; for example, the production of
ethanol in South America [68] is mainly from sugar cane juice whereas molasses, from
autonomous distilleries, is a common fuel source in some European countries [69] and China [9].
The United States is the main corn producer in the globe, and corn is by far the main feedstock for
bioethanol production. From 1 bushel (25.4 kg) of corn 2.87 gallons of bioethanol or 4100 litres
per hectare, 7.21 kg of distillers grains residues for animal feed and other by-products are obtained
[70]. In 2017-2018 around 80 million acres of corn were harvested in the United States with an
average production of 162 bushels/acre [71]. The projected trend is to produce 92.8 million bushels
by 2019-2020. It means 14 bushels more per acre compared to the production in 2017. The corn

26
production is affected by several diseases caused principally by fungi and bacteria. The most
common diseases that are affecting the leaves of the plant are gray leaf spot, eyespot, northern
corn leaf blight southern rust, among others. The stalk is affected by fungus and bacteria as well
but in less proportion than the leaves. Seeds are affected by a disease called Rhizoctonia [72]. The
treatment with fungicides is the most common way to protect the plants against several pests.
Federal incentives such as the biomass crop assistance program are available to help landowners
and operators to produce and deliver eligible biomasses in designated bioethanol production
facilities [65], [73]. The eligible feedstocks produced can retribute up to 50% of the cost of
establishing biomasses feedstock crops. Also, they can receive payments for no more than 15 years
to produce woody feedstocks; these incentives are from $1 us dollar up to $20 for each ton of dry
biomass.
The second main feedstock for bioethanol production is sugarcane, which is mainly harvested in
Brazil. In 2018, 620 million metric tons were harvested in approximately 8.6 million hectares [74].
The yield of TRS or total reducing sugars in that period was 137.52 kg of TRS/metric ton or an
average yield of 6900 litres of bioethanol per hectare [75]. Sugarcane growth in Brazil is affected
by the sugarcane mosaic virus and sugarcane yellow leaf virus. Gonçalves et al. [76] performed a
study about the diseases mentioned above, which are considered endemic in Brazil. Also, the
authors include some controls and elimination treatments.
In Brazil, the government offer incentives and benefits since 1971 for growers and owners of sugar
mills [77]. These benefits intended to relocate and construct mills with better infrastructure in order
to increase their productivity and decreasing transportation costs at the same time.

2.3.2. Second-generation feedstocks.

Second generation feedstocks are non-food biomasses from non-edible crops [13]. Forestry
residues, wood processing by-products, various municipal organic wastes, vegetative grasses,
lignocellulosic crops or by-products from sugar cane, bagasse, cereal straw, among others, are
examples of second-generation biomasses as can be seen in Figure 2.1.
Lignocellulosic materials are considered the most important biomasses for several reasons; first,
they are low cost and the most abundant renewable source with an annual production of around
170 billion tons per year [78], [79] and just 5% of its stored energy s exploited. Additionally,

27
lignocellulosic biomass does not compete with human food crops. The three main sources of
lignocellulosic biomasses are forest, agricultural residues and energy crops [80].
Between them, the most commonly used for biofuel production include pulp and paper industry
residues, bark, wood chips and sawdust, urban and agricultural wastes, crop residues, forest
products such as wood, logging residues, tree branches, from sugar cane by-products (bagasse),
some perennial grasses (Prairie grass, miscanthus, switchgrass), manures, and short-rotation forest
species. Under the appropriate processing conditions, these lignocellulosic materials can be
transformed into valuable compounds like biofuels and chemicals using biochemical or
thermochemical routes [18]. Furthermore, the energy yield of 2G biocrudes produced from
lignocellulosic biomasses is higher in terms of energy produced per hectare of productive land
GJ/ha than the amount of energy obtained from 1G biofuels produced in the same amount of land
with same characteristics [81]. Second generation feedstocks, unlike 1G feedstocks, may also be
grown on agriculturally marginal land. The woody biomasses have been widely used as a second-
generation feedstock [82]. These biomasses are cheap and abundant, making them an attractive
source for the production of value-added products such as biofuels and many chemical products
in a sustainable way [83]. A brief representation of the main chemical components of wood is
presented in Figure 2.2.

Figure 2.2. Chemical wood components adapted from [84].

Regardless of the actual plant source, all lignocellulosic biomass consists of the same basic
building blocks of lignin, hemicellulose and cellulose. The resulting fibrous plant material is

28
composed of 60-80% carbohydrates, and the rest are proteins and lignin, forming a rigid
conglomerated matrix which makes the plant resistant to chemical or biological degradation. The
carbohydrates are composed of polysaccharides, hemicellulose, cellulose inulin and starches [83].

Polyoses (hemicelluloses)
The monomers typically found are C6H10O5 and C5H8O4. Hemicellulose is a polymer composed
of mainly polysaccharides like xylose and arabinose. Its main constituents are five neutral sugars,
the hexoses glucose, mannose and galactose, and the pentoses xylose and arabinose [84]. Some
hemicelluloses may contain uronic acids. Hardwoods contain a higher amount of hemicellulose
than softwoods, and its sugar composition is slightly different.

Cellulose
Its chemical formula is (C6H10O5)n, cellulose is the most abundant biopolymer on earth and is
composed of straight chains of glucose (C6 sugars or hexoses). Almost half of the components of
hard and softwoods are cellulose, which is a linear high molecular weight polymer constituted
exclusively of ß-D-glucose [84], [85]. Cellulose is the main structure of the plant cells as can be
seen in Figure 3.

Lignin
It is a complex phenolic polymer composed of phenylpropanoid units with an amorphous structure.
Lignin’s chemical structure is (C31H34O11)n. However, its structure may vary depending on the
plant species, which is a challenging compound in bioconversion processes [86], lignin gives a
stubborn characteristic and the reason to perform a pre-treatment. Woody biomasses required a
deeper pre-treatment than agricultural biomasses because the first ones have a higher lignin content
making them stronger and denser [87]. Lignin is dissolved under alkaline or acidic conditions
and/or using organic solvents [88]. A representation of the lignocellulose structure found in the
cell wall of plants and its chemical structure and composition is shown in Figure 2.3.

29
Figure 2.3. Structure of lignocellulosic biomass, adapted from [89], [90].

Jensen et al. [21], present a detailed explanation about the composition of woody biomasses and a
list of the most common lignocellulosic biomasses and their cellulose, hemicellulose, lignin and
ash content (wt.%). Alzgameem et al. [79], reported an extensive compendium about different
companies, institutions and pilot plants around the world which utilizes lignocellulosic feedstocks
for the production of hundreds of value-added products such as paper, pulp, lignin-derived
chemicals, lignosulfonates, second-generation biofuels among others. The potential of
lignocellulose biomasses for second-generation biofuels production is a fact because this biomass
is abundant and its production for industrial purposes is continuously growing as is showed
elsewhere (U.S lignin market 2014-2025) [91].

2.3.3. Third-generation feedstocks.

In 1978, the Department of Energy in the United States started a research program called “Aquatic
Species Program” about microalgae biomasses for production of hydrogen initially, and soon
particular attention turns to the biodiesel, bioethanol or methane production from aquatic species
[92].
Microalgae can be found or growth in marine or freshwater, they are unicellular microorganisms
with a fast growth rate and high production capacity of lipids because the aquatic environment
generally has more nutrients and CO2. Microalgae have some advantages compared to terrestrial
plants; for instance, they are more efficient in converting solar energy and can grow in several

30
water sources such as municipal, industrial, and animal wastewaters if the source of carbon,
nitrogen, and phosphorous is enough for them to grow [93]. Algaes are a large group of single or
multicellular plant-like organisms; they can absorb considerable amounts of CO2 and solar energy
to grow via photosynthesis. Algae are divided into two categories algae (unicellular species) and
seaweeds (multicellular species) and some algae species are considered the fastest-growing plants.
Their photosynthetic efficiency is between 3 to 8%, and this is higher than most of the land crops
with around 0.5% efficiency [19],
Microalgae species are used mainly for biodiesel and bioethanol production due to this species
being known to produce large amounts of lipids and fatty acids, which is a potential source for
biodiesel production. According to Alaswad et al. [19] seaweeds or macroalgae, tend to produce
large amounts of sugars and carbohydrates. These sugars are similar to those found in the cellulose
of land plants. The structural features of algae are similar to the land plants structure. However,
macroscopic algae do not require the same strength to resist the gravity force since they are
supported by buoyant forces. In other words, they do not need the same amount of fibrillar material
[94]. Due to its sugar content, this feedstock is also used for bioethanol and biogas production.
Trough biochemical conversion, direct combustion, chemical reaction, biochemical and
thermochemical conversion [20] the biofuels that can be produced are:
• Hydrogen: produced through biomechanical photobiological conversion of lipids.
• Bioethanol: obtained through fermentation processes transforming the carbohydrates with
bacterias, yeast or fungi.
• Biomethane: produced through anaerobic digestion of aquatic biomass.
• Syngas: is obtained through gasification or pyrolysis of microalgae biomasses.
• Biocrude: obtained through liquefaction processes.
• Biodiesel: is produced using transesterification after the extraction of the high contents of
lipids contained in microalgae feedstocks.

Third generation biofuels are a promising technology which is in continuous research and
development. However, some drawbacks associated should be mentioned; for instance, the
production of microalgal biomass is still technologically challenging because the production of
microalgae feedstocks at large-scale production is still not feasible. For example, in the United
States, China, New Zealand, and other developed countries, the production of algal biodiesel is at
a research-level. Even though it is commercially attractive, the technology is still emergent [95].
31
The process of producing microalgae is generally more expensive than growing land plants
because they require appropriate growing protocols such as adequate kind and amounts of nutrients
and rigorous temperature controls [96]. Open-air ponds are an old and extensively studied
technology; lakes and natural ponds have been used to study the production of microalgae intended
to reach a commercial scale. Although, it is the cheapest technology available, this system is
technically challenging because the temperature can not be controlled and the exposure to
changing weather conditions, natural predators, and contamination results in poor biomass
productivity [97].
On the other hand, tubular, flat plates, and column photobioreactors are more efficient. Closed
reactors have the advantage of better control conditions, maintenance, mixing and higher amounts
of CO2 because a small open area releases less CO2 to the environment. Another drawback about
the production of aquatic biomasses is the fact that they can grow in an uncontrolled way if
eutrophication occurs; this is a result of minerals and nutrients excess. As a result, very harmful
consequences to other aquatic species can occur due to the excessive consumption of oxygen
(biomass decomposition) [98], [20].
One of the most important studies about the production of transportation fuels from aquatic
biomasses was performed by the National Renewable Energy Department of the United States
between 1978 to 1996 [92]. After the study, several improvements in lipid productivities were
achieved by performing experiments in ponds with mixing devices and growing the best
microalgae cultures. Moreover, it was found that microalgae produce higher lipid yields than those
produce by macroalgae. Finally, they suggest that more efforts in R&D, including the option of
genetic modifications of microalgae species, are needed to reach an industrial and sustainable
production of biodiesel from microalgae in the future.

2.3.4. Fourth-generation feedstocks

As a result of the extensive research and developments, several findings have been made about
technologies on aquatic biomasses for biofuels production. Microalgae are known by its high lipid
content between 20-70% in dry weight, option desired for biodiesel production [99], [97]. Even
though some species have been highlighted by their high lipid productivity such as green algae
and eustigmatophytes, there is even more room for improvement, in terms of lipid content and

32
lipid productivity [100]. In this regard, fourth-generation feedstocks are microalgae species
subjected to genetic and metabolic modifications in order to improve its lipid content and biomass
yield. The goal is to create microorganisms with high growth rates and higher lipid storage capacity
[95] or in the best-case scenario, develop species able to produce fuels directly from CO 2 [100].
This can be achieved by understanding and modifying the lipid metabolism of the species as well
as its enzymatic and regulatory functions in the photosynthetic cells [101]. In addition, the genetic
modification of microalgae has been focussed in the modification of the CO 2 absorption capacity.

2.4. Production Methods for Biofuels

There are several technologies developed for the production of biofuels; the methodologies can be
grouped according to the processes and technologies employed. The production of biofuels from
renewable sources include chemical, biological and thermochemical routes; regardless of the
pathway chosen, the overall production methodologies involve pre-treatment, reaction and
upgrading.

2.4.1. Pretreatment

The processing of lignocellulosic materials that contain sugars as intermediates require the
modification of their physical and chemical structure. In the past few years, some techniques have
been developed to pretreat the feedstock and make the sugars more accessible to hydrolysis for
conversion to biofuels [102],[103]. The pretreatment is a critical step required to partially separate
the major polymer components (cellulose, lignin, and hemicellulose). It reduces the crystallinity
of the cellulose, increases the access of enzymes to the substrate, enhances the surface area and
the porosity resulting in increased hydrolysis rates [104]. The methods can be classified into four
categories; physical (mechanical), chemical, physicochemical, and biological, which are briefly
summarized below.

Physical Treatment:
The main objective of physical or mechanical treatment is to reduce the cellulose crystallinity
through a combination of grinding, chipping, and/or milling. Besides, it helps to improve the
digestibility and the conversion of saccharides [105]. It has the advantage of not producing

33
inhibitors due to the no addition of chemicals. However, it requires high energy consumption and
long operation times.

Physicochemical Treatment
▪ Steam Explosion:
The biomass is exposed to hot steam (160-290 oC) under pressures between 2000 and 5000 kPa.
Later, the reaction is stopped by slowly depressurizing the system to atmospheric pressure [106].
The expansion of the steam within the lignocellulosic matrix allows separating the individual
fibres. This process is economically attractive due to its a high recovery of xylose (45-65%) [107].
▪ Liquid Hot Water:
The biomass is put in contact with hot liquid water (200-230 °C) at pressures above 5MPa for
about 20 minutes in order to hydrolyze the lignocelluloses and remove lignin [67].
▪ Ammonia fibre explosion:
It involves the use of liquid ammonia and steam explosion where the biomass is treated at high
temperature and pressure followed by a fast depressurization. It has the following advantages: low
generation of inhibitors, efficient removal of lignin and significant retention of carbohydrates in
the substrate [67].

Chemical Treatment:
It involves the use of acids (oxidants or alkalis), ammonia, organic solvent, CO2, or SO2 to break
down the organic compounds that are present in the biomass [108].
▪ The acid pretreatment has been considered one of the most important techniques due to the
production of high yields of sugars. Usually, the sulfuric acid is used diluted or concentrated
(0.2-2.5% w/w) at temperatures between 130 and 210 oC to break the rigid structure of the
biomass [109].
▪ The alkaline pretreatment uses sodium, calcium, potassium, and ammonium hydroxide to
digest the lignin matrix to make the hemicellulose and cellulose available for further
enzymatic degradation [109].

34
Biological Pretreatment
It uses microorganisms such as white rot, brown rot and soft rot fungi to degrade the lignocellulosic
matrix to release the cellulose. This method is environment-friendly, requires low consumption
of energy, and does not need the use of complex procedures and equipment. However, the
hydrolysis rates and yields are low, which impede its implementation [108].
Generally, the biorefinery industry applies a combination of methods to pretreat the biomass for
further conversion to biofuels, and some examples are discussed below.
Zhu and Pan 2010 [110] showed that the pretreatment of woody biomass for biofuel production
requires the application of both physical and thermochemical processes. First, the wood size is
reduced to the level of fibres from chops and then is pretreated at 190 ºC with sulfuric acid (H 2SO4).
However, Zhu et al. [111] proposed a different approach to pretreat the same type of biomass; the
material was first pretreated with H2SO4 followed by disk-milling to generate a solid substrate.
This pretreatment process achieved a high substrate enzymatic digestibility and had less
consumption of energy. Also, Zhang R. and Zhang Z [112] studied the production of biogas from
rice straw where it was found that a combination of grinding, heating, and ammonia treatment gave
a high yield of biogas. For this specific biofuel, the particle size of the substrate plays an important
role in the performance of the anaerobic digestion of municipal solid waste. Regarding the
bioethanol production, the use of agriculture wastes has been extensively studied in the past two
decades [108], [113].
In addition to the above, Chen et al. [114] pretreated the corn straw by applying a mechanical and
a physicochemical method. The biomass was first chopped and then heated at 275 ºC with high-
pressure steam (3.5 MPa). The yield of sugars such as xylose, glucose, and arabinose were reported
to be 23.6, 56.7, and 5.7 g/L, respectively. Finally, Wan and Li [115] treated corn stover using a
microorganism called “Ceriporiopsis subvermispora.” It was shown that this microbial organism
was able to selectively degrade the lignin up to 32 % where the loss of cellulose was less than 6%.

2.4.2. Reaction

The most representative feedstocks for the production of bioethanol, biodiesel and biocrudes were
reviewed in the previous section. Now, the process of transforming those biomasses into value-
added products with higher heating values and/or improved octane number in the case of biofuels

35
requires the application of different reactions depending on the chemical composition of the
feedstocks [58], [116]. Several reaction pathways have been developed to produce compounds as
close as possible to gasoline and diesel fuels or drop-in biocrudes which are chemically
indistinguishable to fossil fuels. When this upgrading is achieved, biofuels can be blended in high
percentages with conventional fossil fuels [117]. The main biomass conversion pathways for
biofuel production include biochemical and thermochemical conversion. Fermentation and
transesterification are biochemical processes vastly used for the production of bioethanol and
biodiesel [118]. On the other hand, pyrolysis, gasification and hydrothermal liquefaction are
thermochemical routes [64]; these reactions pathways are briefly explained in this section.

Fermentation

In order to transform a crop such as sugar cane or corn into ethanol, a fermentation process is used.
Depending on the source of the biomass, the fermentation may also be preceded by a pre-treatment
step. Fermentation is a widely used biochemical process that produces chemical changes in the
organic matter through an enzymatic process. The fermentation is one between several processes
that extract energy from carbohydrates or sugars such as glucose, lactose, starches, among others.
Other technologies for energy production from biomasses are gasification, pyrolysis
(thermochemical routes) [119].
In Figure 2.4, a brief description of the process, including the main steps and reactions of the
distillation products is shown [67]. First, a pre-treatment is performed to release cellulose, starch
or sucrose contained in the biological parts of the raw materials. The next step is the hydrolysis or
saccharification (acidic or enzymatic) of pre-treated raw materials. This step can be done at low
temperatures and high acid concentration or at higher temperatures and low acid concentrations
[120]. The goal of hydrolysis is to increase the production of sugars that will be transformed into
bioethanol through the fermentation process.
After hydrolysis, the fermentation process of pentoses or hexoses is performed by unicellular
microorganisms that process or dissociate the sugars and produce ethanol, CO2, and adenosine
triphosphate. The preferred microorganism for bioethanol production is the yeast Saccharomyces
cerevisiae. This yeast has high ethanol and other inhibitors resistance; it also consumes high
amounts of substrate even in harsh conditions. Furthermore, it is one of the most widely used

36
microorganisms at the industrial level and can resist pH variations [121]. However, this
microorganism can process just hexoses [67].
After fermentation, a distillation step of the fermented broth is carried out, and finally, the
bioethanol is obtained. Optimum fermentation temperatures are 20 to 35 °C; and optimum pH
range of 2.75 to 4.25 for the yeast Saccharomyces cerevisiae. When other microorganisms are
used, those levels can be slightly higher up to 42 °C and a pH of 5.5. Fermentation for bioethanol
production can be performed in batch, semi-batch or continuous operation process reactors.
Bioethanol is produced after a process of 24 to 72 hours in continuous agitation mode [121].

Figure 2.4. Pathway to obtain biofuels through fermentation.

Transesterification

Biodiesel (methyl esters) is produced through a very well-known chemical process called
transesterification, carried out to transform triglycerides from edible and non-edibles oils into fatty
acid esters decreasing the viscosity of triglycerides [122], producing a suitable biofuel for use in
conventional compress diesel engines. The preferred feedstocks for biodiesel production are those
with a high triglyceride content (higher than 90%) and with a low free fatty acid content, which is
the case with soybean, rapeseed, sunflower and canola crops [17]; depending on the nature of the
feedstock oil, a pretreatment step may be required. The transesterification reaction involves the

37
reaction of the feedstock oil with methanol, in the presence of a homogeneous catalyst such as
alkali-based NaOH or KOH. On the other hand, for esterification, the catalyst used is a strong acid,
for example, H2SO4. In industrial production of biodiesel through transesterification, the alcohol
is used in excess to shift the transesterification reaction towards the formation of alkyl esters [123].
During the process, considerable amounts of water are produced. Then, the homogeneous catalyst
tends to decrease its concentration when it is carried by the stream in the subsequent washing steps
[116]. Therefore, high amounts of homogeneous catalysts must be used to achieve high yields of
the ester and to meet glycerine standards (commercial purposes).
When the triglycerides react with the alcohol, the fatty acid chains are separated from the glycerol
frame and combined with the alcohol to yield fatty acid methyl esters (FAME). This reaction is
represented by equation 2.1 adapted from [123].

Equation 2.1.

The process consists of an intense mixing of the homogeneous catalyst with methanol in a closed
reactor. The agitation process takes around 2 hours at 70 °C and ambient pressure. The reaction
products are two liquid phases composed by glycerol with some catalyst and methyl esters. As a
result of the intense mixing, the methyl esters produced from an emulsion with glycerol, as a result
of the mutual solubility of biodiesel and glycerol [124]. After two hours of settling, most of the
products are separated biofuel produced is separated, but not all the glycerol is still recovered. The
biodiesel produced has a high similar chemical structure with the diesel obtained from fossil fuels
[125], [126]. The production of biodiesel has some technical difficulties; the most challenging is
the separation process of final products. When the mixing is poor, or the amount of catalyst is low
undesirables products such as soap are produced. Soaps increase the viscosity, which triggers gum
formations and increases production costs. Some techniques for biodiesel production, such as
supercritical methanol process, can improve the feasibility of the process [127]. This option does
not require a catalyst, the reaction times are lower and consequently, a simpler separation step.

38
Pyrolysis
Is a form of thermochemical decomposition of biomass that breaks down large molecules (like
cellulose) into smaller and more valuable compounds. The pyrolysis process is carried out in an
environment where the oxygen present is near to zero [128]; since no oxygen is present, the organic
material does not combust. Instead, the chemical compounds decompose into three forms: a liquid
bio-oil, solid biochar and/or syngas. The pyrolysis process can be represented with the generic
reaction adapted from [129]:

Equation 2.2

In Figure 5, a typical pyrolysis process is represented. First, a primary decomposition step of the
biomass is performed in a fluidized bed where the solids are heated at the pyrolysis temperature
and finally burned with gas. This step is carried out at atmospheric pressure and limited oxygen
environment, converting lignocellulosic feedstocks into biofuel by heating them rapidly. Next,
char is collected, and the condensable gases produced break down into non-condensable gases
(CO, CO2, H2O and CH4). Then, final products are solid char, liquid products (tar, heavy
hydrocarbons and water) and gases such as CO2, H2O, CH4, C2H6, among others [130].
Heating rate and time are essential factors in pyrolysis technologies. Based on these factors,
pyrolysis is classified into fast and slow pyrolysis. If the process is carried out at a high temperature
and with longer residence times, the process will favour the production of gas products. In contrast,
with short residence times (just a few seconds) can lead to a high yield of liquids [131]. Pyrolysis
is a pre-step in gasification of biomass.

39
Figure 2.5. General scheme of the pyrolysis process adapted from [132].
Some of the critical factors in the process are pressure, temperature and heating rates; the
temperature range of the process is between 300 to 650 °C. The pyrolytic reactor is the most
important part of pyrolysis; they can be either, batch or continuous reactors. For example, a fixed
or moving bed, bubbling fluidized, ultra-rapid reactor, rotating cone, ablative reactor and vacuum
reactor bed. Between them, the fluidized bed bubbling or circulating reactor is the most widely
used [133].
Among all the pyrolysis technologies, fast pyrolysis was developed to obtain higher amounts of
liquid bio-oils. For that reason, it is one of the most studied. An extensive review about the most
relevant features, products, major reaction systems and processes on fast pyrolysis in the last three
decades was performed by Bridgwater and Peacocke [134]. Furthermore, Venderbosch and Prins
revised the current technologies of fast pyrolysis and stablished some challenges that pyrolysis
technology will face in the upcoming years. Another study about the production of biofuels from
lignocellulosic materials through fast pyrolysis was done by Nowakowski et al. [135]. The project
was carried out in collaboration of fourteen laboratories located in eight different countries. Two
purified lignin samples were appropriately characterized and treated in bench-scale bubbling
fluidized bed reactors. Several results were reported, including thermogravimetric analysis. The
best results were bio-oil yield of 49.7 %, 42.3% of char yield and 8% of gas.

40
Gasification

Gasification is another thermochemical pathway for conversion of biomasses into synthesis gases
(CO + H2) using different thermochemical routes. Gasification, instead of breaking bonds as occur
in combustion, the energy is packed into the chemical bonds of the product gas. Moreover,
gasification adds hydrogen and removes carbon from the feedstock. Hence, the gas products have
higher hydrogen to carbon ratio [136].
In order to transform biomasses into gas products, the primary route is always partial oxidation,
pyrolysis or a combination of both [137]. During homogeneous gasification reactions in the gas
phase and heterogeneous reactions between the char (solid) and the gas phase occur.
The process for synthesis gas production from lignocellulosic feedstocks or coal is shown in Figure
2.6. First, the lignocellulosic biomass is pretreated. Next, the biomass with small particle size is
subjected to pyrolysis at relatively low temperature. Next, partial oxidation and hydrogenation
occur at higher temperatures. Finally, value-added products are obtained, the tars are removed, and
after filtration, the produced gas is ready to be used (synthetic diesel, aviation fuels, among
others.). Some gases produced can also be used for generation of electricity or heat [18].

Figure 2.6. Main steps in gasification adapted from [138]

One of the most important gasification technologies is catalytic Fisher-Tropsh synthesis. This
technology consists of reforming of cleaned gases product of gasification (syngas); the gases are
combined with oxygen over nickel catalysts to form a waxy syncrude which is cracked and refined
to produce synthetic or long carbon chains biofuels such as marine or aviation biofuels, biodiesel
as well as chemicals. Fisher-Tropsch is a commercially available technology used by SASOL
(South Africa) for producing diesel fuel [137].

41
Hydrothermal Liquefaction Process (HTL)
HTL is another thermochemical process like pyrolysis, where physical and chemical
transformations of the feedstocks occur. The HTL process requires the presence of solvents
aqueous/organic solutions such as hydroxyl compounds, ethylene glycol, glycerol, low-carbon
alcohols or water. Among them, water is the most common and cheapest solvent available.
Hydrothermal refers principally to water at high temperatures and pressures [139]. Hydrothermal
liquefaction technology is superior to pyrolysis, gasification or combustion for several reasons.
First, a special pre-treatment of the feedstocks in HTL is not needed; HTL is able to transform a
wide range of wet biomasses into high-quality biocrude with improved heating value and lower
oxygen content compared to bio-oils produced from pyrolysis. Also, through pyrolysis and
combustion normally is possible to treat just dry biomasses. For HTL, feedstocks with high
moisture content are instead an advantage because the water in the process is a conversion medium
[140].
HTL mimics the natural process of fossil fuels production through geochemical processes (high
temperatures and pressures) from ancient organic materials. However, HTL has an advantage,
which is the short process times from seconds up to 10 minutes depending on the feedstocks and
conditions [38]. Typical conditions for HTL are temperatures near or below the critical point of
water from 280 to 374 °C and 450-600 °C for supercritical conditions; the pressure range is
between 10 to 35 MPa [21], [141]. Water at high temperatures and pressures provides unique
properties to the HTL process, for example, a high ion product and a low dielectric constant. These
concepts will be explained with more detail in the section about Steeper’s energy HTL
HydrofactionTM technology. Another advantage of water at those conditions is that low amounts
of acid or base catalysts are required for the process.
The reaction mechanism of HTL has been studied by Wang [142] and Gai et al. [143], among
others. The exact pathways are not very clear due to the biochemical complexity of both the
feedstocks and the products, and due to the huge number of intermediate chemical reactions that
take place. Nevertheless, a good approximation is given by Zhang and Chen [144] and Sing et al.
[145]. They concluded that the process consists of three basic HTL reaction pathways:

42
• Depolymerization of lignin, lipids, and proteins
In this stage, several reactions occur, for example hydrolysis and cleavage of ether and C-C bonds,
as well alkylation, and condensation reactions. For instance, carbohydrates undergo rapid
hydrolysis forming saccharides, which are further transformed into oxygenated compounds
(formic acid, lactic acid, levullinic acid).
• Decomposition of biomass monomers like fatty acids, amino acids, and glucose
These compounds are the result of decarboxylation and deamination reactions.
• Recombination and depolymerization of reactive compounds, for instance, radicals such as
hydroxyl groups (OH).
HTL can be performed in batch, semi-batch and continuous reactors. Batch reactors are easy to
handle and operate. Batch reactor systems are appropriate for laboratories and pilot plant scale
studies because the production of biocrude is low to be used in commercial scale. The most
common is the steel batch autoclave batch reactors [146]. The semi-batch reactors have some
advantages over batch reactors, for example, the temperature can be controlled easily and products,
sampling is easy and can be done at any time of the process. This option is for low amounts of
feedstock as well, and the productivity is still for industrial purposes. Finally, continuous mode
operation reactors are tubular steel heated reactors. For continuous operation, a high-pressure
slurry pump is required. This option allows the feeding of high-concentrated slurries and a more
significant scale production [147].
There are several publications available about Hydrothermal Liquefaction process (HTL), some of
the first research publications about HTL were some patents by Doctor Carl Krauch in the 1920s
about the liquefaction of carbonaceous materials [22]. Another research work about the conversion
of wet feedstocks through the HTL process was released back in the 1930s where one large-
commercial-scale process on HTL was leading at Energy Research Centre in Pittsburg 1970 [148].
Since then, this process has been continuously improved by several research institutes, private and
governmental entities around the world. Nowadays, this technology is considered for some authors
as one of the most promising pathways for 2G biofuels production [149].

43
Hydrothermal Liquefaction Process HydrofactionTM (Steeper’s Energy Technology)
In this thesis, the experiments will be performed with a biocrude that has been produced using
Steeper Energy’s patented process known as HydrofactionTM. HydrofactionTM is a thermochemical
technology that is able to transform several lignocellulosic feedstocks into biocrudes. However,
its primary application has been on forestry residues [22], [21].
The basis of HydrofactionTM lays on hydrothermal liquefaction technology, and the most important
features of this technology are explained herein. The goal of Hydrofaction TM is the production of
low oxygen content and high energy-density renewable biocrude, which after an upgrading step
can be compatible with fossil fuels for transportation. To do so, Steeper energy created a novel
process which consists of a combination of water at supercritical conditions and homogeneous
alkaline catalysts.

Figure 2.7. Simplified process diagram about the production of HydrofactionTM Biocrude.
Adapted from [21].
In figure 2.7, the steps of HydrofactionTM process are shown. First, lignocellulosic biomass is
subjected to a size reduction in a crusher and mixed with aqueous products, recycled oil from the

44
process and homogeneous catalysts (potassium carbonate and sodium hydroxide). Sodium
hydroxide is used to maintain the pH of the process at alkaline conditions. Next, the components
are well mixed to form a slurry. The slurry is pressurized (300-350 bar) using a high-pressure
pump; the mixture is pumped to the heaters where is rapidly heated to 390-420 °C. Later, the
preheated slurry enters the reaction zone. In this section, the feed is kept between 10 to 15 min at
the reaction conditions (450 °C and 350 bar). In this stage, the main chemical reactions occurred.
After the reaction stage, the product mixture is quenched to about 150 °C before passing through
a filter. The pressure of filtrated products is performed in a capillary system. Gas products are
collected, measured and analyzed in a gas chromatograph. Finally, the liquid products biocrude
and water are separated by gravity.

Effects of water above supercritical conditions


Water is an excellent solvent, but, above its critical point (374 °C and 220.8 bars), the alkaline
water used in the HydrofactionTM process acquires interesting characteristics. For example, at
those conditions, it acts as a reactant, as a reaction medium, and as a catalyst or even as a catalyst
precursor [140]. Water is a polar molecule by nature; the hydrogen atom is bonded with one
oxygen atom of the molecule through covalent bonds. Also, the same hydrogen molecule is
attracted (hydrogen bonds) at the same time towards a nearby oxygen atom of another water
molecule. These hydrogen bonds attractions are responsible for the actual shape or state of water.
However, at conditions above the supercritical water point, some particular changes occur in its
properties, such as a density and constant dielectric reduction [38]. The dielectric constant of water
plays an essential role in the chemistry of the process; when it decreases, can improve the solubility
of the organic compounds treated, promoting the depolymerization of macromolecules using
hydrolysis and solvolysis [21]. In other words, the water turns into a stronger solvent allowing the
hard compounds of lignocellulose to be miscible and breaking them down easily. Besides, the
reactions that generally happen in a multiphase system will occur in a single one removing the
mass transfer effects that turn the reaction rates of the process difficult [151].

Effect of alkali metal catalysts addition


In conjunction with water at supercritical conditions, potassium carbonate at concentrations around
15 g/l, ammonia hydroxide at1 wt% and sodium hydroxide at 2.5 wt% are used in Hydrofaction TM.

45
At alkaline conditions, water gas shift reactions are promoted. Also, the degradation of
macromolecules is improved as well as decarboxylation and depolymerization reactions. Also,
undesired production of coke, tar and char are inhibited. The potassium carbonate and the water
generate an alkaline environment promoting even more desired reactions such as gasification,
water gas shift (WGS) and steam reforming [140]. Alkaline conditions during the process are
preferred for some reasons; first, the production of hydrogen via WGS reactions is favoured;
second, char is produced just in acid environments, and lastly, the H 2/CO ratio is high at alkaline
conditions, meaning that CO is being consumed and hydrogen is produced [21].

Effects of recirculation of water and biocrude produced


In HydrofactionTM recirculation of water separated by gravity from the biocrude is an important
step in the recycling. The liquid that is rich in water-soluble organics, alcohols, phenols and other
components is reintroduced to the process, concentrating the aqueous products which reduce the
viscosity of the lignocellulosic feedstocks that are pumped in the process. By doing this, the
operating costs decrease by reducing the amount of alkali metal catalysts used [21]. Some
components of the effluent such as phenol are known to foment the depolymerization of lignin
through hydrolysis; phenols can also reduce or inhibit the char and coke formation. On the other
hand, the recirculation of biocrude remarkably improves the Hydrofaction process improving the
rheology of the feedstock, reducing the initial energy required to preheat the feedstock.

Chemical Reactions Involved in HydrofactionTM process

Several chemical reactions occur during the transformation of biomass through HydrofactionTM
process. The chemical reactions occur in three main stages of the process, and the reactions that
take place are tabulated below in Figure 2.8. The pre-treatment performed with sodium hydroxide
and organic solvents, make possible the separation of the main compounds of lignocellulosic
feedstocks [22]. These compounds are hemicellulose, cellulose and lignin. Later, the pre-treated
feedstock reaches high temperatures that depolymerize macromolecules into oligomers and
monomers by the action of hydrolysis and solvolysis reactions.

46
Figure 2.8. Main reactions of HydrofactionTM process, adapted from [21]

The depolymerization occurs first in hemicellulose compounds. Consequently, at higher


temperatures depolymerization of cellulose occurs; also, dehydration of formed monomers and
oligomers take place to form enols, carboxylic acid and aldehydes [21]. Finally, during quenching,
undesired aldol and dehydration reaction occurs.

HydrofactionTM Biocrude
The produced biocrude is a viscous liquid fuel very similar to conventional heavy oils.
HydrofactionTM biocrude has a high energy density or a high amount of energy storage per volume.
It also has low nitrogen and negligible sulphur content and very close chemical characteristics with
conventional fossil fuels. The heavy biocrude obtained requires further upgrading due to its high
viscosity, high aromatic and oxygen content, which is around 10% or more depending on the
characteristics of the feedstock [152]. In Table 2.2, a comparison of properties of some typical
lignocellulosic biocrudes and petroleum crude oil are tabulated. It is noticeable that biocrudes, in
general, require further upgrading before they can be used as transportation fuels either as drop-in
or neat form. The main undesirable properties are a high oxygen content, high Total Acid Number

47
(around 50 mg KOH/g or higher), high viscosity as well as the presence of some thermally unstable
components that generally lead to gum formations [21].

Table 2.2. Properties of typical biocrudes and a Canadian bitumen

Analysis BiocrudeA Alberta BitumenB

Viscosity (cP) 1100-30000 12000-14310


HHV (MJ/kg) 20-38 40-44
Moisture content (wt.%) 1 – 30 0.1-1
Elemental Analysis (wt. %)
Carbon 42.6 - 80.3 82 - 83.9
Hydrogen 5 - 9.4 10.5 - 10.6
Oxygen 9.5 - 40 <1
Nitrogen < 0.2 <1
Sulfur < 0.05 4.2 – 6.0
H/C 1.3 – 1.4 1.5-1.76
Adapted from A: Haghighat et al. [153],[154], Trujillo [155], and Mostafazadeh et al. [156]. B: Gray
[157].

Some recent studies were performed by Albrecht et al. [158]; they investigated the production of
biocrude from aquatic biomasses at 350° C and 20.7 MPa. Another case of study of HTL biocrude
production from microalgae was carried out by Gai et al. [143], they obtained high liquid biocrude
yields (around 50%) when the process was performed at temperatures from 280 to 300° C. These
results agreed with the findings presented by Jena and Kastner [159], whose results were obtained
at similar conditions and same algae feedstock.
Elliot et al. [160] carried out studies about the liquefaction of lignocellulosic biomasses (corn
stover and pine wood) in continuous flow processing systems. The research was carried out at
different temperatures, pressures, as well as the effect of recycling aqueous products.

48
2.4.4. Upgrading

Bio-oils are complex mixtures of water and organic compounds such as alcohols, acids, ketones,
phenols, aldehydes, esters among others. The acids, aldehydes, water and large molecular
oligomers have been identified for being the main compounds responsible for the poor fuel
properties of the bio-oils and biocrudes [161]. Additionally, it has undesired properties for fuel
application such as; high water content, high ash and oxygen content, high corrosiveness and high
viscosity [162].
Therefore, the bio-oil and biocrudes must be adequately upgraded to produce a biofuel of high
quality that can be used and commercially acceptable as a transportation fuel. Several technologies
have been developed to upgrade their physio-chemical properties; those technologies can be
categorized into two main groups: physical and chemical upgrading methods.

Physical Methods

▪ Solvent addition

Organic solvents such as methanol, ethanol and furfural have been used to reduce the viscosity
and increase the heating value of the bio-oil. When the alcohol is added to the bio-oil, reactive
molecules such as organic acids and aldehydes are transformed to esters, and acetals [163]. It
represents a practical approach for bio-oil quality upgrading due to its simplicity and economical
feasible.

▪ Filtration

Bio-oils contains solid particles, alkali metals and biochar that can be reduced through hot-vapour
filtration using granular filters; these physical methods reduced the amount or content of the
undesired solid particles mentioned as no chemical reactions occur. Normally, the ash and alkali
content can be decreased to less than 0.01 % and 10 ppm respectively. Usually, glass wool and
granular filters are employed to upgrade the quality of the bio-oil, including solids content,
viscosity, acidity, and ash content reduction [161].

49
▪ Emulsion

Bio-oils are not miscible with hydrocarbon fuels, but with the help of proper surfactants, they can
be blended. Extensive research studies of bio-oil upgrading through emulsification with diesel has
been done (15). Canmet Energy Technology Centre developed a process to get stable
microemulsions with 5-30% of bio-oil in diesel. They proved that these emulsions were less
corrosive and had promising ignition characteristics [164].

Chemical Methods

• Hydrocracking

This process combines hydrogenation and cracking, and it is carried out at temperatures above 350
ºC and high pressures over bifunctional catalysts (metal/zeolite). The zeolite provides the cracking
function, whereas the metal oxides such as nickel or molybdenum promote the catalytic reactions
towards light hydrocarbons [45]. From a molecular structure point of view, hydrocracking allows
the rearrangement of the carbon chains, which directly affects the octane number, distillation
range, viscosity, freezing point, and cloud point [165].
Catalytic cracking is a mature upgrading technology that is performed to lower the consumption
of energy. Catalytic upgrading of bio-oil includes cracking, decarboxylation, decarbonylation,
hydrocracking, hydro-deoxygenation and hydrogenation reactions [166]. Cracking reactions have
been reported to occur inside the zeolitic catalyst mainly. Twaiq et al. [167] reported the cracking
of palm oil using three types of zeolites named as HZSM-5, HBES and USY. The feed was placed
in a fixed bed reactor in the temperature range of 350-450 ºC. The conversions were reported to
be 99%, 82% and 53% with a gasoline selectivity of 28%, 22% and 7% respectively.
Zhang et al. [168] upgraded bio-oil using Ni/HZSM-5 at a ratio of Si/Al equal to 38. The results
indicated that Ni/HZSM-5 was able to alter the properties of the bio-oil efficiently. The hydrogen
content was increased from 6.2 to 7%, the pH increased to 4.1%, and the heating value went up
from 13.8 to 14.3 MJ/Kg.

50
▪ Catalytic esterification

Bio-oils have a variety of organic acids which are responsible for causing acidity and high
corrosion. Those undesirable organic acids can be converted into esters by catalytic esterification.
This process is carried out in the presence of a catalyst with alcohol [169].
For bio-oils with a significant number of aldehydes, a technology called ozone oxidation can be
applied to convert the aldehydes into acids as discussed by Xu et al. [170]. After the pretreatment,
the acid value increased from 45.4 to 118.4 mg KOH/g. Then, the preprocessed bio-oil was further
reacted with butanol over the NaHSO4 catalyst resulting in better thermal stability and less degree
of the oligomerization.

▪ Hydrotreating

Catalytic hydrogen treating or hydrotreating is a process that removes undesirable materials from
petroleum or biocrudes compounds. The feedstock to be upgraded is selectively reacted with
hydrogen in a reactor at high temperatures and moderate pressures to remove heteroatoms (oxygen,
nitrogen and sulphur, among others) [171]. It is a nondestructive process able to improve the
product quality (higher hydrogen content) of the feedstock without significantly altering its boiling
range.
Sabatier and Senderens back in 1897 [172] published their discovery about how unsaturated
hydrocarbons in the presence of nickel catalysts could be hydrogenated in the vapour phase. The
findings of Sabatier were limited to a few compounds and were performed at low temperatures
and ambient pressure. Later, several experimental contributions carried out at higher hydrogen
pressures, extended the range of feasible hydrogenation reactions. For example, Bergius (1910)
invented a process to produced gasoline by cracking coal and tars and heavy residues in the
presence of catalysts [173]. The production of gasoline through hydrotreating during the second
world war, the beginning of catalytic reforming with the commercial production of hydrogen
(1950), and the increasing interest of the petroleum companies in hydrogen treating were factors
that propel this technology. As a result, in 2001, there was an important installed production
capacity of 39 million barrels/day in 1.600 hydrotreaters constructed all around the world [171].
In the oil-refining industry, hydrotreating is a key process carried out in trickle bed reactors. The
typical catalysts used in hydrotreating are sulphide forms of NiMo/Al2O3 and CoMo/Al2O3, where

51
Al2O3 is impregnated with the promoted systems NiMo and CoMo. The Al2O3 support is known
for its high specific surface area, and good interactions with the promoted systems mentioned
[174]. Fossil fuels contain relatively high sulphur contents, as is shown in Table 2. For that reason,
there is no need for sulphur addition in the feedstock.
Contrary to fossil fuels, biocrudes contained negligible amounts of sulphur (Table 2.2). Then, the
addition of DiMethylDiSulfide or H2S is required to avoid the early deactivation during
hydrotreating of the metallic catalysts mentioned. Low contents of sulphur in the feedstock
promote changes in the structure of the active sites of the catalyst promoting “leaching of sulphur”
[175]. Furthermore, water may also lead to catalyst deactivation [176] then, leaching and water
promote coke formation and subsequent deposit over the surface of the catalyst. Popov et al. [177],
Badawi et al. [178] and Raybaud [174] performed studies and presented a reasonable explanation
about the deactivation effects due to the of lack of sulphur in biocrudes.
Hydrotreating is performed in downflow trickle bed reactors packed with NiMo and CoMo
catalysts allows a three-phase gas-liquid-solid reaction system. In a trickle bed reactors, is possible
to load considerable amounts of catalysts, perform isothermal reactions and manage high hydrogen
pressures [179], [24]. Nonetheless, trickle bed reactors are not suitable for reactions with rapid
deactivation catalysts because it is not very practical. Another drawback is the possibility of
formation of preferred paths of liquid feedstocks inside the catalytic bed [180], this can lead to
excessively heat generations in certain areas and finally, a reactor runaway. A runway occurs
when excessive heat generated in packed bed reactors is not controlled and may lead to explosions.

Hydrotreating Reactions
During hydrotreating, several reactions take place simultaneously. However, three main reactions
that occur during the upgrading off fossil fuels are hydro-desulfurization (HDS), hydro-
dearomatization (HAD) and hydro-denitrogenation (HDN) [181]. Moreover, reactions such as
hydrocracking (HYC), and hydro-demetallization (HDM) take place when the feedstocks are
heavy oils.
Hydro-desulfurization transform sulphur compounds present in the feedstock into hydrogen
sulphide; in hydro-denitrogenation, the nitrogen compounds are converted to ammonia. Hydro-
demetallization converts organo-metallic compounds to metal sulphides.

52
In the upgrading of biocrudes, the oxygen removed could come from different oxygenated
chemical groups. These groups are aldehydes, acids esters, ketones and phenols, among others
[175]. Oxygen excess in biofuels causes instability because they tent to polymerize; it leads to poor
performance when they are used for combustion [4]. Instability is produced from reactions
between aldehydes and organic acids and is more common in pyrolysis bio-oils.
For biocrudes upgrading, hydrotreating deals principally with hydro-deoxygenation. The main
reactions taking place in the catalytic hydrotreating of biocrudes are presented in Figure 2.9. The
oxygen is removed trough three main products; in the form of water from hydro-deoxygenation
reactions, CO2, and CO product of decarboxylation and decarbonylation reaction, respectively.
Even though these reactions do not consume hydrogen, they are undesired in some extent. Jones
et al. [182] investigated the reduction of carbon yield product of decarboxylation and
decarbonylation reactions and how it affects the overall upgrading [183]. Through hydro-
deoxygenation and dehydration reactions, the oxygen is removed in the form of water. During
these reactions, hydrogen is consumed, and it was found that the oxygen removal is more effective
at severe conditions of temperature and pressure [184]. Typical conditions during hydrotreating
are in the range of 50-200 bar and 250-400 °C [156]. At high process temperatures, hydrocracking
reaction occurs transforming large biocrude constituents into smaller refined molecules. Elliot et
al. carried out hydrocracking experiments for the upgrading of pyrolysis bio-oils; they found a
high hydrogen consumption during hydrocracking [185].
Hydrogen consumption is one of the main concerns of hydrotreating. Hydrogen is expensive, and
large amounts must be used when the process required high hydrogen pressures. Grange et al.
[186] studied the influence of temperature on hydrogen consumption during hydrotreating
processes. It was found that hydrogen consumption increases at high operational temperatures. If
total oxygen removal is desired, the process must be accomplished at temperatures above 350 °C.
In the same work, it was concluded that hydrogenation of olefins, aldehydes and ketones is specific
and low hydrogen is consumed. Whereas, the hydrogenation reactions at higher temperatures are
less specific and more hydrogen is consumed because other adjacent groups are hydrogenated at
the same time, such as aromatic rings. At low severity conditions, just partial upgrading of the
biocrudes can be achieved. Also, the hydrogen consumption and its relation with the degree of
deoxygenation in hydrotreating were studied by Venderbosch et al. [187]. They concluded that
hydrogen consumption increases when more than 50% of the oxygen is removed from the

53
feedstock. These conclusions reinforce the fact that hydrogen should be balanced in the
hydrotreating process because it is an expensive input in the process, especially if the intention is
to reach an industrial scale process. A hydrogen balance is performed accounting the hydrogen
present in the feedstock (elemental analysis), the hydrogen fed, also in the products (gases +water
produced + hydrogen remaining in the upgraded products in form of water)
Finally, hydrotreating process has some drawbacks regarding the amount of char, coke (8-25 wt%),
and tar that are produced due to the sulphide catalyst application which may result in the catalyst
deactivation and reactor clogging [45].

Figure 2.9. Main reactions in hydrotreating processes of biocrudes, adapted from [184].

Several research studies have been performed in hydrotreating of high oxygenated renewable
biocrudes and bio-oils; it was concluded that hydrotreating is a viable technology for the
improvement of high oxygenated biocrudes. For example, hydrotreating of pyrolysis bio-oils has
been extensively studied by several authors including Elliot et al. [148], Venderboch et al. [187],
Zhang et al. [188], Mortensen et al. [189], among others.
Hydrotreating is a process that has proved to be more efficient when dual-stage processing is
employed. Elliot [152] proposed two catalytic reaction stages where the feed is first subjected to

54
mild temperatures over cobalt-molybdenum (CoMo), and then the product oil is further treated at
higher temperatures and lower space velocity [150].
Scheele [190] performed studies about the catalytic upgrading of HydrofactionTM biocrude through
the screening of process variables in a continuous mode hydrotreater unit; three different catalysts
produced with in-house technology were studied. Total TAN reduction was achieved as well as a
high viscosity reduction. In a similar upgrading study carried out in continuous mode, Trujillo
[155] performed the upgrading of Hydrofaction in a two-stage process. First, a mild catalytic
upgrading step of the biocrude was performed, achieving a 59% hydro-deoxygenation and more
than 98% TAN reduction. The partially upgraded biocrude was subjected to a catalytic steam
cracking step performed in a different unit where the source of hydrogen was obtained from the
catalytic splitting of water.

2.5. Problem Statement


The HydrofactionTM biocrude to be upgraded, has better characteristics than other biocrudes, for
example, less oxygen content, high stability and high heating value compared to pyrolysis oils or
other biofuels produced with similar thermochemical processes and feedstocks. Nonetheless,
HydrofactionTM biocrude requires improvements that can be accomplished through catalytic
hydrotreating; which is a well-established technology mainly used for upgrading of fossil fuels but
extensible for the upgrading of renewable biocrudes.
The screening of operational conditions such as temperature, pressure, H2/oil ratio, space velocity,
and different configurations with heterogeneous NiMo and CoMo (two-stage process) in trickle
bed reactors, allows the improvement of the feedstock. Then, in order to obtain a product with low
or zero oxygen content, improved distillation profile and low viscosity, the construction of a bench
hydrotreater was required to perform the upgrading of the renewable biocrudes. The understanding
of the hydrodynamics of the unit can be accomplished with the commissioning of the unit. This
step involves the performing of hot and cold tests. For security purposes, leak tests must be
performed.
This research project is in the stage of proof-of-concepts or demonstration at bench-scale the
catalytic hydrotreating of the HydrofactionTM biocrude. Then, the data acquired during the
upgrading process of the feedstock in continuous bench-scale catalytic hydrotreaters will
contribute with the main goals of the research.

55
Besides the upgrading of HydrofactionTM biocrude, another important goal is to scale-up the
upgrading process of the biocrude. Then, the bench-scale continuous mode catalytic hydrotreater
constructed and commissioned at the University of Alberta is a realistic pathway to investigate the
upgrading of the renewable biocrude. Even though, the research project is a proof-of-concept stage
(demonstration bench scale) and is not currently used at an industrial scale with the current
feedstock, the vital information and expertise (know-how) acquired during the upgrading process
performed through this investigation, was intended to demonstrate a potential pathway process for
optimization, scaling-up and upgrading of HydrofactionTM renewable biocrude.

56
Chapter 3: Experimental Methods

This chapter presents the description of materials, experimental set-ups, characterization methods
and experimental procedures used in this research project. The materials, including the
characterization of the feedstocks to be upgraded, the different chemicals and gases used in this
research are presented in Section 3.1. In the apparatus Section 3.2, a detailed description of the
experimental set-up of the bench hydrotreater pilot plant is included, as well as the description of
the different equipment and analytical methods used for characterization of biocrudes and
upgraded samples. Section 3.3 includes a detailed bench hydrotreater operation, reactor assembly
and a brief description of the activation of the catalysts. In Section 3.4, the experimental design
developed to investigate the different operational conditions and catalysts through two independent
sets of experiments is presented.

3.1. Materials

• HydrofactionTM biocrudes

Steeper Energy provided two different biocrudes for upgrading; these feedstocks were obtained
from forestry residue through the Steeper’s energy hydrothermal liquefaction technology
HydrofactionTM. Details about the production of HydrofactionTM biocrude were presented in
chapter 2. The samples were named as biocrude-1 and biocrude-2, and their properties are
presented in Tables 4.1 and 5.1 respectively. The results of higher heating value and elemental
analysis were performed at the University of Calgary. Additional analysis such as density,
viscosity and total acid number TAN were performed to the biocrudes at the University of Alberta
and are presented in the results chapters 4 and 5.

• Catalysts

The catalysts used in this investigation, presented in the Table 3.3 are well known for improving
the properties of the feedstocks by adding hydrogen and improving the physiochemical properties
in this case of the heavy renewable feedstock. Due to a standard confidentiality agreement between
Steeper Energy and various catalyst companies that cannot be mentioned in this document; also,
this manuscript won't include any information of catalysts. No characterization on the fresh or used
catalysts was performed as part of the agreement.
57
Table 3.1. Catalysts tested for the upgrading of HydrofactionTM biocrudes.

Catalyst Type Support


A (CoMo) γ-Al2O3
B (NiMo) γ-Al2O3
C (NiMo) γ-Al2O3
D (NiMo) γ-Al2O3
E*
* Catalyst type was not specified by the catalyst provider.
• Chemicals
The chemicals used in this research project are tabulated in Table 3.4, were used in different
activities which are specified in the table. All the chemicals were stored in the appropriate shelves
for chemicals, and all the material safety data sheets (MSDS) were available in the laboratory. The
biocrude feedstocks are described in chapters 4 and 5, along with the appropriate characterization
results.

Table 3.2. Different chemicals used in different activities during the investigation

Chemical Supplier Grade/purity Purpose


Toluene Fisher Scientific ≥99.5 % Purity cleaning, TAN analysis.

2-propanol Fisher Scientific Purified cleaning, TAN analysis.


Dimethyl sulphide Sigma Aldrich ≥99 % Purity Sulphur agent to spike the
feedstock. Catalyst activation.
Dimethyl Disulphide Sigma Aldrich 99% Catalysts activation
CS2 Sigma Aldrich ≥99 % Purity Simulated distillation analysis

Mineral Oil Fisher Scientific - Commissioning (hot tests)

Pentane Thermo Fisher Extra dry Commissioning (hot tests)

Diesel Husky Premium Catalysts activation

58
• Gases

In this study, high purity gases were used for the investigation. The gases tabulated in Table 3.5
were adequately stored; the high pressure compressed hydrogen was stored in a ventilated shelve.
All the security protocols for the manipulation of compressed gases were followed.
Table 3. 3. Gases used during the investigation

Chemical Supplier Grade Purpose

Hydrogen 99.999 % purity Gas feed for hydrotreating, combustible gas for
SimDist analysis, pressurizing gas.
Nitrogen Praxair 99.999% purity TGA, SimDist analysis, pressurizing gas
Helium > 99% Carrier gas GC and SimDist analysis

Air 99.999% purity SimDist analysis

3.2. Apparatus

3.2.1. Description of the bench hydrotreater pilot plant

The principal equipment used for the consecution of this investigation was the continuous bench
catalytic hydrotreater pilot plant. This hydrotreater was designed and built at the University of
Alberta to upgrade the HydrofactionTM biocrudes. The unit was assembled inside of a modified
fume hood with continuous removal of gases. The main reasons for having the plant in a closed
fume hood with constant extraction was for the risk of working with high hydrogen pressures, as
well as for the risk of the H2S produced during the upgrading processes. The hydrotreater is
divided into three main zones: feed section, reaction section, and product collection; each section
is described below. The P&ID diagram of the hydrotreating plant is shown in Figure 3.1, which
specifies each section. The bench hydrotreater can be operated to a maximum temperature and
pressure of 400 ºC and 100 bar respectively.

59
Figure 3.1. Simplified schematic diagram of the catalytic hydrotreater (double lines represent the
heated and insulated lines in the setup).

60
Feed section:
It consists of a feed steel tank (TK-1), a Teledyne dual syringe pump (P-101 and P-102), hydrogen
mass flow controller (MFC-1), and blowdown drum tank (TK-4). The main components are
described below:

▪ The feed tank is a 7 L stainless steel vessel equipped with an electrical heating blanket and a
pressure relief valve set at 50 psi. It has two inlet ports; one used to inject nitrogen and the
other to place a thermocouple.
▪ A Teledyne Isco 500 D dual pump equipped with two 500 ml cylinders. This pump can deliver
continuous flow in a range from 1µL/min to 204 mL/min with a maximum operating pressure
of 3750 psig. The pump is jacketed and heated with recirculated oil.
▪ The hydrogen smart mass flow controller is a Brooks Smart Instrument 5850 EM well suited
for flow rates from 0.001 standard H2 ml/min to 1000 standard H2 ml/min. The mass flow can
be operated in manual mode or using the LabView interface.
▪ In this section, a solenoid valve SV -1was installed along with a high-pressure gauge HPG-1,
the gauge and the solenoid valve are connected to the OPTO 22 control system (SNAP-PAC-
R1). The desired maximum pressure can be set in the gauge manually, and in the case of
overpressure, the system OPTO 22 reads the signal and automatically close the solenoid valve
which interrupts the hydrogen flow to the hydrotreater unit and stop the high-pressure pump
at the same time.
▪ The blowdown drum is made of stainless steel and has a capacity of 20 L.
▪ For security, the feed section is provided with a rupture disc RD-1 (2500 psi) located in the
high-pressure cross joining the blowdown drum and the feed line located between the pump
and the reactor 1.
▪ The lines are all ¼ inch I.D. stainless steel pipes insulated and heated with heating cords. The
temperature is monitored using wall thermocouples located along the lines and the OPTO 22
system controls the temperature of these lines.

Reaction Section:
This is the principal part of the hydrotreating process, and it includes two down-flow packed bed
catalytic reactors (R1 and R2) that can be operated in series mode. These reactors were

61
disassembled from the bench hydrotreater, repacked when a different catalyst configuration was
tested and then installed again. The tubular reactors have a capacity of 181.5 ml each one, with a
length of 64 cm (25.2 inches) and an internal diameter of 1.9 cm (3/4 inch nominal diameter). The
hydrotreater unit has a bypass located between reactor one and two. The bypass allows operation
in continuous mode with just one reactor when needed, depending on the purpose of the
experiment. The bypass line is heated and well insulated; the temperature of all the ¼ inch I.D.
high-pressure stainless-steel lines is monitored with wall thermocouples. The lines are heated with
heating cables and insulated. The pressure across the reaction section is monitored using
manometers and pressure transducers located at the beginning and the end of each reactor, as
shown in Figure 3.1, whit a pressure transducer PI -4, and with the pressure transducer of the high-
pressure pump; monitored through the LabView program which is linked with the OPTO 22
system. The lines of the reaction section are all ¼ inch I.D. stainless steel pipe.
Figure 3.2 shows the schematic of a packed tubular reactor. The reactors were custom made at the
University of Alberta with ¾” inch high-pressure stainless-steel tube, 0.065” wall, and two high-
pressure Tee connectors.

Figure 3.2. Schematic of the packed tubular reactor

The temperature profile of the catalytic beds was monitored using an OMEGA multipoint
thermocouple placed inside of the reactor. These thermocouples can withstand temperatures

62
between 0 to 650 ºC and were placed at the center of the reactors. These thermocouple sensors
have 6 measurements points, which are spaced, as shown in Figure 3.3. The multipoint probe was
custom made by OMEGA.

Figure 3.3. OMEGA multipoint thermocouple probe. Adapted from [191].

The reactors are rated for a maximum temperature of 400 ºC and are electrically heated with 3
independent band heaters (brand O.E.M. Heaters) located equidistant to each other along the
reactors. The band heaters are controlled with the internal temperature acquisition points TIC 1,
TIC 2, and TIC 3 in reactor one, and TIC 4, TIC 5, and TIC 6 in reactor two (Figure 3.1 and 3.2).
These six points allowed the data acquisition system OPTO 22 installed in the hydrotreater unit to
controlling and monitoring the desired reaction temperature profile along the catalytic beds
packed. The reactors are well insulated with aluminum cases filled with glass wood.
Hydrogen or nitrogen can be supplied directly to reactor two through a dedicated line shown in
Figure 3.1. It can be done by manipulating the three-way valve TWV-1 and the needle valve V-
108; this flow was not controlled with the mass flow controller.

Product Collection Section:


This section was designed to recover the gas, light liquid and heavy liquid products obtained from
the reaction. It consists of the following units: separator (E-3), condenser (E-4), product filters (F-
2 and F-3), back pressure regulator (BPV-1), two light liquid tanks (LTK-3 and LTK-4), two heavy

63
liquid tanks (HTK-1 and HTK-2), and a wet gas flow meter (GFM) located at the end of the exhaust
gas line (vent).
▪ The separator is a 316 stainless steel horizontal vessel (3/4 inch nominal diameter) with a
capacity of 60 ml.
▪ The condenser is a shell-tube heat exchanger made of stainless steel and uses cold water at 4
ºC using a chiller.
▪ Two Swagelok filters made of 316 stainless steel and equipped with a filter cartridge of 15
microns.
▪ Two 316 stainless steel vertical vessels for light products collection (LTK-3 and LTK-4), with
a volumetric capacity of 500 ml and it is rated for 2800 psig.
▪ Two 316 stainless steel vertical vessels for heavy products collection (HTK-1 and HTK-2),
with a volumetric capacity of 300 ml and it is rated for 4000 psig.
▪ A Tescom 26-1700 series back pressure regulator (BPV-1).
▪ A high precision Ritter drum-type wet gas meter.

The double lines shown in figure 3.1 are high-pressure ¼ inch lines heated and well insulated;
other lines of the section are ¼ inch as well. The pressure, in the heavy products collection zone
two pressure gauges were installed and the temperature of the heavy tanks and lines was
monitored using wall thermocouples. Finally, in the zone of light products collection, the system
worked at atmospheric pressure and room temperature.

3.2.2. Analytical equipment used in the characterization of biocrudes and upgraded


samples.
This section is intended to describe the analytical equipment employed to characterize the liquids
and gas samples obtained from the hydrotreater unit along the research study. A brief description
of the analytical techniques used in this investigation is presented.

Equipment used for Gas Samples Analysis

Gas Chromatography: The on-line GC analysis was carried out using a Varian CP-3800 gas
chromatograph equipped with a Thermal Conductivity Detector (TCD), and HP-PLOT/Q column

64
(L: 30m, D:0.32 mm, Film: 20 µm) from Agilent. The GC operational principle is based on the
differences between the thermal conductivity in the gas carrier and the gas samples analyzed.
Every compound has a specific thermal conductivity, which can be detected and measured with
the thermal conductivity detector TCD installed in the GC. The carrier gas used in all the
experiments was helium.

Equipment used for Liquid Samples Analysis

Density: The density measurement was conducted using an Anton Paar DMA 4500M following
the ASTM D5002 [192]. The instrument precision is ±0.00001 g/cm3 with an accuracy of ±0.00005
g/cm3. The density was measured at three temperatures 20, 40, and 60 ºC. The results were reported
with a very low error lower than 0.2%.

Viscosity: Viscosity measurements were carried out at 40 °C with a Brookfield DVE series
viscometer. The instrument is provided with a double jacketed chamber to ensure the isothermal
condition. Water at 40 °C was constantly flowing from a thermal bath to the chamber, the water
was passed through using a silicon hose. Viscosity measurement was conducted at 20, 40 and 60
o
C following the ASTM D 445 [193].

Total Acid Number: TAN analysis was conducted in a digital titrator Mettler Toledo T50
following the Standard Test Method for Acid Number of Petroleum Products by Potentiometric
Titration ASTM D664-11a [194]. The methodology followed consist of weighing in a high
precision scale certain amount of sample (depending on the acidity of the sample). A solution of
50% of toluene, 45% of 2-propanol, and 5% of distilled water was prepared. This solution was the
blank solution used before each analysis. Next, 60 ml of the standard solution was added to the
weighted sample and placed in the titration chamber with constant agitation.

Elemental Analysis CHN: This analysis was performed in a Perkin Elmer 2400 CHN Analyzer,
following the standard ASTM D5291 [195]. A standard was used as a reference to calculate the
carbon, hydrogen and nitrogen composition of the feedstocks and upgraded samples. The oxygen

65
content of the feedstock and upgraded samples was calculated indirectly (by difference) and
reported on a dry basis.

Fourier Transform Infra-Red Spectroscopy Analysis (FTIR): this analytical technique was
performed with a Nicolet 3700 instrument Thermo Electron brand; the source of light installed in
the instrument was ETC EverGlo mid-IR, the spectral signature read from the samples is reported
between 500 to 4000 cm-1. For each analysis, 128 scans were acquired with a resolution of 2cm-1.
The standard followed to perform this analysis was the ASTM D6348 [196].

Higher Heating Value: this analytical technique was achieved in a 6100- calorimeter bomb from
Parr. The standard followed is the ASTM D 2240-17 [197]. The analysis was performed by adding
1 gram of biocrude in a crucible; then the calorimeter was filled with pure oxygen at 30 bars. The
bomb was placed in the adiabatic calorimetric equipment filled with water. The sample was ignited
with electricity. As a result, the temperature of the water placed outside of the chamber increased,
and it allows the calculation of the HHV. The heat capacity of the calorimeter bomb was
determined with benzoic acid as a reference; the errors of the analysis were in the range of ±
60kJ/kg.

High Temperature Simulated Distillation (HTSD): was performed in with a Bruker 450-GC
gas chromatograph following the standard ASTM D 7169-18 [198]. This analytic technique
provides the boiling point distribution of biocrudes up to 750 °C (high-temperature gas
chromatography). The apparatus to perform the analysis is provided with a gas chromatograph
(GC), and a flame ionization detector (FID). The methodology consists of dissolving 0.15 grams
of biocrude in 20 ml of CS2. Next, 1µL of this solution is injected into the temperature-
programmed chromatograph column. The technique is useful for characterization of biocrudes
elucidating the distillation cuts of the upgraded samples in the different established distillation
ranges, such as gasoline, diesel, LVGO, HVGO and residue.

Thermogravimetric Analysis (TGA): is performed to determine the heat and weight changes of
liquid or solid samples when they are under a constant flow of gas and subjected to a stablished
temperature increasing. TGA analysis was carried out in a TA Instruments SDT Q600; the analysis

66
was performed by adding 20 mg of upgraded biocrude into an alumina crucible. Then it was heated
up from room temperature to 900 °C using a linear heating rate of 10 °C/ min. It was accomplished
under a constant nitrogen flow of 100 cm3/min. The standard followed was the ASTM E 2402 -11
[199].

3.3. Experimental Procedure

The bench hydrotreater pilot plant is the principal equipment used through this investigation. This
section is intended to illustrate the general operational procedure to operate the bench hydrotreater
pilot plant.

3.3.1. Hydrotreating unit operation

This section describes the procedure used to start-up and operate the bench hydrotreater pilot plant
(see Figure 3.1). The security protocols such as standard operating procedure and the safe work
procedure prepared for a safe operation were taking into account for every operation of the
hydrotreater; these documents are included in Appendix B. The first step performed was the leak
test; it was achieved by pressurizing the bench-scale pilot plant using hydrogen. It was pressurized
to 100 bars, and the leak test was considered to be successful when the pressure remained constant
for an extended period. During the test, the pressure was monitored with the pressure transducer
and the pressure gauges located at different points of the hydrotreater unit and the pressure
transducer of the pumps. Besides, a hydrogen leak detector was used to verify the presence of any
hydrogen leak across the plant. Finally, the hydrotreater unit was depressurized by slightly opening
the back-pressure regulator (BPV-1), until it reached the subsequent desired experimental pressure.
After finishing a set of experiments, the back-pressure regulator was used to decrease the pressure
of the entire hydrotreater and be able to remove the reactors from the unit and packed them with a
different set of catalysts.
The biocrude feed was charged manually into the feed tank at ambient conditions. The biocrude
was always spiked with 200 ppm of butanethiol before starting any campaign. Then it was mixed
thoroughly using a mixing propeller connected to a manual drill. The sulfur addition is required to
avoid catalysts deactivation because, the catalysts used in this study were sulfided forms of NiMo
and CoMo catalysts.

67
The tank was sealed and pressurized with nitrogen at 30 psi to check leaks. It was then vented, and
more nitrogen was injected up to reach 30 psi again. It was done twice in order to create an inert
atmosphere. The pressure inside the tank was kept at 30 psi by opening the valves RV-1 and V-
101. Also, the nitrogen regulator was adjusted at 30 psi. The valve V-103 was kept close to avoid
sending nitrogen to the reaction zone, as can be seen in Figure 3.1. Later, the feed tank was heated
at 60 ºC using a blanket heater in order to reduce the viscosity of the feedstock and facilitate the
pumping process.
The next step was to pressurize the whole hydrotreater unit by flowing hydrogen. To do so, the
valves RV-2 and TWV-1 were open to provide a continuous gas supply. The mass flow controller
(MFC-1) was set to a hydrogen flow rate depending on the operating hydrogen to oil ratio. The
back-pressure regulator BPV-1 was set at the operating pressure by manually closing it. Once the
operating pressure was reached, the lines before the reactor one were kept at 180 ºC by setting the
temperature in the LabView program linked with the OPTO 22 system which controlled the
temperature set in the whole hydrotreater. The temperatures selected in the different section of the
unit were the result of performing hot and cold tests performed during the commissioning of the
hydrotreater unit.
On the other hand, the temperature of the reactors was ramped up to the reaction temperature. The
lines between the reactors and those after reactor two were heated at 250 °C. The lines located at
the product zone were kept at 180 °C, the heavy tanks were heated as well with heating tape at 80
°C. The objective of heating the tanks was to facilitate the sampling process during mass balances.
Before pumping the feed through the hydrotreater unit, the position of the valves located along the
reaction and the product collection sections were first checked.
In the reaction section, the valves V-105, V-107 and V-109 were opened, and valve V-106 was
closed when two packed reactors R1 and R2 were installed in the unit (series mode process). On
the contrary, valves V-107 and V-109 were closed, and valve V-106 was opened when only reactor
R1 was put in service to run an experiment. To ensure a continuous flow of the feed from the feed
tank to the reactor (R1) valves V-102 and V-105 were kept open.
The hydrotreater unit was designed for prolonged continuous experiments. In the collection
section, four tanks were installed for the collection of the upgraded products during the upgrading
experiments performed. Two tanks were installed in the heavy liquids collection zone and two
more in the lights liquid collection zone, in that way was possible to run experiments in continuous

68
mode. The valves V-111 and V-112 were kept open, and valves, V-113, V-114, V-115, and V-
116 were closed in the side of the heavy tanks. At the same time, for the collection of the light
products, the valves V-119 and V-120 were kept open, while valves V-121, V-122, V-123 and V-
124 were kept closed. Then, in a long experiment where more no more than 600 to 700 ml were
pumped, the products were collected and a mass balance was performed. Details about mass
balance calculation and procedure are presented in Section 3.5. To continue the experiment without
stooping the experiment, the tanks were switched to HTK-2 and LTK-4. The capacity of the tanks
HTK-1 is 300 ml, and LTK-3 is 500 ml. It can be noticed that there was a 100 ml gap for security
purposes.
To switch the tanks, before switching the tanks the pressure in the empty heavy tank should be
increased with nitrogen to the pressure of the experiment by opening the valve V-110. This
procedure was performed just in the heavy tanks because the light tanks worked at atmospheric
pressure.
The tanks HTK-1 and HTK-2 were heated at 80 ºC with heating cords in order to facilitate the
sampling (viscosity reduction) and removing the light ends that might be trapped within the heavy
product.
Once the position of the valves was verified, and the system had no leaks the feed was pumped at
the operating flow rate from the feed tank through the reactors using the high-pressure pumps. The
temperature in the reactors at this stage was already brought up and kept to the desired reaction
temperature. The temperature inside the reactors was measured by OMEGA thermocouples at
points TIC1, TIC2, and TIC3 in reactor one and TIC4, TIC5, and TIC6 in reactor two. The reactors
were considered to have reached reaction temperature when the internal temperatures were within
±0.3% of the setpoint or less than 0.5 °C variation.
The upgraded products entered a horizontal separator (E-3) to remove the produced water from
the upgraded products. The separator was operated 10 °C above the water saturation temperature
at the operating pressure. The dewatered upgraded products were sent to the tank HTK-1 or HTK-
2 where the sample was kept at 80 ºC. Extra gases that may have generated were sent back to the
gas flow line just before entering the condenser E-4. The vapour phase water with the light
products flowed through the condenser were cooled down at 4 ºC. The condensed light products
and gases were passed through a filter (F-2 or F-3) and finally were collected in the tank LTK-3
or LTK-4 at ambient temperature and atmospheric pressure. The gases evolved from the light tanks

69
were later passed through two scrubbers (S-3 and S-4) to remove the hydrogen sulphide (H2S)
produced during the hydrotreating process. The scrubbers were filled with a volumetric solution
at 50/50 of sodium hypochlorite and water that neutralizes the H2S. The composition of the gases
was measured online by a GC chromatograph that was directly connected to the gas stream. The
gas flow rate (mainly hydrogen and produced gases in the process) was measured using a wet gas
flow meter (GFM). The steps mentioned were common for all the experiments performed through
this study.

3.3.2. Reactor assembly.

All the reactors used in this study consisted of three main parts: the top and bottom high-pressure
Tee connectors and the reactor tube. The reactors are assembled with an internal multipoint
thermocouple, which measures the temperature inside the reactor in six different equidistant points
(see Figure 3.5). Furthermore, six ceramic heaters located outside of the wall of the reactor were
used to elevate and maintain the temperature of each reactor during the hydrotreating experiments.

To set up the reactor, a metallic mesh was inserted around the tip of the thermocouple and moved
down until it reaches the bottom of the reactor tube. Then, the OMEGA multipoint thermocouple
was connected to the bottom Tee connector, which was later introduced through the lower part of
the reactor tube. Once the bottom Tee was tightened to the reactor tube, a plastic hose with a
diameter slightly bigger than the thermocouple was placed in the middle of the reactor and outside
of the thermocouple in order to hold the thermocouple at the center of the reactor during the
catalyst loading. Then using a funnel, different layers of pure or diluted catalysts were added
(depending on the experiment). The packed layers were separated using thin layers of glass-wool.
Once the reactor was packed with the desired catalytic beds, another metallic mesh was placed at
the top of the reactor to keep the catalytic bed in a fixed position. Finally, the upper tee connector
was tightened to the reactor. After that, the reactor was placed in the bench hydrotreater pilot plant.
If the experiment requires two reactors in series then, the two reactors were packed following the
same procedure mentioned and then, installed in the hydrotreater.
The configuration of the reactors used in the first set of experiments is presented in Figure 3.4. In
this case, two different scenarios were accomplished. During the first stage scenario carried out

70
with one reactor packed with CoMo catalyst (left-hand-side Figure 3.4), several experiments were
done. In one of the experiments, the conditions were kept and two kilograms of partially upgraded
biocrude were collected.

Figure 3.4. Configuration of the packed bed reactors with CoMo and NiMo catalysts used in the
first set of experiments. The First and second stage were performed independently.

During the second stage scenario of the first set of experiments, a reactor was packed with NiMo
catalyst and used (right-hand-side Figure 3.5) to perform the hydrotreating experiments. The
feedstock used was the two kilograms of partially upgraded biocrude collected in the first stage
scenario. The total amount of CoMo catalyst (A) used in the first stage scenario (left-hand-side
Figure 3.5) was 44.3 g (pure and diluted beds). The catalytic beds were custom-made at the
University of Alberta, consisting of three layers, as presented in Figure 3.5. For both reactors, a
dilute bed of catalyst SiC at a weight ratio of 25/75 respectively, was packed at the top of the
reactors to avoid an abrupt increase in the temperature due to the exothermicity of the reaction, the
second layer was pure CoMo catalysts (A), in the reactor to the left and NiMo catalyst (B) for the

71
reactor to the right-hand-side. Finally, in the bottom pure carborundum was packed in both
reactors.
In the case of the reactors used in the second stage scenario, the reactor to the right (R2) was
packed with 48.8 g of NiMo catalyst (D) diluted with SiC at 75% weight ratio and a configuration
of catalyst E in the bottom of the reactor. The configuration of the packed reactors used is shown
in Figure 3.5. In this set of experiments, the reactors R1 and R2 worked in series mode during all
the experiments accomplished.

Figure 3.5. Configuration of the packed reactors R1 and R2 installed in series mode in the bench
hydrotreater pilot plant. (Second set of upgrading experiments). * Ratio of Catalyst D and E and
the bed configuration cannot be disclosed.

The catalyst C packed in reactor one is a catalyst used in hydrotreatment as guard bed protecting
the main catalytic beds from poisoning or fouling. The main beds, in this case, were packed in the
second reactor. The catalyst C has a hydrotreating function because it is good removing
heteroatoms and metals, but in this study was used for oxygen removal through HDO reactions.

72
As noted in Figure 3.5, the reactor contains a mixture of catalyst D and E arranged in a non-
disclosable configuration.
Catalysts D and E packed in reactor two, are commercial catalysts with high hydrogenation
activity; these catalysts are used in the oil industry for its good heteroatoms removal function, de-
aromatization, and to produce high-quality diesel fraction.
Using the total amount of catalyst added in the different set of experiments, the biocrude ratio to
be pumped was calculated using Equation 3.1. Knowing the amount of catalyst loaded, and the
desired weight hourly space velocity (WHSV), the oil mass flow rate was calculated using the
following equation:

𝐵𝑖𝑜𝑐𝑟𝑢𝑑𝑒 𝑚𝑎𝑠𝑠 𝑓𝑙𝑜𝑤 𝑟𝑎𝑡𝑒


𝑊𝐻𝑆𝑉 = Equation 3.1
𝑀𝑎𝑠𝑠 𝑜𝑓 𝑐𝑎𝑡𝑎𝑙𝑦𝑠𝑡𝑠 (𝑊𝑐𝑎𝑡 )

Where 𝐹𝑜𝑖𝑙 stands for the biocrude flow (g/min) and Wcat for the mass of catalyst (g). For each
experiment, these values were kept constant. A different experiment started when any of the
conditions screened was changed.

3.3.3. Catalyst activation.

Steeper Energy distributed previously established catalyst activation strategies. At the time of
writing, the companies that established these procedures cannot be named in the current document.
Catalyst A and B were provided in an activated state where as catalyst C, D and E were activated
via the procedure described below:
After the reactors were packed and installed in the hydrotreater unit, a leak test was performed. It
was accomplished by maintaining the pressure of the hydrotreater for two hours at 50 bars. The
hydrotreater was pressurized with nitrogen. Next, a purge step was made by flushing more nitrogen
through the reactors for about 30 min. After that, the system was pressurized at a hydrogen pressure
of 45 bars. The pressure was kept verifying that no leaks were present in the unit.
The next step was to subject the catalysts beds to a sulfiding process. It was accomplished by
passing Diesel spiked with 2.5 wt. % of dimethyl disulphide over the catalyst beds in the presence
of hydrogen. The temperature of the reactors was ramped up by following an established
temperature profile. It includes heating ramp intervals, temperature holds, and cooling cycles.

73
3.4. Experimental Design

In this investigation, two independent sets of experiments were carried out in order to evaluate
catalytic hydrotreating processes at different operational conditions and test the performance of
the catalysts used. The two independent sets of experiments accomplished differed in the feedstock
and in the type of catalysts that were used. The experimental protocols proposed for each set of
experiments are provided herein. The experimental protocols developed were based on one factor-
at-the-time design. This methodology was used for this investigation because the length of the
experiments performed was in all cases for more than 24 hours, including the stabilization time,
and several operational conditions were investigated. Each experiment was considered in steady-
state when all the parameters such as temperature, pressure, space velocity and H2/oil ratio were
stable.

3.4.1. Campaign #1: upgrading of biocrude-1 in the presence of pre-sulfided CoMo and
NiMo catalysts.

This campaign consists of two separated hydrotreating stages scenarios each one performed
independently using one catalytic reactor as was explained in the reactor assembly Section 3.3.2.
The experiments in the first stage were achieved using the biocrude-1. This feedstock was
subjected to catalytic hydrotreating using a reactor packed with a pre-sulfided CoMo catalyst A.
The second independent stage scenario of campaign #1, was performed using a partially upgraded
sample produced in the first stage. This feedstock was subjected to catalytic hydrotreating using
one reactor packed with NiMo catalyst B. In Table 3.4, the experimental plan that was designed
and followed to evaluate the effect of the operating conditions over the catalytic hydrotreating of
HydrofactionTM biocrude-1 and the partially upgraded sample is presented.

For the first stage scenario accomplished with CoMo catalyst, eleven experiments at different
conditions were performed as can be seen in the top of Figure 3.6 and are tabulated in the
experimental protocol Table 3.4.

74
Figure 3.6. Operational conditions screened during the catalytic hydrotreating in the presence of
CoMo and NiMo catalysts. First and second stage scenarios were performed independently
(stage one with CoMo and stage two with NiMo).

A partially upgraded sample collected in step # 2 was then used as a feedstock in the second
hydrotreating stage scenario in the presence of NiMo catalyst. In stage 2, two experiments were
performed, and each lasted 24 hours (including stabilization times). The conditions tested are
shown in Table 3.5 and at the bottom of Figure 3.6.

75
Table 3.4. Experimental plan followed for the upgrading of the biocrude-1 in the presence of
pre-sulfided CoMo catalyst.

Step Temperature Pressure WHSV H2/oil Ratio Experiment


# (ºC) (bar) (h-1) (Nm3/m3) duration (h)
1 370 24
95 0.3 900
2 350 192
3 320 274
4 320 12
5 350 36
6 370 80 0.3 900 36
7 350 12
8 350 0.3 16
9 370 0.3 18
95 1300
10 350 0.2 18
11 350
0.5 18

As observed in Table 3.6, the conditions of steps 5 and 7 are the same. It was done to check the
catalyst deactivation over time and to verify the reproducibility of the results. The results of these
experiments are presented in chapter 4. °

Table 3.5. Experimental protocol followed for the upgrading of the biocrude-1 in the presence of
pre-sulfided NiMo catalyst.

Step Temperature Pressure WHSV H2/oil Ratio Experiment


# (ºC) (bar) (h-1) (Nm3/m3) duration (h)
1 320 24
95 0.3 900
2 350 24

The duration time tabulated above means the time or the duration of each experiment at fixed
conditions. The duration time does not include the stabilization time. In most cases, it was the
duration of one or two mass balances taken during the experiment at the same condition. The
details about a mass balance calculation and procedure are presented in Appendix A.

76
3.3.2. Campaign #2: upgrading of HydrofactionTM biocrude-2 using two down-flow
reactors in series in the presence of NiMo Catalysts.
In this campaign, a set of experiments were performed to study the upgrading of the biocrude-2
through a catalytic hydrotreating process performed in series mode with two down-flow fixed bed
reactors. For this set of experiments, biocrude-2 was used because there was an insufficient amount
of biocrude-1 remaining after the first campaign. Also, three different NiMo catalysts were used
for campaign 2, at the request of the sponsor (Steeper Energy).

Figure 3.7. Operational conditions screened during the catalytic hydrotreating performed in
series mode in the presence of activated NiMo catalysts.
The NiMo catalyst C packed in reactor one was diluted at 75 wt% with silicon carbide. The second
reactor with catalysts D and E as can be seen in Section 3.3.2 reactor assembly.
In Figure 3.7 and Table 3.8, the experiments performed, and the conditions investigated in the
second campaign are presented.

77
The reactions of renewable biocrudes with hydrogen are highly exothermic, and the temperature
tends to rise very rapidly in the top of the reactor [200], especially at the beginning of the process
with fresh catalysts. Consequently, side effects may occur during the hydrotreating processes, such
as coke formation, fouling or plugins in the catalyst surface, further pressure drop and finally
deactivation of the catalytic beds. In this research project, the control of temperature is essential
because our study includes temperature screening. Then, it is necessary to perform the upgrading
at a fixed temperature, which must be maintained (isothermic) during the length of the experiment.
Moreover, hot spots in the catalytic beds must be avoided; a solution used to dissipate the heat
generated is diluting the catalytic beds with SiC, which is inert.

Figure 3.8. Experimental conditions proposed for the upgrading of the biocrude-2 in the
presence of pre-sulfided NiMo catalysts.

Step Temperature (ºC) WHSV H2/oil Ratio Experiment


# (h-1) (Nm3/m3) duration
R1 R2 (h)
1 320 24
2 350 12
320 0.3 900
3 370 24
4 370 36
5 370 12
320 0.5 900
6 350 12
7 320 350 12
0.2 900
8 320 370 12
9 320 350 12
0.2 1300
10 320 370 12
11 320 12
350 0.3 900
12 350 36
13 350 36
350 0.3 900
14 390 48

The conditions screened during campaign # 2 and the number of experiments are presented in
Table 3.8, where R1 and R2 correspond to reactors 1 and 2 placed in series during this campaign
# 2.

78
Chapter 4: Upgrading of Biocrude-1 through catalytic hydrotreating in the presence of
commercial CoMo and NiMo catalysts (campaign #1)

In this chapter, the upgrading results of HydrofactionTM renewable biocrude-1 through two
independent hydrotreating stages are presented. Typical hydrotreating operational conditions such
as temperature, space velocity, H2/oil ratio and pressure, were investigated following the
experimental protocols designed for each stage shown in Section 3.4.1 of chapter 3. The biocrude-
1 was produced from forestry residue through Steeper energy’s patented (HTL) HydrofactionTM
technology. Preliminary experiments (commissioning) were performed in the hydrotreater unit;
these results are presented in Section 4.1 of this chapter.
The hydrotreating catalysts NiMo and CoMo used in the experiments of campaign #1 have a high
hydro-deoxygenation activity. The amount of coke produced was found to be negligible after
completing all the catalytic experiments included in this chapter. Very few solids (coke-like) were
found inside the reactor, in the filters before the backpressure, and in the heavy and light product
tanks.
During campaign #1 (first and second stage scenarios), 13 experiments were performed. To do so,
the bench hydrotreater pilot plant ran smoothly for 24 days. All the mass balances taken to perform
further analysis were above 98% in specified time intervals, usually 8 to 12 hours. The details
about the calculation procedure for the mass balances can be found in Appendix A.
The screening results of temperature, space velocity, pressure and H 2/oil ratio over the upgrading
are included in this chapter. A study of the activity of the catalyst through experiments performed
at different times is included. A comparison of the results between the first stage scenario
performed with CoMo catalysts, and the results of upgrading results of the second stage scenario
performed with NiMo catalysts and partially upgraded biocrude-1, are included as well. Finally,
the results of the thermal effect (upgrading without catalysts) performed at four temperature levels
are discussed in this chapter.

4.1. Commissioning and equipment verification

The bench hydrotreater pilot plant was built at the University of Alberta, the unit was assembled
with 90% of new installations, and the rest are existing parts from previous hydrotreating setups.
The commissioning of the equipment was accomplished through hot and cold tests. In both cases,

79
the reactors 1 and 2 of the hydrotreater were packed with SiC. The unit was cold tested using
toluene and nitrogen with an approximate flow to the ones used in normal operations, for example,
space velocity of 0.3, and N2/oil ratio of 900 Nm3/m3; ambient temperature and atmospheric
pressure were used during the cold tests. In average, 95% of the toluene fed was recovered, and
100% of the gas was detected in the wet gas meter. The hot tests were achieved at conditions of
300 °C for both reactors packed with SiC, pressure of 30 bars, H2/oil ratio of 1500 Nm3/m3 of oil,
a feed flow of 0.6 ml/min and WHSV of 0.17 h-1The conditions temperature and feed flow were
manipulated to verify the response time of the bench hydrotreater. The feed for the hot tests was a
volumetric solution 90/10 of heavy mineral oil and pentane and the gas phase fed was hydrogen.
The pentane was added to check the performance of the separator and condenser. These tests were
performed with liquids with similar characteristics of the biocrude, trying to mimic the properties
of the biocrude. The goal of performing hot and cold tests was to learn about the behaviour of the
hydrotreating unit at different temperatures and pressures, as well as to understand the
hydrodynamics of the process.
After performing the hot tests, it was determined that the temperature of the heavy products tanks
is critical to obtain proper mass balances. The temperature should be controlled in such a way that
is high enough to facilitate the oil flow, but not too high, so evaporation or decomposition takes
place in those tanks. Furthermore, more than 95 % of heavy mineral oil was recovered, but no
pentane was collected in the light products tanks. It was found using the online GC, that all the
pentane was evaporated in the tanks due to the gas flow (hydrogen) through the liquid phase. At
the end, all the pentane fed was accounted. Finally, during the hot test, it was also determined that
the heating cords used are not homogenous and tend to deliver more heat in the middle, especially
during start-up. Most of the heating cords were re-installed, taking into account this heterogenicity.
The temperature control system was tuned by recording the temperatures read in all the wall
thermocouples and reactor heaters installed in the reactors during the hot tests. After that, the
control system technician from the University of Alberta adjusted the system OPTO 22 with the
information recorded. As a result, the heaters of the reactors and heating cords installed in the
bench hydrotreater were considered within specs and were tuned to a precision of ±0.5 oC when
the unit is at steady state.

80
4.2. Characterization of HydrofactionTM biocrude-1.

Some preliminary analysis performed to the feedstocks are presented in Table 4.1. FTIR, HHV,
elemental analysis were performed at University of Calgary. Density, viscosity and TAN were
carried out at University of Alberta.

Table 4.1. Characterization of HydrofactionTM Biocrude-1

Analysis Value Units Method

Density @ 20 oC 1055.1 kg/m3 ASTM D5002


Viscosity @ 40 oC 3975 cP ASTM D445
TAN 63.2 mg KOH/g oil ASTM D664-11a
HHVA 37.6 MJ/kg ASTM D240
Elemental Analysis
CA 79.4 wt%
A
H 9.1 wt%
B
ASTM D5291
O 10.9 wt%
H/C 1.37 mol
A
On dry basis. B Oxygen by difference.

4.3. Upgrading of Biocrude #1 in the Presence of CoMo Catalyst

4.3.1. Effect of temperature on catalytic hydrotreating of HydrofactionTM biocrude-1.


The temperatures investigated in catalytic hydrotreating are typically elevated temperatures when
the goal is to upgrade highly oxygenated feedstocks and produce drop-in or biocrudes with boiling
point fractions with specifications to be used as transportation biofuels [201], [156], [202]. In order
to evaluate the effect of temperature over the catalytic hydrotreating, three reaction temperatures
were screened 320, 350, and 370 °C; the upper and lower limits were studied in steps 1 and 3 of
the experimental protocol presented in chapter 3 Section 3.4.1. The level of upgrading at different
temperatures was investigated tracking the improvement of physio-chemical properties of the
biocrude-1. In Table 4.2, the results of density, viscosity, total acid number (TAN), elemental

81
analysis and other properties are tabulated. For comparison purposes, the properties of
HydrofactionTM biocrude-1 are included as well.

First, the viscosity of the biocrude was highly improved with reductions of 16.5% at 320 °C,
91.47% at 350 °C, and 98% at 370 °C taking as a baseline the viscosity of the biocrude-1.
Furthermore, through the catalytic hydrotreating at 370 °C, the TAN of the biocrude-1 was reduced
from 63.2 KOH/g oil in the feedstock to 6.1 mg KOH/g oil at 350 °C and 2.2 mgKOH/g oil at 370
°C.
Table 4.2. Effect of temperature on the upgrading of Biocrude-1. Fixed conditions: P=95 bar,
WHSV=0.3 h-1, H2/oil=900 Nm3/m3oil

Analysis Biocrude-1 T = 370 °C T = 350 °C T = 320 °C


Density @ 20 oC (kg/m3) 1055.1 1022.8 991 965.9
TAN (mgKOH/g) 63.2 6.1 8.2 13.3
HHVA (MJ/kg) 37.6 41.4 40.6 39.4
CA (wt%) 79.4 86.1 85.3 84.2
HA (wt%) 9.1 11 10.8 10.4
OB (wt%) 10.9 2.2 3.2 4.7
H/C (molar ratio) 1.37 1.53 1.52 1.48
Water Yield (wt%) - 8.9 7.3 5.3
A
On dry basis. B Oxygen by difference.

Total acid number reduction was consistent with elemental analysis; for example, the initial
oxygen content was 10.9% in the biocrude-1, and it decreased to 2.2 wt% in the upgraded biocrude
at 370 °C. It demonstrates the positive effect of temperature over the deoxygenation reactions
taking place in the process. In addition, the water produced also confirms the progress of hydro-
deoxygenation (HDO) in the feedstock; the yield of water increased with the temperature from
5.3% at 320 °C to 8.9% at 370 °C as shown in Table 4.2. The hydrogen content in the upgraded
biocrude-1 was improved at higher reaction temperatures; the H/C ratio increased from 1.37 in the
biocrude-1 to 1.48 at 320 °C and 1.53 at 370 °C.
On the other hand, the distribution of gas compounds in the exit gas after performing catalytic
hydrotreating at different reaction temperatures is presented in Figure 4.1. The analysis of these

82
results gives good signs about the level of HDO of the biocrude-1 after hydrotreating; the HDO
reactions remove the oxygen in the form of water. First, the hydrogen content (green line in Figure
4.1) remaining in the gas samples decreased at higher reaction temperatures; at 320 °C, it was
97.8%, decreased to 97.4% at 350 °C, and 96.5%, at 370 °C. These results are consistent with the
level of hydro-deoxygenation (HDO), which was increased in the upgraded samples at higher
temperatures. Comparing the oxygen content of the biocrude-1 with the oxygen content remaining
in upgraded biocrude samples and measuring the amount of water produced during the
hydrotreating process, an estimation of the contribution of the HDO reaction over the total oxygen
removed can be evaluated. Then, it can be concluded that at 370 °C, more than 90% of the oxygen
present in the biocrude-1 (oxygen in the feedstock calculated with elemental analysis) was
removed via HDO reactions in the form of water. The rest of the oxygen remaining in the biocrude-
1 was removed through decarboxylation and decarbonylation reactions. The amount of CO 2
produced (around 0.3 %) remains unchanged at the three temperatures studied (see Figure 4.1).

CH4 C2H6 C3 C4 C5 CO2 H2


1.2 98.0
97.8

1.0 97.5
Product (mole %)

97.4

H2 mol %
0.8 97.0

0.6 96.5 96.5

0.4 96.0

0.2 95.5
320 350 370
Temperature (oC)
Figure 4.1. Distribution of gas compounds at different reaction temperatures. Fixed conditions
P=95 bar, H2/oil =900 Nm3/m3oil, and HWSV= 0.3 h-1

It is necessary to clarify that some percentage of CO2 and water produced in the process (presented
in Table 4.2 and Figure 4.1), may be produced by side reactions such as water gas shift and

83
methanation of CO and CO2 taking place during the catalytic hydrotreating. Andreev et al. [203]
reported a total CO conversion in WGS reactions in a catalytic process with pre-sulfided NiMo at
400 °C.
In overall, the hydro-deoxygenation is progressed in agreement with the water yield production or
oxygen reduction in the upgraded biocrude-1 at higher temperatures. In the gas analysis figure 4.1,
it can be observed that light alkanes production (C2H6, C3, C4, and C5) increased with the
temperature.
In Figure 4.2, the thermo-gravimetric analysis performed in a nitrogen atmosphere to the upgraded
samples at different temperatures is presented. These thermal analyses performed, provide
information about the thermal decomposition (vaporization) of the components of biocrudes as the
temperature is increased. The decomposition is tracked by measuring weight loss.

120 0.6

Feed Feed
100 0.5
320°C 320°C
Deriv. Weight (%/°C)

80 350°C 0.4
Weight Loss(%)

350°C
370°C 370°C
60 0.3

40 0.2
9.63%
20 8.21% 0.1
5.98%
3.15%
0 0
0 200 400 600 800 1000 1200 0 200 400 600 800 1000
Temperature (°C) Temperature (°C)

Figure 4.2. TGA Analysis of Hydrofaction biocrude at different temperatures. Left hand-side:
weight loss. Right hand-side: derivative weight loss. Fixed conditions: P=95 bar, H2/oil =900
Nm3/m3oil, and HWSV= 0.3 h-1

On the left-hand side of Figure 4.2., the residue fraction form TGA analysis (not the same
mentioned ahead for SimDist analysis) found in the feedstock without upgrading was 9.63%. The
residue is the part of the sample which did not decompose or vaporize after 1000 °C. Then, in the
same plot, it can be observed that after catalytic hydrotreating at 370 °C, the residue content was
reduced to 3.15%. In the right-hand side figure, the differential thermal gravimetric results are
presented. These results are the analysis of the first derivative of the mass loss curves on the left-
hand side; derivative loss analysis is useful when multiple overlapping reactions occur. For

84
example, when samples have very similar weight loss over time, it is not possible to differentiate
the curves, however, with the first derivative of the weight loss curve or the inflection point; the
most significant changes in loss weight curve will be represented by a high peak in the derivative
plot, which is easy to differentiate from the others. In the figure to the right, all the samples showed
two representative peaks. The peaks on the left side are located in the temperature range between
180 to 270 °C, and the peaks in the right side are located between approximately 420 to 430 °C.
The higher peaks of upgraded samples are progressively shifted to the left to lower temperatures
range. The reason is that the lighter materials of upgraded samples are thermally decomposed at
lower temperatures, as opposed to the residue fraction, which decomposes at high temperatures.
The green line, which is the upgraded biocrude-1 at 370 °C, shows a higher peak at around 180 °C
suggesting the presence of carbonaceous fractions, which volatize easier at lower temperatures
compared with the peak of the feedstock and the other two upgraded samples. In the range of 420
°C (right-hand side figure), the same upgraded sample at 370 °C contains less carbonaceous
products which start cracking around 400 °C compared with the feedstock, which contains a high
amount of coke or low volatility compounds (higher peak).

Figure 4.3. Boiling point distribution (SimDist) results of upgraded samples at different
temperatures. Fixed conditions P=95 bar, H2/oil =900 Nm3/m3oil, and HWSV= 0.3 h-1

85
In figure 4.3, the simulated distillation (SimDist) analysis of the biocrude-1 and upgraded samples
are shown. The analysis of boiling point distribution is an essential parameter for characterizing
biocrudes and its hydrotreated products. For example, in Figure 4.3, the progress of upgrading can
be observed when the range of gasoline, diesel and LVGO products is increased and the amount
of residue decrease; in this case, it occurs at higher reaction temperatures like 370 °C. This
performance shown in the curves of upgraded samples is in agreement with the TGA results
presented in Figure 4.2. The upgraded sample at 370 °C (green line) distillate almost 60% of value-
added products in the boiling range of gasoline and diesel at temperatures below 350 °C in the
SimDist analysis. Also, around 28 % of upgraded products which distillate in the range of LVGO
and HVGO, and the vacuum residue fraction (+550 °C) reduced to around 12% contrary in the
feedstock biocrude-1, where the amount of residue fraction was around 35%. Through catalytic
hydrotreating at 370 °C, the fractions which distilled into gasoline (10%) and diesel (49%) were
found to be doubled compared to the same fractions distilled from the feedstock with just 5% of
gasoline and 23% of diesel fractions. In overall, by increasing the reaction temperature, the amount
of value-added products with lower boiling points was improved (shifted towards gasoline and
diesel fractions) due to hydro-deoxygenation, hydrogenation, cracking, among other reactions
taking place.
Results of boiling point distribution of bio-oils produced from microalgae and wood feedstocks
reported by Biller and Roth [204] are similar to the boiling point distribution of HydrofactionTM
biocrude-1 presented in Figure 4.4. However, the boiling points distribution of the hydrotreated
biocrude-1 at 350 °C reported herein shows high-quality products compared to the boiling point
distribution of hydrotreated biocrude reported in Biller and Roth study. In both studies, the
upgrading was performed in the presence of CoMo catalysts. It is important to mention that Biller
studies were carried out in a batch reactor.

86
Figure 4.4. FTIR analysis for renewable biocrude-1 and upgraded products at different
temperatures. Fixed conditions: 95 bar, H2/oil=900 Nm3/m3oil, and HWSV= 0.3 h-1

In Figure 4.4, the spectrums of spectroscopic analyses performed on biocrude-1 and upgraded
samples at different temperatures are presented for comparison. The infrared analysis is useful to
track the effect of catalytic hydrotreating over the reduction of oxygen-containing and aromatic
functional groups present in hydrotreated biocrudes. On the left-hand side of the figure, the
spectrums in the range of 3200 to 3600 cm-1 represent the O-H stretch of the feedstock (red line),
which represents the high content of oxygenated alcohols groups and phenols. After hydrotreating,
these oxygenated groups were highly reduced. It can be observed if we compared the spectrum of
the upgraded biocrude at 370 °C (green line) with the spectrum of the feedstock.
In the spectrum around 3000 to 3100 cm-1, reduction of aromatic C-H stretch groups is observed,
one more time, the upgraded sample at 370 °C shows the higher reduction compared with the
spectrum of the feedstock. The peak found in the C-C stretch peak between 1700-1500 cm-1 is
another sign of high aromaticity of the feedstock. Though, high reduction of these groups was
achieved by increasing the reaction temperature, furthermore, in Figure 4.4 the alkyl C-H stretch
groups are observed in the range of 2950-2850 cm-1 [205]; the C=O stretch peak shown around
1710 cm-1 referring to carboxylic acids present in the feedstock as well as ether groups (R-O-R)
were observed in the spectra of biocrude-1 and upgraded samples around 1100 to 1300 cm-1 [205].
Reduction of the groups mentioned was achieved by increasing the reaction temperature. Finally,

87
upgrading of biocrude-1 was confirmed through the reduction of carboxylic acids, alcohol and
phenol groups, as well as aromatic groups. The results showed in Figure 4.4. are in agreement with
hydrogenation reactions occurring during catalytic hydrotreating that will also lead to higher H/C
ratios at higher conversions, as was shown in Table 4.2.

4.3.2. Space velocity screening over the catalytic hydrotreating of HydrofactionTM


biocrude-1.
In this study, three different space velocities were investigated during the catalytic hydrotreating
of HydrofactionTM biocrude-1. The weight hour space velocity (WHSV) is the inversion of
residence time. In other words, a low space velocity means that the feedstock will spend more time
passing through the catalytic bed. As a result, more reactions will occur due to a higher exposure
of the biocrude-1 with the surface of the catalyst. Space velocity is an important operational
parameter that relates the desired oil mass flow rate (g/h) divided by the amount of catalyst packed
(g) as can bee seen in chapter 3 Section 3.3.2. Furthermore, space velocity is also an important
parameter for developing kinetic models of chemical reactions [206].

Table 4.3. Effect of space velocity on properties of upgraded products. Fixed conditions:
T=350oC, P=95 bar and H2/oil =1300 Nm3/m3oil.

Analysis Biocrude-1 0.2 h-1 0.3 h-1 0.5 h-1

Density @ 20 oC (kg/m3) 1055.1 980.2 989.5 1001.5


Viscosity @ 40 oC (cP) 3975 173 268 480
TAN (mgKOH/g) 63.2 3.7 7 8.5
HHVA (MJ/kg) 37.6 41.1 40.4 40
CA (wt%) 79.4 85.9 85.4 84.8
HA (wt%) 9.1 11 10.9 10.7
OB (wt%) 10.9 2.4 3.1 3.8
H/C (molar ratio) 1.37 1.54 1.53 1.51
Water Yield (wt%) - 8 7.5 6.6
A
On dry basis. B Oxygen by difference.

88
In this study, the space velocity screening was performed at three levels 0.2, 0.3, and 0.5 h -1. The
level of upgrading at different space velocities was studied by tracking the improvement of physio-
chemical properties of the upgraded biocrude-1. It was found that by decreasing the space velocity
from 0.5 to 0.2 h-1, the viscosity of the upgraded biocrude decreased from 480 to 173 cP.
By comparing the viscosity of biocrude-1, with the upgraded sample at 0.2 h-1 tabulated in Table
4.3, the reduction achieved was of 95.6%. The TAN reduction accomplished was from 8.5 to 3.7
mg KOH/g oil at space velocities of 0.5 and 0.2 h-1, respectively. Other parameters were improved
as well by decreasing the space velocity, for example, the H/C ratio was increased from 1.37 in
the feedstock to 1.51 in the upgraded sample at 0.5 h-1, 1.53 at 0.3 h-1 and finally to 1.54 at 0.2 h-
1
. Some authors have also investigated the effect of space velocity, for instance, Churin [207]
studied the effect of space velocity on oxygen removal using different fast pyrolysis bio-oils. Two
different feedstocks were used to produced bio-oils, wood and olive husks; the last one rich in
lignin. He concludes that a total oxygen removal in those pyrolysis bio-oils can be achieved at very
low space velocities around 0.1 h-1.

80

70 66.2 65.6 64.9


0.2 h-1
60
*Selectivity %

0.3 h-1
50
0.5 h-1
40

30 26.4
24.4 23.3

20
10.1 11.8
10 7.4

0
CO2 CH4 C2-C5

Figure 4.5. Selectivity of gas products at different space velocities. Fixed Conditions: T=350oC,
P=95 bar and H2/oil =1300 Nm3/m3oil.

𝑐𝑜𝑛𝑐𝑒𝑛𝑡𝑟𝑎𝑡𝑖𝑜𝑛 (𝑚𝑜𝑙) 𝑜𝑓 𝑥 𝑖𝑛 𝑔𝑎𝑠


∗ 𝑠𝑒𝑙𝑒𝑐𝑡𝑖𝑣𝑖𝑡𝑦 𝑜𝑓 𝑥 = 100 × 𝑐𝑜𝑛𝑐𝑒𝑛𝑡𝑟𝑎𝑡𝑖𝑜𝑛 (𝑚𝑜𝑙) 𝑜𝑓 𝑎𝑙𝑙 𝑔𝑎𝑠 𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠 𝑒𝑥𝑐𝑙𝑢𝑑𝑖𝑛𝑔 𝐻 Equation 4.1
2

89
In figure 4.5, the results of the selectivity of different gases produced during the upgrading process
are presented. First, the CO2 selectivity increased from 7.4 to 11.8% at higher space velocities. In
other words, CO2 production decreased when the space velocity decreased (at higher residence
times). The decarboxylation reactions do not progress at higher exposure times as the other
reactions that produces methane and light hydrocarbons. Furthermore, the production of light
compounds C2-C5 was favoured when the residence time increased or at low space velocity.

The calculation of H2 consumption for hydrogenation shown in Figure 4.6 was determined using
the hydrocarbon yield, the initial amount of H2 provided, and the elemental analysis results. In
Figure 4.6, it can be observed that total hydrogen consumption increased from 7 mg/g of biocrude
at 0.5 h-1 to 18 mg/g of biocrude at 0.2 h-1. Regarding hydrogen consumption in catalytic
hydrotreating processes, Albrecht et al. [158] reported hydrogen consumptions between 26 to 51
mg H2/g bio-oil after the catalytic hydrotreating of microalgae biofuel in the presence of CoMo
catalysts. In a similar investigation, Elliot et al. [208] reported results of hydrogen consumption
after upgrading of several aquatic biofuels and the typical values reported were in the range of 25
to 42 mg H2/g bio-oil. The results of different authors are mentioned to explain that the hydrogen
consumption depends on the type of process (one or two catalytic stages) used for upgrading, the
feedstock to be upgraded, and the operational conditions.

0.028 Total H2 comsumption 99.0


H2 comsumption for hydrogenation
0.024 H2 percentage in exit gas
H2 Consumption (g/g oil)

98.6
H2 in exit gas (mol %)

0.020
0.018
0.016 98.2
0.014

0.012 97.8
0.008 0.007
97.4
0.004

0.000 97.0
0.1 0.2 0.3 -1
0.4 0.5 0.6
WHSV (h )

Figure 4.6. Hydrogen consumption at different space velocities. Fixed Conditions: T=350oC,
P=95 bar and H2/oil =1300 Nm3/m3oil.

90
Typical values reported of hydrogen consumption of renewable crude oils during catalytic
hydrotreating may vary between 9 to 50 mg H2/g bio-oil. In the current study, the hydrogen
consumption results are lower than those reported for hydrotreating of HTL bio-oils from wood or
algae feedstocks [148].

4.3.3. Pressure and H2/Oil ratio screening on catalytic hydrotreating of HydrofactionTM


biocrude-1.
The last two parameters studied in the process of catalytic hydrotreating of renewable biocrudes
were the effect of hydrogen pressure and H2/Oil ratio. The objective of adjusting the H2/Oil ratio
was to study the level of improvement or conversion of the upgraded products as more hydrogen
was added to the hydrotreatment process. The improvements are related to the percentage of
oxygen removed from the feedstock, the viscosity reduction, the total acid number (TAN)
reduction, among other properties of the biocrude-1. In catalytic hydrotreating, the hydrogen must
always be present in high concentrations, and it should not be a limiting reactant in the process.
The pressure also plays an essential role in catalytic hydrotreating. The reason is because three
phases are in contact inside the reactor; the biocrude (liquid phase) which covers the catalyst
surface (solid phase), and the hydrogen molecules (gas phase) that must be dissolved in the liquid
phase to reach the active sites of the catalyst allowing all the reactions involving hydrogen to occur.
To date, there have been no reports about the solubility of hydrogen in similar HTL biocrudes. It
suggests a further investigation about the influence of the hydrogen solubility over the upgrading
of biocrudes. The hydrogen pressure of the system is adjusted with the back-pressure regulator,
and the amount of hydrogen available per gram of oil in the system is adjusted through the desired
H2/oil ratio.
To study the influence of pressure over the upgrading of biocrude, a set of experiments were
carried out at hydrogen pressures of 80 and 95 bars (both gauge pressures). In table 4.4, the analysis
performed are tabulated.

91
Table 4.4. Effect of pressure over the upgrading of HydrofactionTM biocrude-1. Fixed
parameters: WHSV=0.3 h-1, H2/oil = 900 Nm3/m3oil.

Biocrude-1 T=350 °C T=370 °C


Analysis
P=80 bar P=95 bar P=80 bar P=95 bar
Density @ 20 oC (Kg/m3) 1055.1 1004.8 991 968.8 965.9
Viscosity @ 40 oC (cP) 3975 375 339 90 78
TAN (mgKOH/g) 63.2 9.2 8.2 7.8 6.1
HHVA (MJ/Kg) 37.6 40.5 40.6 41.3 41.4
CA (wt%) 79.4 84.9 85.3 85.9 86.1
HA (wt%) 9.1 10.5 10.8 11 11
OB (wt%) 10.9 3.8 3.2 2.5 2.2
H/C (molar ratio) 1.37 1.49 1.52 1.53 1.53
Water Yield (wt%) - 7.2 7.3 8.6 8.9
A
On dry basis. B Oxygen by difference.

Some improvements in the physio-chemical characteristics of the biocrude-1 were confirmed when
the operational hydrogen pressure was increased from 80 to 95 bar at a fixed temperature of 350
°C. For example, the TAN was improved from 9.2 mg KOH/gr oil at 80 bars to 8.2 mg KOH/gr
oil at 95 bars. Also, the viscosity of the upgraded biocrude-1 was improved from 375 cP to 339 cP
at the same conditions. Oxygen content, HHV, and H/C ratio were also slightly improved by
increasing the pressure at a fixed temperature of 350 °C. At 370 °C, the effect of increasing the
hydrogen pressure did not show any considerable improvements. The physio-chemical
characteristics of the biocrude-1 were just slightly improved when the pressure was increased.
In table 4.5, the results of experiments performed at two different H2/oil ratios are presented. In a
similar way to the results mentioned above, the effect of two levels of H 2/Oil ratio was studied at
two fixed temperatures, 350 and 370 °C. The properties presented in the table, demonstrate a
negligible effect on the physio-chemical characteristics of upgraded samples when the H 2/oil ratio
was increased from 900 to 1300 Nm3/m3oil. Results of TAN, HHV, and elemental analysis did not
change more than 1 % comparing the results at 900 and 1300 Nm3/m3oil in the scenarios at fixed
temperatures of 350 °C and 370 °C.

92
Table 4.5. Effect of H2/oil ratio over the upgrading of HydrofactionTM biocrude-1. Fixed
parameters: WHSV=0.3 h-1, H2/oil = 900 Nm3/m3oil.

Analysis Biocrude-1 T=350 °C T=370 °C

H2/oil Nm3/m3 oil 900 1300 900 1300


Density @ 20 oC (Kg/m3) 1055.1 991 989.5 965.9 964.5
Viscosity @ 40 oC (cP) 3975 339 268 78 67
HHVA (MJ/Kg) 37.6 40.6 40.4 41.4 41.5
CA (wt%) 79.4 85.3 85.4 86.1 86.3
HA (wt%) 9.1 10.8 10.9 11.0 11.2
OB (wt%) 10.9 3.2 3.1 2.2 1.9
H/C (molar ratio) 1.37 1.52 1.53 1.53 1.56
Water Yield (wt%) - 7.3 7.5 8.9 9.2
A
On dry basis. B Oxygen by difference.
The results presented in table 4.4 and 4.5 indicate that the effect of increasing the operating
pressure and the H2/oil ratio has a lower impact over the upgrading of biocrudes compared to the
results found after increasing the temperature and the space velocity.

4.3.4. Study of catalyst activity.

The catalysts activity was studied by tracking some properties such as density, viscosity, total acid
number, HHV and elemental analysis of the upgraded samples at different operational times. The
comparison of the results obtained was done under the same upgrading conditions; in other words,
at the same operational pressure, temperature, space velocity an H 2/oil ratio. In Table 4.6, a
comparison of results obtained from mass balances taken at different days of operation under the
same operational conditions is shown. The samples used to track the activity of the catalyst were
collected during the experiment # 3 during the upgrading of biocrude-1 in the presence of pre-
sulfided CoMo catalysts. It can be observed in the table that all the properties remain almost in the
same range after 260 hours of operation.

93
Table 4.6. The comparison of different mass balances taken at different times during experiment
# 3 of the experimental protocol. Fixed conditions: T= 350 oC, P= 95 bar, WHSV= 0.3 h-1 and
H2/Oil= 900 Nm3 H2/m3 oil.

Analysis Mass balance


Experiment duration (h) 100 170 260
Density @ 20 oC (kg/m3) 991.2 990.8 991
Viscosity @ 40 oC (cP) 339 332 336
TAN (mgKOH/g) 8.2 8.2 8.2
HHVA (MJ/kg) 40.6 40.6 40.6
OB (wt%) 3.2 3.2 3.2
Water Yield (wt%) 7.3 7.3 7.3
A
On dry basis. B Oxygen by difference.

A similar comparison of properties was performed with the upgraded samples with CoMo catalysts
obtained during steps five and seven as can be observed in the experimental protocol tabulated int
Chapter 3 Section 3.4.1. The goal of this comparison was not only to check the activity of the
catalyst but also to verify the reproducibility of results. Experiment duration showed in Table 4.6
means the time that the hydrotreater unit was operated; in this case, the time on stream includes
the performing of different experiments, the stabilization time and the time of the mass balances
taken at steady-state conditions. A mass balance time was performed after reaching the
stabilization time, it is the time counted when the bench hydrotreater was operated at the same
operational conditions at steady state, usually during 6 to 12 hours which is the time required to
collect enough amount of sample for further analysis details about mass balance were presented in
chapter 3 Section 3.5. In the time on stream was also explained.

94
Table 4.7. Reproducibility results obtained from step 5 and step 7 of upgrading in the presence
of CoMo catalyst. Fixed conditions: T= 350 oC, P= 80 bar, WHSV=0.3 h-1 and H2/oil= 900 Nm3
H2/m3 oil.

Analysis Biocrude-1 Step 5 Step 7*


Density @ 20 oC (kg/m3) 1055.1 1004.8 1004.2
Viscosity @ 40 oC (cP) 3975 375 371
TAN (mgKOH/g) 63.2 9.2 99
HHVA (MJ/kg) 37.6 40.51 40.58
40.58
OB (wt%) 10.9 3.8 3.7
40.58
55555555444444
Water Yield (wt%) - 7.2 7.2
A
3.7
On dry basis. B Oxygen by difference.
* The slight difference between the properties of biocrude obtained at different steps could be the product of
evaporation of light products during the sampling step. 7.2

In Table 4.7, the results one more time showed the same catalyst activity after more than three
days of the catalytic upgrading process. From the results shown in Table 4.7, the reproducibility
of the results was confirmed.

4.4. Upgrading of Partially Upgraded Biocrude in the Presence of NiMo Catalyst

In campaign #1 the first stage was performed by screening several conditions in a packed bed
downflow reactor in the presence of CoMo catalyst presented in the experimental protocol
included in Section 3.4.1 of chapter 3. At step two of the first hydrotreating stage (experimental
protocol Section 3.4.1), the conditions were kept at 95 bars of pressure, space velocity of 0.3 h -1,
and H2/oil ratio of 900 Nm3/m3 oil, as can be observed in chapter 3 Figure 3.7 and experimental
protocol Table 3.4. At the fixed conditions mentioned, several mass balances were taken. As a
result, more than 2 kg of products were collected in this experiment. After removing the water
produced from the light products of each mass balance, the heavy and light products (water-free)
were well mixed manually using a glass stick, avoiding any loss of sample. A representative
sample named (1st upgrading sample) was taken and a full characterization was performed. The
water produced by HDO reactions can be an issue since it may promote catalyst deactivation. Some
authors reported studies about the deactivation mechanism of sulfided catalysts on the presence of
water [176], [205] in catalytic processes. For those reasons, the water produced in the first

95
upgrading stage was removed from the two kilograms sample. It was done to avoid a possible
catalysts deactivation in the subsequent hydrotreating stage in the presence of pre-sulfided NiMo
catalyst.
The characterization results of the 1st upgrading sample is tabulated in Table 4.8. The results
suggest that a deep upgrading was not achieved at the conditions established with one
hydrotreating stage in the presence of CoMo catalyst. Particularly the TAN (8.2 mg KOH/g oil)
and the oxygen content (3.2 wt%) of the upgraded sample are still relatively high. For the reasons
mentioned, the 2 kg sample collected was considered as a partially upgraded feedstock.

Table 4.8. Properties of biocrude-1 and upgraded samples after first and second upgrading steps.
Fixed conditions: WHSV = 0.3 h-1, P = 95 bar and H2/oil = 900 Nm3/m3 oil

1st upgrading 2nd upgrading


Analysis Biocrude-1
(CoMo) (NiMo)
Reaction temperature (°C) - 350 320 350
Density @ 40 oC (kg/m3) 1043.3 977.6 950.6 923.7
Viscosity @ 40 oC (cP) 3975 339 23 9
TAN (mgKOH/g) 63.2 8.2 3 1.4
HHVA (MJ/kg) 37.6 40.6 42.4 42.7
OB (wt%) 10.9 3.2 2.3 1.8
H/C (molar ratio) 1.37 1.52 1.6 1.62
H2 Consumption (gH2/g oil) - 0.015 0.003 0.008
H2 Consumption 1st + 2nd
- - 0.018 0.023
upgrading stage
A
On dry basis. B Oxygen by difference.

Therefore, the partially upgraded biocrude was first spiked with dimethyl-sulphide to 200 ppm. It
was used as a feedstock in the second hydrotreating stage in the presence of pre-sulfided NiMo
catalyst. In the second hydrotreating stage, two temperatures were investigated; it was found that
the physio-chemical properties of biocrude-1 were highly improved. For example, the TAN
decreased from 8.2 in the 1st upgrading stage (with CoMo catalyst) to 2.3 and 1.4 mg KOH/g oil
at 320 and 350 °C respectively in the second hydrotreating stage. The viscosity was highly

96
improved as well; it decreased from 339 cP in the 1st upgrading stage to 23 cP at 320 °C and 9 cP
in the 2nd stage at 350 °C.
Moreover, the final oxygen content found in the upgraded sample after 2nd stage hydrotreating was
found to be 1.8 wt%. In terms of hydrogen consumption, the results showed that higher
consumption occurs in the 1st stage. For example, at first hydrotreatment stage, the hydrogen
consumed was 0.015 g H2/g oil and after performing a second upgrading stage the total H2
consumption increases to 0.018 g H2/g oil when the reaction temperature was 320 °C in the 2nd
stage. Also, at 350 °C, the hydrogen consumption was 0.008 g H 2/g oil, accounting a total H2
consumption after the first and second stage of 0.023 g H2/g oil. The hydrogen consumed in the
second upgrading stage was mainly in HDO reactions. An oxygen reduction was observed from
10.9 to 3.2 wt% after performing the first upgrading stage. After performing these two
hydrotreating stages, very small amounts of coke were produced, and there was no reactor
plugging observed in the experiments performed.

Figure 4.7. Boiling point distribution (Simdis) of biocrude-1 and upgraded samples after two
upgrading stages. Fixed conditions: P=95 bar, H2/oil =900 Nm3/m3oil, and HWSV= 0.3 h-1

The boiling point distributions results shown in Figure 4.7, demonstrate considerable
improvements in the quality of upgraded products. For example, the gasoline and diesel fractions

97
found in the sample produced after the second upgrading stage at 350 °C were 12 and 53%
respectively. The 1st upgrading stage sample yield just 6 % of gasoline fractions and 42% of diesel
fractions. Furthermore, the amount of residue with a boiling point above 550 °C found in the 1 st
upgrading stage sample was around 20%. On the other hand, the sample submitted to a second
upgrading stage yield just a 5% of residue.

Finally, the results are shown in Table 4.8 and Figure 4.7, demonstrate that the two-stage upgrading
process can be an alternative which can lead to the production of improved biocrudes. The
conditions screened in association with the catalyst studied played a vital role and gave very good
signs to perform further catalytic hydrotreating experiments. It is important to mention that the
NiMo catalyst used in the second upgrading stage was re-activated and no information about that
was provided. Then, it is necessary to perform a further experimental campaign with fresh and
activated catalysts to improve even more the physical-chemical properties of HydrofactionTM
renewable biocrude.

4.5. Thermal Experiments Performed Without Catalysts

The effect of the catalysts used in the hydrotreating processes must be evaluated. One useful way
to determine the real effect over the upgrading is by performing thermal studies without the
presence of catalysts. These experiments were performed following the same procedure and same
conditions than those used during the catalytic hydrotreating in the presence of CoMo catalysts
with one stage upgrading scenario (see Section 4.3); first, the reactor was packed with inert silicon
carbide free of metals and impurities. Then the reactor filled up with SiC was used to run
experiments at four different reaction temperatures; 320, 350, 370, and 400°C, H2/oil ratio of 900
Nm3/m3, HWSV= of 0.3 h-1, and pressure of 95 bars. The reactor packed with SiC intends to
simulate the same diluted bed packed for the catalytic experiments, emulating the flow pattern of
the biocrude inside the packed reactor.

98
Table 4.9. Temperature effect over the upgrading of biocrude-1. Analysis performed without
catalysts. Fixed conditions: P=95 bars, H2/Oil =900 Nm3/m3, and HWVS= 0.3 h-1

Analysis Biocrude-1 T=320 °C T=350 °C T=370 °C T=400 °C

Density @ 40 oC (kg/m3) 1043.3 1046.91 1046.88 1039.99 1023.02

TAN (mgKOH/g) 63.2 59.2 53.8 39.74 22

HHVA (MJ/kg) 37.6 38.14 38.43 38.93 39.63

CA (wt%) 79.4 82.3 83.5 83.6 85.8

HA (wt%) 9.1 9.3 9.3 9.3 9.2

OB (wt%) 10.9 8.1 6.5 6.9 4.7

H/C (molar ratio) 1.37 1.35 1.34 1.33 1.29

A
On dry basis. B Oxygen by difference.

In Table 4.9, improvements in the physio-chemical characteristics of the biocrude-1 were


achieved. For example, the total acid number was gradually reduced when the process temperature
was increased. For example, there was a reduction of 37% or 39.74 mgKOH/g at 370 °C, and 65%
or 22 mg KOH/g oil at 400 °C, compared with the TAN of the biocrude-1. The HHV was improved
as well when the temperature in the process increased, in the experiment at 370 °C was 38.93
MJ/Kg at 370 °C; and at 400 °C, the HHV was increased to 39.63 MJ/Kg. The oxygen content of
the biocrude was progressively reduced from 10.9 in the feedstock to 4.7 wt% in the sample
subjected to thermal experiment at 400 °C. The improvements mentioned are the result of cracking
reactions taking place at high temperatures. The H/C ratio of the samples subjected to thermal
experiments was decreased with the temperature. Even though the thermal experiments were
performed at an H2/oil ratio of 900 Nm3/m3 oil, no significant signs of hydrogenation were seen
after the experiments. By performing hydrotreating experiments at temperatures between 350 to
400 °C and pressures between 80 and 90 bars in the presence of NiMo or CoMo catalysts,
improvements in the H/C ratio have been reported by some authors [209], [210],[207].
Comparing the results of thermal experiments (Table 4.9) with the results of upgraded samples
with pre-sulfided and reactivated CoMo catalysts; even at the lower temperature 320 °C screened

99
in the presence of CoMo catalysts (see Table 4.3). It is noticeable that all the physio-chemical
properties such as density, total acid number, H/C ratio and elemental analysis were highly
improved compared with those obtained after the thermal experiments at the lower temperature
320 °C and at the higher temperature of 400 °C.
It is important to mention that the viscosity of thermally upgraded samples is not presented because
the viscosity after the thermal process was highly increased; it was not possible to perform the
viscosity analysis with the equipment available in our laboratory. The samples studied in the
process without catalysts, experience a physical degradation or polymerization at higher
temperatures, for example, in the case of the sample produced in the experiment at 400 °C. To
perform the density analysis, it was necessary to heat the U-tube sensor to 60 °C, and also heat the
sample up to 70 °C to decrease the viscosity and be able to take the sample with a syringe. Then,
after injecting the sample to the densimeter and measure at 60 °C, during the cleaning step, it was
extremely difficult to remove the sample from the equipment. After trying to remove all the sample
with different solvents, the U-tube made with borosilicate glass was broken and had to be replaced.
Finally, after performing the thermal experiments, the upgrading of the sample was achieved. The
amount of water produced during the thermal process was negligible compared with the yield of
water during catalytic hydrotreating. For example, during the thermal experiment at 370 °C, the
yield of water was approximately 1 wt%, and around 3 wt% at 400 °C. Possibly, the oxygen
removed from the sample was mainly by de-hydration reactions due to the thermal effect.

100
Chapter 5: Two-Stage Catalytic Hydrotreating of HydrofactionTM Biocrude-2 in the
presence of commercial NiMo Catalysts (campaign #2)

In this chapter, the upgrading results of HydrofactionTM renewable biocrude-2 provided by Stepper
energy are presented. The biocrude-2 was produced from woody feedstocks. The catalysts used in
this campaign were three different commercials NiMo catalysts named C, D, and E. The catalysts
were activated in the hydrotreater pilot plant following a protocol provided by the sponsor Steeper
Energy; some details about the activation step are included in chapter 3 Section 3.3.3. The catalytic
hydrotreating experiments were carried out in the hydrotreater unit using two downflow reactors
in series and following an experimental protocol described in Chapter 3 Section 3.4.2. The
characterization analysis of the feedstock biocrude-2 are included in this chapter.
The results of the operational conditions screening such as temperature, pressure, space velocity
and H2/oil ratio over the upgrading of biocrude-2 investigated, are included in this chapter. In this
campaign, the water produced as a result of HDO reactions was separated at the end of each
experiment when the biocrude-2 was upgraded passing through both reactors. Finally, an
evaluation of the deactivation of the catalyst is presented.

5.1. Characterization of HydrofactionTM biocrude-2


The results tabulated below in Table 5.1. were performed at the University of Alberta and
University of Calgary. The physio-chemical characteristics of biocrude-2 were slightly lower than
those found in biocrude-1. The feedstock used to produce both biocrudes was woody biomasses.

101
Table 5.1. Characterization of HydrofactionTM Biocrude-2

Analysis Value Units Method

Density @ 20 oC 1051.1 kg/m3 ASTM D5002


Viscosity @ 40 oC 1146 cP ASTM D445
TAN 62 mg KOH/g oil ASTM D664-11a
HHVA 37.6 MJ/kg ASTM D240
Elemental Analysis
CA 80.3 wt%
A
H 9.4 wt%
ASTM D5291
B
O 9.5 wt%
H/C 1.41 mol
A
On dry basis. B Oxygen by difference.

5.2. Effect of temperature in catalytic hydrotreating of HydrofactionTM biocrude-2 (two


reactors in series process)

The temperature screening was done by first keeping the reactor one at a fixed temperature of 320
°C. Meanwhile, the effect of varying the temperature in reactor two from 320 to 370 °C over the
upgrading of biocrude-2 was explored. Moreover, the effect of high severity conditions, in this
case, is related to higher reaction temperatures such as 350 °C in reactor one and 390 °C in reactor
two.
In table 5.2, the results of density, viscosity, TAN and other properties of upgraded products are
shown. It was fount that, increasing the reaction temperature the physio-chemical properties of
biocrude-2 were remarkably improved. Even at low operating temperatures of 320 °C in both
reactors, the TAN was decreased around 97% from 62 mgKOH/g to 1.9 mgKOH/g. The viscosity
was decreased by almost 99% from 1146 cP to 14 cP, and the oxygen content was highly reduced
to 1.7 wt% in the upgraded biocrude-2 at the same conditions.

102
Table 5.2. Effect of temperature in reactors 1 and 2 over the upgrading of biocrude-2. Fixed
conditions: P=90 bars, WHSV=0.3 h-1, H2/oil=900 Nm3/m3oil

T1=320 T1=320 T1=320 T1=350 T1=350


Analysis Biocrude-2
T2=320 T2=350 T2=370 T2=350 T2=390
Density @ 20 oC (kg/m3)
1051.1 929.9 926.2 920.9 923.2 918

Viscosity @ 40 oC (cP) 1146 14 11 9 9 6

TAN (mgKOH/g) 62 1.9 0.5 0.2 0.2 0.06


HHVA (MJ/kg)
37.6 43.3 43.5 43.7 43.5 43.7

CA (wt%) 80.3 86.1 86.9 87.2 86.8 87.1

HA (wt%) 9.4 11.9 11.4 12.1 12.3 12.4


OB (wt%)
9.5 1.7 0.7 0 0.4 0

H/C (molar ratio) 1.41 1.66 1.58 1.67 1.69 1.71


A
On dry basis. B Oxygen by difference.

Furthermore, when the temperature in the second reactor was increased to 370 °C, the elemental
analysis results shown total oxygen remotion, and the viscosity of the upgraded sample was
decreased to 9 cP at 370 oC. Finally, the results at high severity conditions 350 oC in reactor one
and 390 oC in reactor two a dramatic improvement in the physical-chemical properties of biocrude-
2 were achieved. For example, the total acid number TAN was decreased to 0.06 mgKOH/g oil.
In overall, the results at higher operational temperatures in both reactors 350 and 390 °C, were
slightly better than those reported at 320 oC in reactor one and 370 oC in reactor two.

The distribution of gas compounds and the percentage of hydrogen consumed after the screening
of three temperatures in reactor two and keeping fixed the temperature in reactor one at 320 oC are
shown in Figure 5.1 light hydrocarbon gases were produced in noticeable amounts at the three
temperatures studied, except the low amount of C3 produced when the temperature was at 320 oC
in both reactors. Also, methane was produced in a relatively high percentage in the three scenarios
studied; other hydrocarbons like ethane and propane could be produced during decarboxylation
reactions of carboxylic acids present in the biocrude. The production of CO 2 confirmed the
presence of the decarboxylation reaction in the process. However, as the reaction temperature

103
increased the CO2 production seems to be decreased, but recalling the selectivity of gas products
(chapter 4) the CO2 was produced during the tree temperatures studied, but it was not produced as
much as the other compounds were produced during the experiments at 350°C and 370 °C.
Possible methanation of CO2 could occur during the hydrotreating process too. Methanation of the
CO2 occurred when the CO2 produced reacts with hydrogen present in high concentrations during
the process. For example, in Figure 5.1 it can be observed that the CH 4 was increased with the
temperature and the CO2 decreased as a result of higher yield of other hydrocarbons or
methanation.
Similar results were found by Yakovlev et al. [211] in a study about the upgrading of biocrudes in
the presence of NiMo catalysts. On the other hand, an opposite effect was observed with the
selectivity to hydrocarbons, which increased as the temperature of the process increased. The
importance of selectivity to CO2 production at high upgrading temperatures in this specific
upgrading process is not relevant, as was explained in chapter 4. The level of hydro-deoxygenation
of products was improved when the temperature of reactor two was increased; the reason is that
the catalysts packed in reactor # 2, catalysts D and E are known for having very high hydrotreating
activity. In addition, the effect of temperature benefits the hydro-deoxygenation of the biocrude-
2. The results mentioned, can be correlated with the results shown in Figure 5.1 in the right-hand
side, where less amount of hydrogen present in the gas phase was observed at higher temperatures;
it is a result of high hydrogen consumption at elevated temperatures.

104
Figure 5.1. Selectivity of compounds in the gas phase at different temperature in reactor 2. Left-
hand side: produced gases, right-hand side: remaining hydrogen in the exit gases. Fixed
conditions: TR1=320 °C, P=90 bars, WHSV=0.3 h-1, H2/oil=900 Nm3/m3oil

The product yield results shown below in Figure 5.2 indicate that by increasing the temperature of
the second reactor from 320 oC to 370 oC, the production of water increased from 8 wt% at 320 oC
to 8.9 wt% at 370 oC. Moreover, Hydrogen consumption increased from 0.019 to 0.023 gH2/g oil
when the temperature in the second reactor increased from 320 oC to 370 oC. The aforementioned
water production and hydrogen consumption confirmed the progress of HDO reactions at higher
temperatures.

105
Liquid oil Water Gas H₂
100 0.4 0.7 0.7 0.024
8.0 8.2 8.9
0.0234
0.023

H2 consumption ( g H2/g oil)


Prodcut Yiled (wt%) 88

0.022

75
0.0212
91.6 90.3 0.021
91.1

63
0.0199 0.020

50 0.019
320 °C 350 °C 370 °C

Figure 5.2. Products yield and hydrogen consumption by the effect of different temperatures in
reactor 2. Fixed conditions: TR1=320 °C, P=90 bars, WHSV=0.3 h-1, H2/oil=900 Nm3/m3oil

Next, for the set of experiments at high severity conditions (350 °C in reactor one and 390 °C in
reactor two), the temperature of reactor one was increased to 350 oC to study the effect of higher
operating temperatures in both reactors. In table 5.1, the results at these conditions showed
considerable improvements in the physical-chemical properties of the upgraded feedstock.
Nevertheless, comparing the results obtained when reactor 1 was at 320 oC and reactor two at 370
o
C (step 3 of the protocol in chapter 3), it can be concluded that total oxygen removal and higher
TAN reduction were reached when the temperature was increased at 370 oC in the reactor 2.
Furthermore, slightly better results like total oxygen removal, improved H/C ratio and almost zero
TAN were obtained when the reactor 2 was a 390 oC. By increasing the temperature in reactor two,
the performance of the catalysts D and E was highly improved. It was found that the cracking of
heavy fractions is favoured when the catalysts were kept in a sulphided state [82]; in this
investigation, the feedstock was kept sulphided by adding to the feedstock 200 ppm of dimethyl
sulphide; this is a required step in the upgrading of biocrudes processes because the sulphur content
in this kind of samples is negligible. The need for sulphur addition to the feedstock was explained
in chapter 3, Section 3.3.1.
The gas analysis results presented below in figure 5.3 indicated that some fractions of light liquid
products were converted into gases at the highest reaction temperatures screened (350 oC and 390

106
o
C). The gas yield observed at 320 oC in reactor one and 390 oC in reactor two was found to be
three times higher than that produced when the temperatures in both reactors were 350 oC. These
results indicate the progress of cracking reactions with temperature increasing. The hydrogen
consumption also increased from 0.019 at 320 oC in both reactors to 0.0326 or 33 mgH2/g oil at
high severity conditions. Even though the properties of the biocrude-2 were highly improved,
reaching total oxygen reduction and TAN of zero at the most elevated temperatures screened, the
yield of liquid products slightly decreased, as can be seen in figure 5.3. Then, by performing
catalytic hydrotreating at high temperatures in both reactors, 350 oC in reactor one and 390 oC in
reactor two, and at the fixed conditions screened (pressure of 90 bars, WHSV of 0.3 h-1, and H2/oil
of 900 Nm3/m3 oil) is not necessarily the best pathway for the process since it is important to
maximize the production of liquid hydrocarbons as well. Further investigations are required to
improve the yield of value-added with improved physio-chemical characteristics in the products
and decreasing the amount of gas produced during the upgrading of Hydrofaction TM biocrudes.
Finally, it must be taken into account that the improvements achieved are significant and
approached the main objectives of the investigation presented in chapter one.

Liquid oil Water Gas H₂


100 0.4 0.7 2.1 0.035

H2 consumption ( g H2/g oil)


8.0 8.4 0.0326 9.0
0.031
Prodcut Yiled (wt%)

88
0.0287
0.027

75
91.6 91.0 0.023
88.9

63 0.0199
0.019

50 0.015
T₁=320 °C, T₂=320 °C T₁=350 °C, T₂=350 °C T₁=350 °C, T₂=390 °C

Figure 5.3. Products yield and hydrogen consumption at severity conditions, higher temperatures
in both reactors. Fixed conditions: P=90 bars, WHSV=0.3 h-1, H2/oil=900 Nm3/m3oil

107
The results presented in Figure 5.4 below, demonstrated the improvement in the boiling point
distribution of the products, especially for the upgraded products at higher temperatures in both
reactors. In this case, the progress of hydrogenation and cracking reactions took place during the
process resulted in lower boiling point distribution of the products. For example, the boiling point
cuts with light fractions below 454 °C were shifted to the left when the temperatures in the reactors
increased. The gasoline fractions of the three reaction temperature scenarios showed in figure 5.4.
were in the same range (13 to 15%). The distillate fractions in the diesel range (180 to 343 °C)
were highly improved as well, for instance the upgraded sample at 350 °C in reactor one and 390
°C in reactor two distillated almost 50% in the diesel range, similar fractions were obtained in the
other two upgrading scenarios at lower temperatures in both reactors 320 °C in R1 and 350 °C in
R2. Finally, the vacuum residue fraction found was above 35% in the biocrude-2, and it was
decreased to 7 % in the upgraded samples at 350 °C in reactor 1 and 390 °C in reactor 2. In this
set of experiments, the catalysts C, D, and E were used in the process to promote oxygen and
carboxylic acid remotion by HDO reactions (hydrotreating function) and adding hydrogen as well.
The upgrading of biocrude produced from miscanthus feedstock was performed by Castello et al.
[212]. Miscanthus is a lignocellulosic perennial grass with similar biomass composition to the
TM
feedstock used to produce Hydrofaction biocrude. As a result of the investigations
accomplished by Castello et al. in the presence of NiMo catalysts, the boiling point distribution
found in the upgrading process of miscanthus biocrude indicates similar gasoline fractions than
those shown in figure 5.4 at higher temperatures. However, the recovered diesel fractions obtained
in our study and shown in figure 5.4 were higher in all the temperature ranges investigated than
those presented in Castello studies.

108
Figure 5.4. Boiling point distribution (Simdis) results of upgraded samples at different
temperatures in reactors 1 and 2. Fixed conditions: P=90 bars, WHSV=0.3 h-1, H2/oil=900
Nm3/m3oil

5.3. Effect of Space Velocity on Catalytic Hydrotreating of HydrofactionTM Biocrude-2 (two


reactors in series process)
Another critical parameter in catalytic hydrotreating is the effect of space velocity. This parameter
determines the time that the biocrude-2 will spend passing through the catalytic beds. The results
of the screening of three different space velocities 0.2, 0.3, and 0.5 h-1 are shown below. The effect
of decreasing the space velocity (or increase the resident time of biocrude-2 in the reactor) over
the process was reflected in the improvement of the hydrotreated products. In table 5.2., the results
of two scenarios at a fixed temperature in reactor one (320 oC) and 2 different temperatures in
reactor two (350 and 370 oC) are shown. Then, the results obtained when the temperature in reactor
two was 350 oC, and by decreasing the space velocity at 0.2 h-1, the oxygen content (from elemental
analysis) was reduced to 0.1 wt%.
Moreover, the TAN of biocrude-2 was decreased to 0.2 mgKOH/g at the same conditions. Further
improvements were obtained when the reactor one was heated at 320 °C and the reactor two was
kept at 370 oC; for instance, the TAN decreased from 62 mmgKOH/g in the renewable biocrude-

109
2 to less than 0.1 mgKOH/g. In addition, the total acid number TAN achieved was very close to
zero. Besides, other physio-chemical properties were also highly improved. For example, the
viscosity decreased to 10 cP when the temperature was 350 oC in the reactor one, and as low as 8
cP at a higher temperature in reactor two (370 oC). High heating value and H/C ratio were highly
improved in both scenarios studied 350 oC and 370 oC in reactor 2 and at the lowest space velocity
tested of 0.2 h-1.

Table 5.3. Effect of space velocity on properties of upgraded products. Fixed conditions: TR1
=320 oC P=90 bar, and H2/oil = 900 Nm3/m3oil.
Analysis Density Viscosity TAN HHVA C H OA,B H/C
@ 20oC @ 40 oC (mgKOH/g) (MJ/kg) (wt %) (wt %) (wt %) (mol)
(kg/m3) (cP)
Biocrude-2 1051.1 1146 62 37.6 80.3 9.4 9.5 1.41

WHSV Temperature reactor 2 = 350 °C


(h-1)
0.5 945.2 20 3.9 42.6 86.7 11.3 1.4 1.57
0.3 926.2 11 0.5 43.5 86.9 11.4 0.7 1.58
0.2 925.2 10 0.2 43.6 87.2 12.3 0.1 1.69
WHSV Temperature reactor 2 = 370 °C
(h-1)
0.5 929.3 10 1.7 43.4 87.2 11.8 0.3 1.62
0.3 920.9 9 0.2 43.7 87.2 12.1 0 1.67

0.2 920.0 8 < 0.1 43.8 86.9 12.4 0 1.73


A
On dry basis. B Oxygen by difference.

From figure 5.5 below, it can be observed that, as the space velocity was decreased from 0.5 to 0.2
h-1, the hydrogen consumption increased at the two temperatures investigated. The hydrogen
consumption at 350 °C in reactor 2 increased from 0.019 wt% at 0.5 h -1 to 0.024 wt% at 0.2 h-1
(Figure 5.5. A). Similar behaviour was found in the experiment performed at 370 °C in reactor 2
(Figure 5.5. B), where the hydrogen consumption increased as the space velocity decreased.
Nonetheless, the highest yield of water was not obtained at a space velocity of 0.2 and 370 oC;
instead, it was obtained at 0.3 h-1 and 320 oC in reactor one and 370 oC in reactor two.

110
A Liquid oil Water Gas H₂ B Liquid oil Water Gas H₂
100 0.5 0.7 0.6 0.030 100 0.7 0.7 0.6 0.030
7.9 8.2 8.0 8.5 8.9 8.2

H2 consumption (g H2/g oil)

H2 consumption ( g H2/g oil)


90 0.028
Product Yiled (wt%)

88

Product Yiled (wt%)


0.025 0.027
80 0.024
0.026

91.1 75

70 91.6 0.021 91.4 90.8 90.4 91.2 0.024


0.020 0.023

0.019 63
60 0.022

0.021
50 0.015 50 0.020
0. 5 0.3 0.2 0.5 0.3 0.2
WHSV (h-1) WHSV (h-1)

Figure 5.5. Products yield and hydrogen consumption at different space velocities. Left hand-
side figure (A) TR2=350 °C; right-hand side (B) TR2= 370 °C. Fixed conditions: TR1= 320 C,
P=90 bars, and H2/oil=900 Nm3/m3oil

5.4. Effect of H2/oil Ratio on Catalytic Hydrotreating of HydrofactionTM Biocrude-2 (two


reactors in series process)
In the process of catalytic hydrotreatment of the HydrofactionTM renewable biocrude, the objective
of increasing the H2/oil ratio was to improve the quality of the upgraded products as more hydrogen
was added to the process; in other words, more hydrogen volume per volume of biocrude added
(Nm3 H2/m3 oil). The high heating value of upgraded products is not only improved by oxygen
removal through hydro-deoxygenation, decarboxylation, or decarbonylation reactions tacking
place in the catalytic upgrading; it is also improved by increasing the hydrogen added to the
upgraded samples through saturation reactions (hydrogenation). For these reasons, it is essential
to study how much hydrogen is required in the process in order to achieve the desired levels of
upgrading, such as total oxygen removal and improved distillation curve.
In Table 5.3, the results of upgraded products at two different H2/oil ratios are shown. It was found
that by increasing the H2/oil ratio from 900 to 1300 scc/cc, the physio-chemical characteristics of
the biocrude-2 were not improved significantly. In other words, there were improvements in the
TAN number of the feedstock, the density and viscosity were improved as well as the elemental
analysis of the upgraded samples when the H2/oil ratio in the process was 900 Nm3/m3 oil in both

111
scenarios investigated at 350 °c and 370 °C in reactor two. Nevertheless, results in the same range
were obtained when the H2/oil ratio was increased to 1300 Nm3/m3 oil. From the table 5.3, it can
be observed at a H2/oil ratio of 900 Nm3/m3, and temperature in reactor 2 of 370 °C, the viscosity
was decreased to 8 cP, the TAN was reduced to almost zero and the oxygen was totally removed
from the biocrude-2 at the conditions showed in table 5.3. No further improvements were observed
by increasing the H2/oil ratio above 900 Nm3/m3. It seems that the amount of hydrogen used for
HDO reactions, plus the amount added to increase the molecular hydrogen in the upgraded
biocrude-2 was enough at 900 Nm3 /m3. The last observation can be reinforced with the H/C ratio
results obtained at 1300 Nm3 as shown in table 5.3.

Table 5.4.Effect of H2/oil ratio on properties of upgraded products. Fixed conditions: TR1 =320
o
C P=90 bar, and WHSV= 0.2 h-1.

Analysis Density @ Viscosity TAN HHVA C H OA,B H/C


20oC o
@ 40 C (mgKOH/g) (MJ/Kg) (wt (wt (wt (mol)
(kg/m3) (cP) %) %) %)
Biocrude-2 1051.1 1146 62 37.6 80.3 9.4 9.5 1.41

H2/Oil
Temperature reactor 2 = 350 °C
(Nm3/m3 oil)
900 925.2 10 0.2 43.6 87.2 12.3 0.1 1.69
1300 923.3 10 0.2 43.6 86.7 12.2 0.1 1.69
H2/Oil
(Nm3/m3 oil)
Temperature reactor 2 = 370 °C
900 920.8 8 < 0.1 43.8 86.9 12.4 0 1.71
1300 916.6 7 < 0.1 43.9 87.1 12.6 0 1.73
A
On dry basis. B Oxygen by difference.

The yields are shown in figure 5.6. are in the same range for the two scenarios studied (left and
right-hand side results). The hydrogen consumption increased in both scenarios; at 350 °C, from
0.024 to 0.029 gH2/g oil when the H2/oil was increased to 1300 Nm3/m3 oil. However, in the
experiments performed at a higher temperature of 370 °C in reactor two, there was a slightly higher
hydrogen consumption from 0.027 at 900 Nm3/m3 oil to 0.034 gH2/g oil at 1300 Nm3/m3 (right-
hand side).

112
Liquid Oil Water Gas H2 consumption Liquid Oil Water Gas H2 consumption

100 0.6 0.9 0.035 100 0.6 1.2 0.04


8 8.8 8.2 8.9
0.03 0.034 0.035

H2 consumption ( g H2/g oil)


0.029

H2 consumption ( g H2/g oil)

Product Yiled (wt%)


Product Yiled (wt%)

88 88 0.03
0.024
0.025
0.027
0.025
T=350oC 0.02 T=370oC
75 75 0.02
91.4 90.3 0.015 91.2 89.9 0.015
0.01
63 63 0.01
0.005 0.005

50 0 50 0
900 1300 900 1300
WHSV (h-1) WHSV (h-1)

Figure 5.6. Products yield and hydrogen consumption at H2/oil ratios. Left hand-side figure
TR2=350 °C; right-hand side TR2= 370 °C. Fixed conditions: TR1= 320 C, P=90 bars, and
HWSV =0.2 h-1

5.5. Effect of Hydrogen Flow Pattern over the Catalytic Hydrotreating of HydrofactionTM
Biocrude-2 (two reactors in series process)
In figure 5.7, a representation of the study performed about the addition of hydrogen into reactor
one and two is shown. To do so, two scenarios were studied; in the first one, the hydrogen for the
process was supplied through reactor one, as shown in the left-hand-side of figure 5.7. In the
second scenario, the hydrogen was split and supplied to both reactors as shown in the right had-
side of the figure; in this experiment, fresh hydrogen was sent to reactor one and two at the same
flow rate. In both scenarios, the hydrogen mass flow was the same; the only difference is that it
was split to be fed to both reactors. The analysis performed to the properties of upgraded samples
collected from the two scenarios described does not show any important difference in the upgraded
products. In the table below, some analysis performed is shown. Additional analyses were
performed such as FTIR, and SimDist and no considerable differences were seen either. It can be
concluded that any improvement was not found by altering the way of adding the hydrogen during
the reaction.

113
Figure 5.7. Sketch of the hydrogen flow patterns investigated in the catalytic hydrotreating of
biocrude-2.

The properties of upgraded biocrude-2 after two different hydrogen addition flow patterns are
tabulated in Table 5.4. In overall, the feedstock was highly upgraded after the two upgrading
scenarios described in figure 5.7 but, by manipulating the hydrogen addition to the process (split

Table 5.5. Effect of hydrogen flow patterns over the catalytic hydrotreating of biocrude-2. Fixed
conditions: TR1=320 oC TR2=370 oC P=90 bar, HWSV=0,3 h-1 H2/Oil= 900 Nm3/m3 oil

Analysis Biocrude-2 Non-split Split


Density @ 20 oC (kg/m3) 1051.1 923.4 922.6
Viscosity @ 40 oC (cP) 1146 9 9
TAN (mgKOH/g) 62 0.8 0.7
HHVA (MJ/kg) 37.6 43.5 43.5
CA (wt%) 80.3 87.2 87.4
HA (wt%) 9.4 12.1 12.3
OB (wt%) 9.5 0.5 0.4
H/C (molar ratio) 1.41 1.67 1.69
Liquid yield (wt%) - 90.5 90.4
Water yield (wt%) - 8.8 8,9
Gas yield (wt%) - 0.7 0.7
H2 Consumption (gH2/g oil) - 0.023 0.022
A
On dry basis. B Oxygen by difference.

114
to both reactors and directly trough reactor one), no significant differences in the properties of
upgraded biocrude-2 were observed. The yield of water found in upgraded biocrude-2 samples,
produced gases and hydrogen consumed during the process was similar in both scenarios
investigated.

5.6. Evaluation of Catalyst Deactivation

The activity of the catalyst was studied in the first set of experiments, or campaign # 1 (chapter 4).
Similarly, in campaign # 2, after 200 hours of continuous operation, experiments were performed
to track the catalytic activity. The first set of experiments were performed with the upgraded
samples obtained after 24 hours of operation in the catalytic hydrotreater pilot plant. The samples
for characterization were collected after reaching the stabilization time.

Table 5.6 Evaluation of catalysts deactivation. Fixed conditions: TR1=320 oC TR2=350 oC P=90
bar, HWSV=0,3 h-1 H2/Oil= 900 Nm3/m3 oil

Analysis Biocrude-2 Sample 1 Sample 2


Experiment duration (h) - 24 200
o 3
Density @ 20 C (kg/m ) 1034.8 912.4 917.5
Viscosity @ 40 oC (cP) 1146 11 13
TAN (mgKOH/g) 62 1 1.2
HHVA (MJ/kg) 37.6 43.5 43.9
CA (wt%) 80.3 86.9 86.8
HA (wt%) 9.4 11.4 11.2
OB (wt%) 9.5 0.7 1.2
H/C (molar ratio) 1.41 1.58 1.55
A
On dry basis. B Oxygen by difference.
At the beginning of the process, the catalysts were fresh and highly active, and the results obtained
were slightly better than those obtained from the sample collected after 200 hours of continues
operation.
According to the results shown in table 5.5, slightly less catalyst activity after 8 days of continuous
operation was observed. One possible explanation can be that the catalysts lost their activity with
the processing time when the surface starts becoming covered with some coke deposits that
accumulate over time. The physio-chemical properties presented in table 5.5., are presented. Even

115
though the results of density, viscosity, TAN, and HHV were better in the 24 hours sample, the
elemental analysis results did not change significantly.

116
Chapter 6: Conclusions and Recommendations

A bench hydrotreater pilot plant with two continuous downflow packed bed reactors in series was
constructed and commissioned at the University of Alberta; the hazard assessment and standard
operational procedure were established for the secure operation of the unit. In the bench
hydrotreater pilot plant, prolonged experiments were carried out in the presence of CoMo and
NiMo commercial catalysts and substantial quantities of hydrogen. In this investigation, two
independent hydrotreating campaigns were accomplished to study the effect of operational
conditions during the upgrading of renewable HydrofactionTM biocrudes. In both campaigns, the
performance of the catalysts over time was studied. Also, the thermal effects on the upgrading of
biocrudes were studied; it was accomplished by performing experiments in continuous mode
without catalysts.
Through this investigation, the physio-chemical properties of the biocrudes subjected to catalytic
hydrotreating were highly improved. The data obtained in this study is relevant not only for
estimating the feasibility of the upgrading process. Also, the results are crucial to scale-up the
upgrading process of HydrofactionTM biocrudes. Once the upgrading of the renewable biocrudes
reaches a commercial scale, in the form of drop-in biocrudes or value-added finished fuels for
transportation; it will contribute with a significant decrease in harmful emissions.

6.1. Conclusions

• Conclusions campaign # 1:

One important conclusion of this investigation is that the upgrading performed in one single
catalytic stage with CoMo catalyst A did not result in a significant improvement of the properties
of HydrofactionTM biocrude-1 compared with the results obtained in the process performed with
two reactors in series campaign #2. After the second stage, performed with NiMo catalyst B in
campaign #1, slightly better characteristics in the upgraded samples were achieved.
Some representative results obtained in the first stage scenario in the presence of CoMo catalyst
A at fixed conditions of 370 °C, 95 bars, 0.3 h-1, and 900 Nm3/m3 oil were obtained. For example,
the total acid number was reduced to 6.1 mgKOH/g (around 90 %), the oxygen content was

117
reduced to 2.2 wt% (80%), the viscosity was highly reduced to 78 cP (99.5%), and the H/C ratio
was improved to 1.53. The baseline for comparison was the properties of the biocrude-1. Recall
that, the second scenario was accomplished in the presence of NiMo catalyst B, and the feedstock
was the partially upgraded sample obtained in the first upgrading stage scenario with CoMo
catalyst A. Then, the total acid number was reduced to 1.4 mgKOH/g oil or 98 % compared to the
TAN of the biocrude-1. The viscosity was reduced by 99.8%, from 3975 cP in the biocrude-1 to 9
cP, and the oxygen content in the sample upgraded with catalyst NiMo B at 350 °C was decreased
to1.8 wt%.

The screening of different operational conditions demonstrated that temperature had the greatest
influence on the improvement of physio-chemical properties of renewable biocrude-1. Also, the
distillation curves of the upgraded products at the higher temperature explored 370 °C, showed
that the distilled fractions in the gasoline and diesel range were increased. These results are
consistent with the greater H/C ratios, and heating values (HHV) obtained.
Space velocity was the second most important parameter in the upgrading of biocrude-1. It was
observed in the experiments performed in the presence of CoMo catalysts A, and stage two in the
presence of NiMo catalysts B. When the space velocity was decreased, all the physio-chemical
properties of the upgraded samples were improved.

The results obtained after the screening of operational pressure and H2/oil ratio demonstrated that
these parameters were less effective in terms of the upgrading of biocrude, compared to the effect
of temperature and space velocity. Negligible improvements were found when the pressure was
increased from 80 to 95 bars, and no difference in the properties of upgraded products was
observed when the H2/oil ratio was increased from 900 to 1300 Nm3/m3 oil.

Finally, the thermal experiments (without catalysts), confirmed that hydro-deoxygenation (HDO)
and hydrogenation reactions do not take place in the absence of catalysts. Because, the H/C ratio
was not improved, neither water was produced, nor hydrogen was consumed during the thermal
experiments.

118
• Conclusions Campaign # 2:

In this campaign, after performing the activation of three commercial catalysts C, D, and E, the
first parameter screened was the temperature. It was accomplished by first keeping the temperature
in reactor one at 320 °C and then varying the temperature in reactor two from 320 to 370 °C.
Further screening of temperature was carried out at high severity conditions (R1 at 350 °C and R2
at 390 °C).

The best upgrading results of the second campaign were observed after the screening of
temperature at high severity conditions of temperature, the physio-chemical properties of the
biocrude-2 were dramatically improved. For instance, the TAN was reduced as low as 0.06
mgKOH/g oil, and the viscosity was reduced to 6 cP. Besides, a total oxygen reduction in the
sample was found with elemental analysis; the HHV of the upgraded sample was highly improved
to 43.7 MJ/kg, as well as the H/C of 1.71. The results of upgrading at high severity conditions
were slightly better than those obtained after performing upgrading with reactor one at 320 °C and
reactor two at 370 °C. From the experimental results it can be concluded that, when the hydrogen
consumption increased during the catalytic hydrotreating process, the oxygen removal, water
production, and the H/C ratio in the products were improved simultaneously, which means that the
hydro-deoxygenation (HDO) and hydrogenation reaction progressed at the same time.

The distillation curve (boiling point distribution) of upgraded samples were highly improved;
specifically, the upgraded sample at high severity conditions of temperature. It was found a very
low vacuum residue in this sample of 7% compared to around 37% in the biocrude-2. In addition,
all the light fractions with a boiling point below 454 °C were increased at high severity conditions
(high temperatures) in both reactors; it was found that more than 50% of the sample distilled in
the diesel and gasoline range. From the results obtained after the temperature screening in both
reactors, it was concluded that increasing the reaction temperature in the reactor two, the properties
of the upgraded samples were further improved. It was expected because the catalysts D and E
packed in reactor two are catalysts designed for very high hydrotreating activity (at high
temperatures and pressures), contrary to the catalyst C in reactor one which has a low hydrotreating
function and was used as a guard bed of catalysts D and E.

119
After performing this second hydrotreating campaign, it was concluded that temperature and space
velocity were the most important parameters over the progress of upgrading of the biocrude-2.
Finally, from the evaluation of catalysts over time, it was observed a slightly lower catalytic
activity after eight days of continuous operation. The reason could be that the catalysts started
being covered with some coke-like formations (gums) after a prolonged operational time.
Additionally, the temperature of 390 °C screened in reactor two could trigger the formation of
undesirable compounds such as gums in the surface of the reactors.

Final remarks:
• The temperature and the space velocity were the operational conditions with stronger
influence over the upgrading of HydrofactionTM biocrudes. It was observed in the results
obtained in both catalytic campaigns performed.

• In the bench hydrotreater pilot plant, it was possible to perform extended time experiments,
and screen several operational conditions in continuous mode with two downflow reactors.
Very few amounts of coke-like solids were found after performing prolonged catalytic
hydrotreating experiments.

• The catalytic hydrotreating performed to biocrude-2 in campaign #2, showed promising


results; for instance, the initial oxygen content of biocrude-1 was totally removed, and the
total acid number of the biocrude was decreased to almost zero, through the catalytic
hydrotreating performed. It can be concluded that a two-stage hydrotreating process (two
reactors in series) was more effective than one catalytic stage process.

• The configuration of commercial catalysts C, D, and E used in campaign #2, activated and
tested in the bench hydrotreater pilot plant (campaign #2), was highly effective performing
HDO and hydrogenation reactions. The configuration of catalysts can be contemplated as a
possible option to perform further hydrotreating experiments due to the promising results
achieved.

120
• The knowledge and expertise acquired after performing the catalytic hydrotreating
experiments are important for further improvements in the process.

6.2. Recommendations

The upgrading of renewable HydrofactionTM biocrudes was successfully performed, and reliable
data was obtained at different conditions in the presence of different catalysts. However, further
experiments are required to develop a suitable kinetic study which is necessary to scale-up the
process. Also, the upgrading process of HydrofactionTM biocrude can be modelled only if kinetic
data about the process is available.
In addition to the kinetic studies, it is recommended to evaluate the stability of the upgraded
biocrudes. That information is essential in the case that the upgraded biocrudes requires storage
for extended periods, for transportation by pipelines or in the case that the upgraded biocrude is
going to be used as drop-in biocrude.
The dilution of hydrogen with the biocrude for this process is not well understood, it is
recommended for further campaigns perform some studies to acquire a better understanding about
the effect of different conditions such as temperature and pressure over this process.
The catalytic configuration used in the reactors one and two in campaign #2 was maintained fixed
during the screening of several conditions. One option to be explored could be increasing the
amount of catalyst packed modifying the proportions (SiC/catalysts). Another attractive option
could be the exploration with a different set of hydrotreating catalysts with similar catalytic bed
proportions (SiC/catalysts) than those used in this investigation.

121
REFERENCES

[1] O. Edenhofer, R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A.


Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C.
von Stechow, “Technical Summary in Climate Change 2014: Mitigation of Climate
Change.,” 2014.
[2] A. K. D. S.N. Naik, Vaibhav V. Goud, Prasant K. Rout, “Production of first and second
generation biofuels: A comprehensive review,” p. 15, 2009.
[3] D. Manfre, “Analytical and numerical modeling of the Cyclic ES-SAGD process,” Univ.
Calgary, 2019.
[4] E. Furimsky, “Catalytic hydrodeoxygenation,” Appl. Catal. A Gen., vol. 199, no. 2, pp.
147–190, 2000.
[5] P. E. Savage, “Hydrothermal Processing of Biomass for Biofuels,” in Biofuel Research
Journal, no. 1, 2010, pp. 192–221.
[6] M. Kumar, A. Olajire Oyedun, and A. Kumar, “A review on the current status of various
hydrothermal technologies on biomass feedstock,” Renew. Sustain. Energy Rev., vol. 81,
no. November 2016, pp. 1742–1770, 2018.
[7] E. P. E. A. Hawkins Ortega Suckling Schurer Hegerl, “GLOBAL TEMPERATURE
SINCE THE PREINDUSTRIAL PERIOD,” Am. Meteorol. Soc., no. September, pp.
1841–1856, 2017.
[8] C. Environment and Climate Change, “National Inventory Report 1990-2016 -
Greenhouse Gas Sources and Sinks in Canada: Executive Summary,” 2018.
[9] S. Phillips, B. Flach, S. Lieberz, and A. Rossetti, “EU Biofuels Annual 2017.” 2017.
[10] G. R. Timilsina and A. Shrestha, “How much hope should we have for biofuels?,” Energy,
vol. 36, no. 4, pp. 2055–2069, Apr. 2011.
[11] P.-C. Framework, PAN ‑ CANADIAN FRAMEWORK on Clean Growth and Climate
Change. 2018.
[12] Goverment of Canada, “CONSOLIDATION Renewable Fuels Regulations,” 2018.
[13] R. A. Jansen, Second Generation Biofuels and Biomass, no. October. 2012.
[14] U. Rashid, F. Anwar, B. R. Moser, and G. Knothe, “Moringa oleifera oil: A possible
source of biodiesel,” Bioresour. Technol., vol. 99, no. 17, pp. 8175–8179, 2008.
[15] S. F. V. Byran R. Moser, “Evaluation of alkyl esters from Camelina sativa oil as biodiesel

122
and as blend components in ultra low-sulfur diesel fuel,” Int. J. Appl. Eng. Res., vol. 10,
no. 17, pp. 37724–37730, 2015.
[16] U. Rashid, F. Anwar, A. Jamil, and H. N. Bhatti, “Jatropha curcas seed oil as a viable
source for biodiesel,” Pakistan J. Bot., vol. 42, no. 1, pp. 575–582, 2010.
[17] J. J. Cheng and G. R. Timilsina, “Status and barriers of advanced biofuel technologies: A
review,” Renew. Energy, vol. 36, pp. 3541–3549, 2011.
[18] R. E. H. Sims, W. Mabee, J. N. Saddler, and M. Taylor, “An overview of second
generation biofuel technologies,” 2009.
[19] A. Alaswad, M. Dassisti, T. Prescott, and A. G. Olabi, “Technologies and developments
of third generation biofuel production,” Renew. Sustain. Energy Rev., vol. 51, pp. 1446–
1460, 2015.
[20] F. Alam, S. Mobin, and H. Chowdhury, “Third generation biofuel from Algae,” Procedia
Eng., vol. 105, no. Icte 2014, pp. 763–768, 2015.
[21] C. U. Jensen, J. K. Rodriguez Guerrero, S. Karatzos, G. Olofsson, and S. B. Iversen,
“Fundamentals of HydrofactionTM: Renewable crude oil from woody biomass,” Biomass
Convers. Biorefinery, vol. 7, no. 4, pp. 495–509, 2017.
[22] C. U. Jensen, “Aalborg Universitet PIUS - Hydrofaction ( TM ) Platform with Integrated
Upgrading Step Jensen , Claus Uhrenholt,” 2018.
[23] T. Moraes Righi, “Evaluation of Catalytic Steam Cracking Process for Total Acid Number
Reduction of Heavy Oils,” Univ. Calgary, vol. 3, no. 1, p. 183, 2016.
[24] A. D. P. Gruia J. Jones Pujado, “Handbook of Petroleum Processing.” pp. 1–82, 2001.
[25] W. Mu, H. Ben, A. Ragauskas, and Y. Deng, “Lignin Pyrolysis Components and
Upgrading-Technology Review,” Bioenergy Res., vol. 6, no. 4, pp. 1183–1204, 2013.
[26] T. J. Benson, P. R. Daggolu, R. A. Hernandez, S. Liu, and M. G. White, Catalytic
deoxygenation chemistry: Upgrading of liquids derived from biomass processing, 1st ed.,
vol. 56. Elsevier Inc., 2013.
[27] N. Doassans-Carrère, J.-H. Ferrasse, O. Boutin, G. Mauviel, and J. Lédé, “Comparative
Study of Biomass Fast Pyrolysis and Direct Liquefaction for Bio-Oils Production:
Products Yield and Characterizations,” Energy & Fuels, vol. 28, no. 8, pp. 5103–5111,
Aug. 2014.
[28] R. Shakya, S. Adhikari, R. Mahadevan, E. B. Hassan, and T. A. Dempster, “Catalytic

123
upgrading of bio-oil produced from hydrothermal liquefaction of Nannochloropsis sp.,”
Bioresour. Technol., vol. 252, no. November 2017, pp. 28–36, 2018.
[29] J. Li, X. Fang, J. Bian, Y. Guo, and C. Li, “Microalgae hydrothermal liquefaction and
derived biocrude upgrading with modified SBA-15 catalysts,” Bioresour. Technol., vol.
266, no. May, pp. 541–547, 2018.
[30] D. C. Elliott, “Historical developments in hydroprocessing bio-oils,” Energy and Fuels,
vol. 21, no. 3, pp. 1792–1815, 2007.
[31] N. R. Singh, Biofuels, no. 5. 2011.
[32] P. Grammelis, Solid Biofuels for Energy. 2011.
[33] H. D. Gesser, “Gaseous Fuels,” in APPLIED CHEMISTRY, 2002, pp. 93–113.
[34] C. B. Britt Rob, “Plants at the Pump Biofuels, Climate Change, and Sustainability.”
[35] S. A. Markov, “Biofuels and Synthetic Fuels,” Applied Science. .
[36] M. R. D. Riazi Chiaramonti, Biofuels Production and Processing Technology. 2018.
[37] P. K. Sarangi, S. Nanda, and P. Mohanty, Recent Advancements in Biofuels and Bioenergy
Utilization. .
[38] A. A. Peterson, F. Vogel, R. P. Lachance, M. Fröling, M. J. Antal, and J. W. Tester,
“Thermochemical biofuel production in hydrothermal media: A review of sub- and
supercritical water technologies,” Energy Environ. Sci., vol. 1, no. 1, pp. 32–65, 2008.
[39] R. C. Ã. Saxena, D. K. Adhikari, and H. B. Goyal, “Biomass-based energy fuel through
biochemical routes : A review,” vol. 13, pp. 167–178, 2009.
[40] G. Ragauskas, Arthur J. Williams, Charlotte K. Davinson, Brian H. Britovsek, “The Path
Forward for Biofuels and Biomaterials,” Am. Assoc. Adv. Sci., vol. 311, no. 274, pp. 484–
489, 2006.
[41] M. Balat, Fuels from Biomass – Overview, no. 5. 2011.
[42] K. Dutta, A. Daverey, and J. G. Lin, “Evolution retrospective for alternative fuels: First to
fourth generation,” Renew. Energy, vol. 69, pp. 114–122, 2014.
[43] G. A. Ban-Weiss, J. Y. Chen, B. A. Buchholz, and R. W. Dibble, “A numerical
investigation into the anomalous slight NOxincrease when burning biodiesel; A new (old)
theory,” Fuel Process. Technol., vol. 88, no. 7, pp. 659–667, 2007.
[44] M. Patel and A. Kumar, “Production of renewable diesel through the hydroprocessing of
lignocellulosic biomass-derived bio-oil: A review,” Renew. Sustain. Energy Rev., vol. 58,

124
pp. 1293–1307, 2016.
[45] S. Xiu and A. Shahbazi, “Bio-oil production and upgrading research: A review,” Renew.
Sustain. Energy Rev., vol. 16, no. 7, pp. 4406–4414, 2012.
[46] A. Ray and A. Anumakonda, Hydrocarbon Fuels, 1st ed. Elsevier Inc., 2011.
[47] P. C. Hallenbeck, Microbial Technologies in Advanced Biofuels Production. .
[48] K. Chandrasekhar, Y. Lee, and D. Lee, “Biohydrogen Production : Strategies to Improve
Process Efficiency through Microbial Routes,” pp. 8266–8293, 2015.
[49] BP Energy Economics, “BP Energy Outlook 2019 edition,” 2019.
[50] ExxonMobil, “The Outlook for Energy : A View to 2040,” Outlook, p. 43, 2012.
[51] W. E. Outlook, “World Energy Outlook 2018,” 2018.
[52] Equinor, “Energy Perspectives 2018,” 2018.
[53] D. O. Hall and P. De Groot, “Biomass energy Lessons from case studies in developing
countries,” 1992.
[54] T. A. Mckeon, D. G. Hayes, D. F. Hildebrand, and R. J. Weselake, Introduction to
Industrial Oil Crops. Elsevier Inc., 2016.
[55] D. G. Hayes, Polymeric Products Derived From Industrial Oils for Paints , Coatings , and
Other Applications. AOCS Press., 2016.
[56] A. Demirbas, “Characterization of biodiesel fuels,” Energy Sources, Part A Recover. Util.
Environ. Eff., vol. 31, no. 11, pp. 889–896, 2009.
[57] R. Friedt, Wolfgang,. Snowdon, Handbook of Plant Breeding Oil Crops. 2009.
[58] C. J. H. Luque, R. Campelo, J.M., Handbook of Biofuels Production Processes And
Technologies.pdf. 2011.
[59] I. H. Williams, Biocontrol-Based Integrated Management of Oilsee Rape Pets. 2010.
[60] A. C. W. Metzger, “Biocontrol of Oilseed Rape Pests,” vol. 97, no. 5, 2019.
[61] A. R. C. Stephenson and H. N. Pitre, “Influence of Soybean Maturity Group and Row
Width on Bean Leaf Beetle ( Coleoptera : Chrysomelidae ) and Bean Pod Mottle Disease
in an Early Season Production System.”
[62] A. L. S. Hesler, “Inventory and Assessment of Foliar Natural Enemies of the Soybean
Aphid ( Hemiptera : Aphididae ) in South Dakota Inventory and Assessment of Foliar
Natural Enemies of the Soybean Aphid ( Hemiptera : Aphididae ) in South Dakota,” vol.
43, no. 3, pp. 577–588, 2019.

125
[63] A. K. F. Chang, S. F. Hwang, H. U. Ahmed, S. E. Strelkov, and M. W. Harding, “Disease
reaction to Rhizoctonia solani and yield losses in soybean Disease reaction to Rhizoctonia
solani and yield losses in soybean,” vol. 98, no. 1, pp. 115–124, 2019.
[64] A. Demirbas, Biodiesel, A Realistic Fuel Alternative for Diesel Engines. 2008.
[65] U. S. D. of A. (Farm S. Agency), “ENERGY Biomass Crop Assistance Program for Fiscal
Year 2017,” 2017.
[66] N. Koehler, J. Mccaherty, C. W. Rfa, and G. Cooper, “2019 Ethanol Industry Outlook
(RFA),” 2019.
[67] L. Canilha et al., “Bioconversion of sugarcane biomass into ethanol: An overview about
composition, pretreatment methods, detoxification of hydrolysates, enzymatic
saccharification, and ethanol fermentation,” J. Biomed. Biotechnol., vol. 2012, 2012.
[68] T. Drapcho, Caye M. Nghim, Nhuan. Walker, “Biofuel feedstocks,” Access Eng., pp. 1–2,
2012.
[69] D. R. Weightman, Richard M, Kindred, “MEETING EU BIOETHANOL TARGETS
THROUGH IMPROVEMENTS IN WHEAT GRAIN QUALITY AND
PRODUCTIVITY IN EUROPE Richard,” 2018, pp. 49–73.
[70] R. F. ASSOCIATION, “ETHANOL INDUSTRY OUTLOOK 2018,” 2018.
[71] T. Capehart, O. Liefert, and D. Olson, “Feed Outlook First Projection for 2019 / 20 : Large
Crop and Low,” no. March 2020, 2019.
[72] BAYER Crop Science United States, “Corn Diseases Threaten Yields | Crop Science US,”
May 6 2019. [Online]. Available: https://www.cropscience.bayer.us/learning-
center/articles/corn-diseases-threaten-yields#phcontent_5_divAccordion. [Accessed: 10-
Jun-2019].
[73] Biomass Crop Assistance Program. United States Department of Agriculture Farm Service
Agency., “Energy-Programs,” 2019. [Online]. Available:
https://www.fsa.usda.gov/programs-and-services/energy-programs/BCAP/index.
[Accessed: 10-Jun-2019].
[74] S. Barros, “Sugar Annual Report 2019 (USDA),” Sugar Annu., 2009.
[75] D. Pimentel and T. W. Patzek, “Ethanol production using corn, switchgrass, and wood;
biodiesel production using soybean and sunflower,” Food, Energy, Soc. Third Ed., vol. 14,
no. 1, pp. 311–332, 2007.

126
[76] Ma. C. Gonçalves, L. R. Pinto, S. S. Creste, and M. G. A. Landell, “Virus Diseases of
Sugarcane . A Constant Challenge to Sugarcane Breeding in Brazil,” no. January 2015,
2012.
[77] D. Zilberman, R. Goetz, and A. Garrido, Production of Ethanol from Sugarcane in Brazil.
2014.
[78] M. Balat and G. Ayar, “Biomass Energy in the World , Use of Biomass and Potential
Trends Biomass Energy in the World , Use of Biomass and,” vol. 8312, 2006.
[79] A. Alzagameem et al., “applied sciences Low-Input Crops as Lignocellulosic Feedstock
for Second-Generation Biorefineries and the Potential of Chemometrics in Biomass
Quality Control,” 2019.
[80] D. Mohan, C. U. Pittman, and P. H. Steele, “Pyrolysis of Wood / Biomass for Bio-oil : A
Critical Review,” no. 4, pp. 848–889, 2006.
[81] W. M. Sims, Ralph E H, Michael Taylor and Jack Saddler, “FROM FIRST TO SECOND
GENERATION TECHNOLOGIES REPORT Biofuel Technologies,” no. November,
2008.
[82] C. U. Jensen, J. K. R. Guerrero, S. Karatzos, G. Olofsson, and S. B. Iversen,
HydrofactionTM of forestry residues to drop-in renewable transportation fuels. Elsevier
Ltd., 2018.
[83] E. De Jong and R. J. A. Gosselink, Lignocellulose-Based Chemical Products. Elsevier,
2014.
[84] D. Fengel and G. Wegener, Wood: chemistry, ultrastructure, reactions. Berlin, 1989.
[85] P. Mäki-arvela, I. Anugwom, P. Virtanen, R. Sjöholm, and J. P. Mikkola, “Dissolution of
lignocellulosic materials and its constituents using ionic liquids — A review,” Ind. Crop.
Prod., vol. 32, no. 3, pp. 175–201, 2010.
[86] E. W. Qian, Pretreatment and Saccharification of Lignocellulosic Biomass. Elsevier,
2014.
[87] A. García, M. G. Alriols, and J. Labidi, “Evaluation of different lignocellulosic raw
materials as potential alternative feedstocks in biorefinery processes,” Ind. Crop. Prod.,
vol. 53, pp. 102–110, 2014.
[88] T. Ikeda, K. Holtman, J. F. Kadla, and H.-Mi. Ckang, “Studies on the Effect of Ball
Milling on Lignin Structure Using a Modified DFRC Method,” 2002.

127
[89] F. H. Isikgor and C. R. Becer, “Lignocellulosic biomass: a sustainable platform for the
production of bio-based chemicals and polymers,” Polym. Chem., vol. 6, no. 25, pp. 4497–
4559, Jun. 2015.
[90] P. Chen and L. Peng, “The Diversity of Lignocellulosic Biomass Resources and their
Evaluation for Use as Biofuels and Chemicals,” in Royal Society of Chemistry, no. 10, .
[91] Grand View Research, “Lignin Market Size | Global Industry Growth Analysis Report,
2019-2025,” March 2019, 2019. [Online]. Available:
https://www.grandviewresearch.com/industry-analysis/lignin-
market?utm_source=google&utm_medium=cpc&utm_campaign=AdWords_Lignin_Type
3_CMFE&gclid=EAIaIQobChMIht2XndXh4gIVkvhkCh2QBgTxEAAYASAAEgJsVvD
_BwE. [Accessed: 11-Jun-2019].
[92] J. Sheehan, T. Dunahay, J. Benemann, and P. Roessler, “A Look Back at the U . S .
Department of Energy ’ s Aquatic Species Program : Biodiesel from Algae,” 1998.
[93] F. Alam, A. Date, R. Rasjidin, S. Mobin, and H. Moria, “Biofuel from algae- Is it a viable
alternative ?,” vol. 49, pp. 221–227, 2012.
[94] T. Widmar and F. Loewus, Plant Carbohydrates (Encyclopedia of Plant Physiology).
Regensburg/Washington, 1981.
[95] J. Singh and S. Gu, “Commercialization potential of microalgae for biofuels production,”
Renew. Sustain. Energy Rev., vol. 14, no. 9, pp. 2596–2610, 2010.
[96] G. Dragone and B. D. Fernandes, “Third generation biofuels from microalgae Third
generation biofuels from microalgae,” no. January, 2010.
[97] L. Brennan and P. Owende, “Biofuels from microalgae — A review of technologies for
production , processing , and extractions of biofuels and co-products,” vol. 14, pp. 557–
577, 2010.
[98] G. S. Murthy, Overview and assessment of algal biofuels production technologies, 1st ed.,
no. Table 1. Elsevier Inc., 2011.
[99] E. Suali and R. Sarbatly, “Conversion of microalgae to biofuel,” Renew. Sustain. Energy
Rev., vol. 16, no. 6, pp. 4316–4342, 2012.
[100] J. Lu, C. Sheahan, and P. Fu, “Metabolic engineering of algae for fourht generation
biofuels production,” Energy Environ. Sci., pp. 2451–2466, 2011.
[101] R. Radakovits, R. E. Jinkerson, A. Darzins, and M. C. Posewitz, “Genetic Engineering of

128
Algae for Enhanced Biofuel Production ᰔ,” vol. 9, no. 4, pp. 486–501, 2010.
[102] S. Sasmal and K. Mohanty, “Pretreatment of Lignocellulosic Biomass Toward Biofuel
Production,” no. March 2019, 2018.
[103] P. Kumar et al., “Methods for Pretreatment of Lignocellulosic Biomass for Efficient
Hydrolysis and Biofuel Production Methods for Pretreatment of Lignocellulosic Biomass
for Efficient Hydrolysis and Biofuel Production,” pp. 3713–3729, 2009.
[104] Z. Tong, N. Cheng, and P. Pullammanappallil, “Pretreatment of Ligno-cellulosic Biomass
for Biofuels and Bioproducts,” 2019.
[105] E. Khullar, B. S. Dien, K. D. Rausch, M. E. Tumbleson, and V. Singh, “Effect of particle
size on enzymatic hydrolysis of pretreated Miscanthus,” Ind. Crop. Prod., vol. 44, pp. 11–
17, 2013.
[106] J. Sanchez and C. A. Cardona, “Trends in biotechnological production of fuel ethanol
from different feedstocks,” Bioresour. Technol., vol. 99, pp. 5270–5295, 2008.
[107] M. A. das Neves, T. Kimura, N. Shimizu, and M. Nakajima, “State of the Art and Future
Trends of Bioethanol Production State of the Art and Future Trends of Bioethanol
Production,” Dyn. Biochem. Process Biotechnol. Mol. Biol., no. June 2014, 2007.
[108] N. Sarkar, S. K. Ghosh, S. Bannerjee, and K. Aikat, “Bioethanol production from
agricultural wastes : An overview,” Renew. Energy, vol. 37, no. 1, pp. 19–27, 2012.
[109] M. Ferdes, G. Paraschiv, N. Ungureanu, and V. Vladut, “Lignocellulosic biomass
pretreatment for biofuel production,” Bucharest, 2017.
[110] J. Y. Zhu and X. J. Pan, “Bioresource Technology Woody biomass pretreatment for
cellulosic ethanol production : Technology and energy consumption evaluation q,”
Bioresour. Technol., vol. 101, no. 13, pp. 4992–5002, 2010.
[111] J. Y. Zhu, X. Pan, R. S. Z. Jr, and W. Forest, “Pretreatment of woody biomass for biofuel
production : energy efficiency , technologies , and recalcitrance,” pp. 847–857, 2010.
[112] Z. Zhang, “Biogasification of rice straw with an anaerobic-phased system solids digester,”
Bioresour. Technol., vol. 68, pp. 236–245, 1999.
[113] A. Abraham, A. K. Mathew, R. Sindhu, A. Pandey, and P. Binod, “Bioresource
Technology Potential of rice straw for bio-refining : An overview,” Bioresour. Technol.,
vol. 215, pp. 29–36, 2016.
[114] M. Chen, J. Zhao, and L. Xia, “Enzymatic hydrolysis of maize straw polysaccharides for

129
the production of reducing sugars,” vol. 71, pp. 411–415, 2008.
[115] C. Wan and Y. Li, “Bioresource Technology Microbial pretreatment of corn stover with
Ceriporiopsis subvermispora for enzymatic hydrolysis and ethanol production,”
Bioresour. Technol., vol. 101, no. 16, pp. 6398–6403, 2010.
[116] A. Kamari and S. Ishak, “A review of optimum conditions of transesterification process
for biodiesel production from various feedstocks,” Int. J. Environ. Sci. Technol., vol. 16,
no. 5, pp. 2481–2502, 2019.
[117] N. Savage, “The ideal biofuel,” Nature, vol. 474, no. 7352, pp. S9–S11, 2011.
[118] J. Sun, S. Ding, and J. O. Y. Doran-peterson, “Biomass and its Biorefinery : Novel
Approaches from Nature-Inspired Strategies and Technology,” in Royal Society of
Chemistry, no. 10, pp. 1–13.
[119] M. M. Kucuk and A. Demirbas, “Biomass conversion processes,” Energy Convers., vol.
38, no. 2, pp. 151–165, 1997.
[120] C. A. Cardona, O. J. Sanchez, and L. F. Gutierrez, PROCESS SYNTHESIS FOR FUEL
ETHANOL PRODUCTION. 2010.
[121] S. Hajar et al., “Yeasts in sustainable bioethanol production : A review,” vol. 10, no.
March, pp. 52–61, 2017.
[122] J. Otera, “Transesterification,” Chem. Rev., vol. 93, no. 4, pp. 1449–1470, 1993.
[123] L. C. Meher, D. V. Sagar, and S. N. Naik, “Technical aspects of biodiesel production by
transesterification — a review,” vol. 10, pp. 248–268, 2006.
[124] I. M. Atadashi, M. K. Aroua, and A. A. Aziz, “Biodiesel separation and puri fi cation : A
review,” ScienceDirect, vol. 36, pp. 437–443, 2011.
[125] A. Birla, B. Singh, S. N. Upadhyay, and Y. C. Sharma, “Bioresource Technology Kinetics
studies of synthesis of biodiesel from waste frying oil using a heterogeneous catalyst
derived from snail shell,” Bioresour. Technol., vol. 106, pp. 95–100, 2012.
[126] B. Y. E. Lotero, J. G. Goodwin, D. A. Bruce, K. Suwannakarn, Y. Liu, and D. E. Lopez,
The Catalysis of Biodiesel Synthesis, vol. 19, no. Table 2. 2006.
[127] A. Demirbas, “Progress and recent trends in biodiesel fuels,” Energy Convers. Manag.,
vol. 50, no. 1, pp. 14–34, 2009.
[128] W. S. Donahue, J. C. Brandt, W. S. Donahue, and J. C. Brandt, “Pyrolysis: Types,
Processes, and Industrial Sources and Products,” p. 563, 2009.

130
[129] P. Basu, “Pyrolysis,” in Biomass Gasification, Pyrolysis and Torrefaction : Practical
Design and Theory, Elsevier Inc., 2019, pp. 147–176.
[130] A. V Bridgwater, “Renewable fuels and chemicals by thermal processing of biomass,”
vol. 91, pp. 87–102, 2003.
[131] C. Hognon et al., “Comparison of pyrolysis and hydrothermal liquefaction of
Chlamydomonas reinhardtii. Growth studies on the recovered hydrothermal aqueous
phase,” Biomass and Bioenergy, vol. 73, pp. 23–31, Feb. 2015.
[132] P. Basu, “Biomass Gasification, Pyrolysis and Torrefaction,” 2013.
[133] M. N. Uddin, K. Techato, J. Taweekun, and M. Rahman, “An Overview of Recent
Developments in Biomass Pyrolysis Technologies,” 2018.
[134] A. V Bridgwater and G. V. C. Peacocke, “Fast pyrolysis processes for biomass,” Renew.
Sustain. Energy Rev., vol. 4, 2000.
[135] D. J. Nowakowski, A. V Bridgwater, D. C. Elliott, D. Meier, and P. De Wild, “Journal of
Analytical and Applied Pyrolysis Lignin fast pyrolysis : Results from an international
collaboration,” Renew. Sustain. Energy Rev., vol. 88, no. 1, pp. 53–72, 2010.
[136] P. Basu, “Gasification Theory,” in Biomass Gasification, Pyrolysis and Torrefaction,
Elsevier Inc., 2013, pp. 199–248.
[137] S. De, A. K. Agarwal, V. S. Moholkar, B. Thallada, R. Advances, and F. Challenges, Coal
and Biomass Gasification. 2018.
[138] C. Higman and S. Tam, “Advances in coal gasification, hydrogenation, and gas treating
for the production of chemicals and fuels,” Chem. Rev., vol. 114, no. 3, pp. 1673–1708,
2014.
[139] R. Singh, A. Prakash, and B. Balagurumurthy, “Hydrothermal liquefaction of agricultural
and forest biomass residue : comparative study,” J. Mater. Cycles Waste Manag., pp. 442–
452, 2015.
[140] P. Basu, “Hydrothermal Conversion of Biomass,” in Biomass Gasification, Pyrolysis and
Torrefaction, 2018, pp. 331–371.
[141] S. S. Toor, L. Rosendahl, and I. Sintamarean, “Recipe-based co-HTL of biomass and
organic waste,” in Direct Thermochemical Liquefaction for Energy Applications, Aalborg:
Elsevier Ltd., 2018, pp. 169–189.
[142] Z. Wang, “REACTION MECHANISMS OF HYDROTHERMAL LIQUEFACTION OF

131
MODEL COMPOUNDS AND BIOWASTE FEEDSTOCKS,” University of Illinois,
2011.
[143] C. Gai, Y. Zhang, W. T. Chen, P. Zhang, and Y. Dong, “An investigation of reaction
pathways of hydrothermal liquefaction using Chlorella pyrenoidosa and Spirulina
platensis,” Energy Convers. Manag., vol. 96, pp. 330–339, 2015.
[144] Y. Zhang and W. Chen, “Hydrothermal liquefaction of protein-containing feedstocks,” in
Direct Thermochemical Liquefaction for Energy Applications, no. 2, Elsevier Ltd., 2018,
pp. 127–168.
[145] R. Singh, A. Prakash, B. Balagurumurthy, and T. Bhaskar, “Hydrothermal Liquefaction
of,” Elsevier, 2015, pp. 269–291.
[146] N. Kobayashi, N. Okada, Y. Tanabe, and Y. Itaya, “Fluid behavior of woody biomass
slurry during hydrothermal treatment,” Ind. Eng. Chem. Res., vol. 50, no. 7, pp. 4133–
4139, 2011.
[147] R. Singh, T. Bhaskar, and B. Balagurumurthy, “Hydrothermal Upgradation of Algae into
Value-added Hydrocarbons,” in Biofuels from Algae, Elsevier B.V., 2013, pp. 235–260.
[148] D. C. Elliott, “Historical developments in hydroprocessing bio-oils,” Energy and Fuels,
vol. 21, no. 3, pp. 1792–1815, 2007.
[149] A. R. K. Gollakota, N. Kishore, and S. Gu, “A review on hydrothermal liquefaction of
biomass,” Renew. Sustain. Energy Rev., vol. 81, no. April 2017, pp. 1378–1392, 2018.
[150] A. Montanez, “Paper Review: ‘Application of UNIQUAC Model to Predict Liquid-Liquid
Equilibrium,’” pp. xiii–xiv, 2017.
[151] P. E. Savage and C. Rearrangements, “Organic Chemical Reactions in Supercritical
Water,” 1999.
[152] D. C. Elliott, “Historical developments in hydroprocessing bio-oils,” Energy and Fuels,
vol. 21, no. 3, pp. 1792–1815, 2007.
[153] P. Haghighat, A Montanez, G. Aguliera, J. K. Rodriguez, S. Karatzos, M.A. Clarke, and
W. McCaffrey “Hydrotreating of Hydrofaction TM biocrude in the presence of presulfided
commercial catalysts †,” pp. 744–759, 2019.
[154] P. Haghighat, A. Montanez, J. Rodriguez, S. Karatzos, and W. Mccaffrey, “Two-Stage
Catalytic Hydrotreating of HTL Biocrude,” Edmonton, 2019.
[155] M. Trujillo, “Catalytic Upgrading Process of Ligno-cellulose Derived Heavy Crude Oil,”

132
Univerity of Calgary, 2018.
[156] A. Khosravanipour Mostafazadeh, O. Solomatnikova, P. Drogui, and R. D. Tyagi, A
review of recent research and developments in fast pyrolysis and bio-oil upgrading, vol. 8,
no. 3. Biomass Conversion and Biorefinery, 2018.
[157] M. R. Gray, Heavy Oil-Refining. Edmonton: The Universsity of Alberta Press, 2015.
[158] K. O. Albrecht et al., “Impact of heterotrophically stressed algae for biofuel production
via hydrothermal liquefaction and catalytic hydrotreating in continuous-flow reactors,”
Algal Res., vol. 14, pp. 17–27, 2016.
[159] U. Jena, K. C. Das, and J. R. Kastner, “Effect of operating conditions of thermochemical
liquefaction on biocrude production from Spirulina platensis,” Bioresour. Technol., vol.
102, no. 10, pp. 6221–6229, 2011.
[160] D. C. Elliott, P. Biller, A. B. Ross, A. J. Schmidt, and S. B. Jones, “Hydrothermal
liquefaction of biomass: Developments from batch to continuous process,” Bioresour.
Technol., vol. 178, 2015.
[161] H. Yang, J. Yao, G. Chen, W. Ma, B. Yan, and Y. Qi, “Overview of upgrading of
pyrolysis oil of biomass,” Energy Procedia, vol. 61, pp. 1306–1309, 2014.
[162] X. Li et al., “Upgrading of bio-oil into advanced biofuels and chemicals. Part III. Changes
in aromatic structure and coke forming propensity during the catalytic hydrotreatment of a
fast pyrolysis bio-oil with Pd/C catalyst,” Fuel, vol. 116, pp. 642–649, 2014.
[163] S. Xiu and A. Shahbazi, “Bio-oil production and upgrading research: A review,” Renew.
Sustain. Energy Rev., vol. 16, no. 7, pp. 4406–4414, 2012.
[164] M. Ikura, S. Mirmiran, M. Stanciulescu, and H. Sawatzky, “United States Patent #
5820640,” 1998.
[165] Y. Shi, E. Xing, K. Wu, J. Wang, M. Yang, and Y. Wu, “Recent progress on upgrading of
bio-oil to hydrocarbons over metal/zeolite bifunctional catalysts,” Catal. Sci. Technol.,
vol. 7, no. 12, pp. 2385–2415, 2017.
[166] F. Shi, P. Wang, Y. Duan, D. Link, and B. Morreale, “Recent developments in the
production of liquid fuels via catalytic conversion of microalgae: Experiments and
simulations,” RSC Adv., vol. 2, no. 26, pp. 9727–9747, 2012.
[167] F. A. Twaiq, N. A. M. Zabidi, and S. Bhatia, “Catalytic conversion of palm oil to
hydrocarbons: Performance of various zeolite catalysts,” Ind. Eng. Chem. Res., vol. 38,

133
no. 9, pp. 3230–3237, 1999.
[168] J. Zhang, Z. Luo, Q. Dang, J. Wang, and W. Chen, “Upgrading of bio-oil over
bifunctional catalysts in supercritical monoalcohols,” Energy and Fuels, vol. 26, no. 5, pp.
2990–2995, 2012.
[169] J. WANG, J. CHANG, and J. FAN, “Catalytic esterification of bio-oil by ion exchange
resins,” J. Fuel Chem. Technol., vol. 38, no. 5, pp. 560–564, 2010.
[170] X. Guangwen, L. Xiang, X. Yuan, and Y. Han, “Progress on Upgrading methods of bio-
oil: A review,” Arch. Thermodyn., vol. 33, no. 4, pp. 23–40, 2012.
[171] A. Gruia, “An introduction to crude oil and its processing,” in Handbook of Petroleum
Processing, Dordrecht, 2006, p. 1349.
[172] M. Che, “Nobel Prize in chemistry 1912 to Sabatier: Organic chemistry or catalysis?,”
Catal. Today, vol. 218–219, no. April, pp. 162–171, 2013.
[173] W. O. Library, “Bergius Process,” Hawley’s Condensed Chemical Dictionary, 15-Mar-
2007. [Online]. Available: http://doi.wiley.com/10.1002/9780470114735.hawley01791.
[Accessed: 20-Jun-2019].
[174] P. Raybaud, “Understanding and predicting improved sulfide catalysts: Insights from first
principles modeling,” Appl. Catal. A Gen., vol. 322, no. SUPPL., pp. 76–91, 2007.
[175] X. Li et al., “Heterogeneous sulfur-free hydrodeoxygenation catalysts for selectively
upgrading the renewable bio-oils to second generation biofuels,” Renew. Sustain. Energy
Rev., vol. 82, no. November 2017, pp. 3762–3797, 2018.
[176] M. Badawi et al., “Effect of water on the stability of Mo and CoMo hydrodeoxygenation
catalysts: A combined experimental and DFT study,” J. Catal., vol. 282, no. 1, pp. 155–
164, 2011.
[177] A. Popov et al., “Bio-oil hydrodeoxygenation: Adsorption of phenolic compounds on
sulfided (Co)Mo catalysts,” J. Catal., vol. 297, pp. 176–186, 2013.
[178] M. Badawi et al., “Hydrodésoxygénation de composés phénoliques en présence de
catalyseurs sulfurés (Co)Mo/Al2O3: Une étude expérimentale et théorique,” Oil Gas Sci.
Technol., vol. 68, no. 5, pp. 829–840, 2013.
[179] M. H. Al-dahhan, F. Larachi, M. P. Dudukovic, and A. Laurent, “High-Pressure Trickle-
Bed Reactors : A Review,” pp. 3292–3314, 1997.
[180] R. L. McManus, G. A. Funk, M. P. Harold, and K. M. Ng, “Experimental Study of

134
Reaction in Trickle-Bed Reactors with Liquid Maldistribution,” Ind. Eng. Chem. Res., vol.
32, no. 3, pp. 570–574, 1993.
[181] R. Boesen, “Investigation and Modelling of Diesel Hydrotreating Reactions,” 2010.
[182] S. B. Jones et al., “Production of Gasoline and Diesel from Biomass via Fast Pyrolysis,
Hydrotreating and Hydrocracking: A Design Case,” 2009.
[183] A. A. Forghani, “Catalytic Hydro-cracking of Bio-oil to Bio-fuel,” University of Adelaide,
2014.
[184] A. H. Zacher, M. V Olarte, D. M. Santosa, D. C. Elliott, and S. B. Jones, “A review and
perspective of recent bio-oil hydrotreating research.”
[185] D. C. Elliot, todd R. Hart, G. Neunenschwander, L. Rotness, and A. Zacher, “Catalytic
Hydroprocessing of Biomass Fast Pyrolysis Bio-Oil to Produce Hydrocarbon Products,”
Environ. Prog. Sustain. Energy, vol. 00, no. 00, pp. 1–10, 2009.
[186] P. Grange, E. Laurent, R. Maggi, A. Centeno, and B. Delmon, “Hydrotreatment of
pyrolysis oils from biomass: Reactivity of the various categories of oxygenated
compounds and preliminary techno-economical study,” Catal. Today, vol. 29, no. 1–4, pp.
297–301, 1996.
[187] R. H. Venderbosch, A. R. Ardiyanti, J. Wildschut, A. Oasmaa, and H. J. Heeres,
“Stabilization of biomass-derived pyrolysis oils,” J. Chem. Technol. Biotechnol., vol. 85,
no. 5, pp. 674–686, 2010.
[188] L. Zhang, R. Liu, R. Yin, and Y. Mei, “Upgrading of bio-oil from biomass fast pyrolysis
in China: A review,” Renew. Sustain. Energy Rev., vol. 24, pp. 66–72, 2013.
[189] P. M. Mortensen, J. D. Grunwaldt, P. A. Jensen, K. G. Knudsen, and A. D. Jensen, “A
review of catalytic upgrading of bio-oil to engine fuels,” Appl. Catal. A Gen., vol. 407, no.
1–2, pp. 1–19, 2011.
[190] E. Maria and S. Ferreira, “Assessment of Stage 1 in a Novel Bio-Oil Upgrading Process:
Catalytic Hydrotreating,” 2018.
[191] OMEGA Engineering Canada, “Thermocouple Profile Probes Made from High-Accuracy
Special Limits-of-Error Wire (SLE),” 2003. [Online]. Available:
https://www.omega.ca/en/sensors-and-sensing-
equipment/temperature/sensors/thermocouple-probes/p/PP3-PP6-PP10. [Accessed: 25-Jul-
2019].

135
[192] ASTM International, “Standard Test Method for Density and Relative Density of Crude
Oils by Digital Density Analyzer,” Man. Hydrocarb. Anal. 6th Ed., vol. 05, no.
September, pp. 778-778–5, 2008.
[193] ASTM International D445-06, “Standard Test Method for Kinematic Viscosity of
Transparent and Opaque Liquids (and Calculation of Dynamic Viscosity).,” Man.
Hydrocarb. Anal. 6th Ed., pp. 1–10, 2008.
[194] ASTM International, “Standard Test Method for Acid Number of Petroleum Products by
Potentiometric Titration,” Man. Hydrocarb. Anal. 6th Ed., vol. i, pp. 159-159–7, 2008.
[195] ASTM International, “Standard Test Methods for Instrumental Determination of Carbon ,
Hydrogen , and Nitrogen in Petroleum Products and Lubricants 1,” vol. 02, no.
Reapproved, pp. 1–7, 2007.
[196] ASTM International, “Standard Test Method for Determination of Gaseous Compounds
by Extractive Direct Interface Fourier Transform Infrared ( FTIR ) Spectroscopy 1,” vol.
03, no. 2010, pp. 1–16, 2019.
[197] ASTM International, “Standard Test Method for Heat of Combustion of Liquid
Hydrocarbon Fuels by Bomb,” pp. 1–10, 2017.
[198] ASTM International, “Standard Test Method for Boiling Point Distribution of Samples
with Residues Such as Crude Oils and Atmospheric and Vacuum Residues by,” pp. 1–20,
2018.
[199] ASTM International, “Standard Test Method for Mass Loss and Residue Measurement
Validation of,” vol. 11, no. Reapproved 2017, pp. 2–7, 2017.
[200] R. M. Egeberg and L. Skyum, “Novel hydrotreating technology for production of green
diesel,” 2009.
[201] F. D. M. Mercader, M. J. Groeneveld, S. R. A. Kersten, R. H. Venderbosch, and J. A.
Hogendoorn, “Pyrolysis oil upgrading by high pressure thermal treatment,” Fuel, vol. 89,
no. 10, pp. 2829–2837, 2010.
[202] A. H. Zacher, M. V. Olarte, D. M. Santosa, D. C. Elliott, and S. B. Jones, “A review and
perspective of recent bio-oil hydrotreating research,” Green Chem., vol. 16, no. 2, pp.
491–515, 2014.
[203] A. A. Andreev, V. J. Kafedjiysky, and R. M. Edreva-kardjieva, “Active forms for water-
gas shift reaction on NiMo-sul ® de catalysts,” vol. 179, no. September 1998, pp. 223–

136
228, 1999.
[204] P. Biller and A. Roth, “Hydrothermal Liquefaction : A Promising Pathway Towards
Renewable Jet Fuel,” in Biokerosene, 2018, pp. 607–635.
[205] P. Haghighat, “Processing of Solubilized Asphaltene in Aqueous Media,” University of
Calgary, 2016.
[206] H. S. Fogler, “Distributions of Residence Times for Chemical Reactors 13,” in Elements
of Chemical Reaction Engineering, 2008, pp. 867–944.
[207] E. Churin, “Upgrading of Pyrolysis Oils by Hydrotreatment E.,” in Biomass Pyrolysis
Liquids Upgrading and Utilisation, A. V. Bridgwater and G. Grassi, Eds. Brussels and
Luxembuourg, 1991, p. 363.
[208] D. C. Elliott et al., “Process development for hydrothermal liquefaction of algae
feedstocks in a continuous- fl ow reactor,” ALGAL Res., vol. 2, no. 4, pp. 445–454, 2013.
[209] M. S. Haider, D. Castello, and K. M. Michalski, “Catalytic Hydrotreatment of Microalgae
Biocrude from Continuous Hydrothermal Liquefaction : Heteroatom Removal and Their
Distribution in Distillation Cuts,” Energies, 2018.
[210] J. A. Ramirez, R. J. Brown, and T. J. Rainey, “A review of hydrothermal liquefaction bio-
crude properties and prospects for upgrading to transportation fuels,” Energies, vol. 8, no.
7, pp. 6765–6794, 2015.
[211] V. A. Yakovlev et al., “Development of new catalytic systems for upgraded bio-fuels
production from bio-crude-oil and biodiesel,” Elsevier, vol. 144, pp. 362–366, 2009.
[212] D. Castello, M. S. Haider, and L. A. Rosendahl, “Catalytic upgrading of hydrothermal
liquefaction biocrudes: Different challenges for different feedstocks,” Renew. Energy, vol.
141, pp. 420–430, 2019.

137
Appendix A: Mass Balance Calculation

After reaching steady-state at fixed operating conditions, mass balances of the reactants coming in
and the products coming out from the process were performed over a specified time interval. For
this investigation, the time interval of the mass balance was defined as follows: time zero
corresponds to the time when the empty heavy and light tanks started the collection of reacted
products. The final time of the mass balance corresponds to the time when the lecture of the wet
gas meter was recorded, and the collection of the products were stopped by switching the tanks in
service to the empty ones.
The mass balances, including heavy products, light products (lights +water produced) and gases,
were calculated at different intervals of time applying the following equation:

𝐹𝑒𝑒𝑑 (𝑖𝑛) + 𝐻𝑦𝑑𝑟𝑜𝑔𝑒𝑛 (𝑖𝑛) = 𝑇𝑜𝑡𝑎𝑙 𝐿𝑖𝑞𝑢𝑖𝑑 𝑃𝑟𝑜𝑑𝑢𝑐𝑡 (𝑜𝑢𝑡) + 𝐺𝑎𝑠𝑒𝑠 (𝑜𝑢𝑡) Equation A.1

To calculate the mass flow rate of the feed, the volumetric flow given by the pump was converted
to mass flow, and it was multiplied by the density of the feedstock in kg/cm3.

The hydrogen mass flow rate coming into the reaction unit was determined by converting the
volumetric flow rate provided by the mass flow controller to mass flow. Assuming ideal gas, the
following equation was applied:
𝑄𝐻2 ∗𝑃∗𝑀𝑊
𝑊𝐻2 = Equation A.2
𝑅∗𝑇

Where:
𝑊𝐻2 = Mass flow rate of hydrogen (g/min)
𝑄𝐻2 = Volumetric flow rate of hydrogen (Nm3/min)
𝑃= Pressure (Pa)
𝑀𝑊 = Molecular weight of the hydrogen (g/mol)
𝑅 = Universal gas constant = 8.314 J/mol*K
T = Temperature (K)
Ideal gas was assumed because around 99% of the gas was composed of hydrogen; it was detected
with the gas analysis composition from the GC analysis.

138
The total liquid product includes the heavy and light products generated during the process. They
were collected from the tanks HTK-1 and 2 and LTK-3 and 4. When the mass balance time was
completed, the liquid samples from the heavy and light tanks were collected in glass vessels
previously labelled and weighted. Then, the total mass of the collected sample was calculated by
difference. The liquid mass balance was calculated according to Equation 3.4:

𝑇𝑜𝑡𝑎𝑙 𝑙𝑖𝑞𝑢𝑖𝑑 𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠 (𝑔)


𝐿𝑖𝑞𝑢𝑖𝑑 𝑚𝑎𝑠𝑠 𝑏𝑎𝑙𝑎𝑛𝑐𝑒(%) = 𝑇𝑜𝑡𝑎𝑙 𝑙𝑖𝑞𝑢𝑖𝑑 𝑓𝑒𝑒𝑑 𝑝𝑢𝑚𝑝𝑒𝑑 (𝑔) × 100 Equation A.3

In the case of the produced gases, the volumetric flow reading taken from the wet test meter was
converted to molar flow using the ideal gas equation. Due to the gas stream is mainly composed
of hydrogen, the molar flow was multiplied by the hydrogen molecular weight to determine the
corresponding molar flow.
It was assumed that the hydrotreater unit was at steady-state conditions when two or three
consecutive mass balances were around 98%. Afterwards, the collected light products were
transferred to a 300 ml glass jar (weighted). It was settled at ambient conditions for about 1 hour
in order to separate the sample into two phases. The upper phase, which was mainly composed of
light upgraded products, was withdrawn using a pipette. The remaining water phase was quantified
by weighing the jar again. The water yield was determined by dividing the mass of water by the
total amount of liquid product.
Next, the dewatered light products were mixed with heavy products. The blended samples were
used to perform the final analyses that were described in chapter 3.

139
Appendix B: Standard Operational Procedure

Process: Hydrotreater operation


Department/Faculty: Chemical and Materials Department / Faculty of Engineering
Location:5-120 CME
Supervisor: Dr. William C. McCaffrey Phone: 780 492 6733 Email: wcm@ualberta.ca
Emergency Contact: Anderson Montanez After-hours Contact: 403 903 9344

SCOPE:
This process is used to investigate the upgrading of HydrofactionTM Biocrude or similar through
hydrotreatment processes using different catalysts and hydrogen. The setup includes dossification,
reaction, and separation stages. The equipment operates at high pressure and high temperature.

Hazard Identification:
• Use of high-pressure hydrogen. Danger of explosion due to leaks.
• Damage to the equipment due to overpressure.
• Loss of airflow in the fumehood can cause accumulation of highly inflammable/toxic gases.
• Condensation of water in cold part of the setup can cause short-circuits if the water gets in
contact with electrical components.
• High temperature causing debilitation of the components leading to leaks and explosion.
• Production of H2S during the reaction causes high concentrations of H2S in the fumehood
and/or lab space.
• Sampling at high pressure causing exposure to chemicals and/or physical injuries.
• The high temperature in the lines can transfer to the valves’ handles causing burns during
sampling.
• The feed can be corrosive leading to premature failure of components.
• Accumulation of liquid in the unit can cause personnel exposure while removing the catalyst
from the packed reactor.
• Spent catalyst may be pyrophoric.
• Heating cords can become damaged and short circuit.
• During the feed preparation, operators can be exposed to harmful chemicals.

140
Training Required:

• General lab safety training


• Safe handling of gas cylinders
• Safe handling of pyrophoric substances

Control/Protective Measures:

• Properly fitted lab coat, safety glasses, floor-length pants and closed-toe shoes should always
be worn
• Long hair should be tied up
• The equipment is installed inside a fumehood. All gas venting should be done inside the
fumehood.
• All substances should be handled inside a fumehood. Transfer of substances from one
fumehood to another should be done in closed containers.
• Generation of H2S is possible during the reaction. If H2S formation is suspected; the personnel
should:
- Always Wear a personal H2S detector.
- Check the H2S detector in front of the door before entering the lab
- Verify that the setup stops automatically and vacate the lab if a H2S alarm is triggered.

Emergency Procedures:
• The unit is controlled through the opto22 and computer (labview). Several interlocks are in
place, including:
• The syringe pump will stop if one or more of the following conditions are detected
• Overpressure as detected by the pump sensors
• Low hydrogen flow
• High H2S concentration inside the fumehood
• The hydrogen flow will stop if overpressure is detected.

141
• All the heating zones but the reactor will stop heating if one or more of the following
conditions are detected:
- Faulty temperature sensor if the sensor is tied to a control loop (excluding the reactors)
- Overtemperature in any sensor (excluding reactors)
• Reactors’ heating zones will stop heating if one or more of the following conditions are
detected:
- Faulty temperature sensor if the sensor is tied to a reactor control loop
- Overtemperature on any reactor sensor.
• All the heating equipment is protected through fuses located in the Opto22 cabinet. All the
equipment is grounded.
• In case of leakage (NOTE: When working around heated zones, stop the electricity flow to
that zone either by turning off the appropriate panel or disconnecting the cable from the panel.
Always inspect the heating cords for damage. If a short circuit is suspected, use a multimeter
to confirm and correct before turning on the power to the heaters)
• If the leak is small, try to correct it if it is possible to do so without putting the personnel and
equipment at risk
• If the leak is bigger, stop the hydrogen and liquid flow. Depressurize the equipment through
the scrubbers. Correct the leak. Replace the faulty parts if needed. Cleanup as much as possible
before restarting
• In case of spill of HTL oil, use a blend of 2-propanol / toluene to clean it.
• H2S could be produced during the reactions at certain conditions.
• If high concentrations of H2S are detected inside the fumehood, the feeding pump will stop
automatically. DO NOT RESTART THE PUMP UNTIL THE CAUSE OF THE HIGH H 2S
CONCENTRATION IS REMOVED.
• If H2S production is suspected, wear a personal H2S detector at all times while inside the lab.
Evacuate the lab if H2S is detected outside the fumehood.
• The personal H2S detectors should be hanged in front of the door to the lab (inside the lab)
while not in use. Before entering the lab, make sure they are not signaling high concentrations
of H2S

Step by Step Procedure of the Process

142
NOTE: All the valves might be hot during operation. Use adequate gloves while manipulating the
valves.

Catalyst loading
• Make sure there is no pressure inside the reactor. Verify ALL the pressure indicators
• Make sure the unit is cold. Turn off the power to the heaters using the switches in the Opto22
cabinet
• Carefully remove the insulation on the top and bottom of the reactors. Remove the heating
cords from the top and bottom tees.
• Disconnect all the power cables (3 each reactor) and thermocouples (7 each reactor)
• Loosen the 3/8 nut on the side of the upper and bottom tee of the reactor.
• Make sure there is space available to put the reactor once removed from the fumehood
• Remove the top reactor disconnecting first the top tube and then, the bottom tube. Be carefully
with the wires hanging from the reactor
• Remove the old catalyst if present. NOTE: some spent hydrotreating catalyst are pyrophoric.
Contact the supplier of the catalyst and/or your supervisor to confirm. If the catalyst if
pyrophoric, make sure is always wet with water.
• To remove the catalyst, both tees might need to be removed from the reactor. Be careful while
removing the bottom tee in order not to damage the multipoint thermocouple.
• Catalyst should be collected and weighted.
• Once all the catalyst is removed, thoroughly clean the reactor. Put special attention to the
sealing surface of the fittings. Inspect the reactor and replace tubing and/ or fittings if they are
damage.
• Replace and tighten the bottom tee. Make sure the mesh to keep the catalyst inside the reactor
is in place.
• Put the reactor in a vertical position.
• Using a plastic hose, keep the multipoint thermocouple in the middle of the reactor and start
loading the catalyst / carborundum. Contact your supervisor for information on how to
assemble the different catalysts beds.
• Once the catalyst is in place, reinstall the top tee and tighten

143
• Remove the bottom reactor and repeat the previous steps
• Once the catalyst is loaded into the second reactor, reinstall the bottom reactor. Tighten up the
3/8 nuts.
• Install the top reactor. Make sure everything is tight
• Make sure all the valves are correctly aligned. Pressurize the equipment to the maximum
operating pressure using N2. Close the valves and check for leaks.
• Once no more leaks are detected, vent the equipment. Pressurize with H2 and check for leaks
using the “Leakator” (combustible gas detector)
• Once the unit is leak-free, reinstall the heating cords and insulation. If the heating cord looks
damaged, check for short circuits to the piping using a multimeter. If short circuits are detected
and cannot be repaired, replace the heating cord.
• Verify with your supervisor if the catalyst needs to be activated previously to the reaction.
Since each catalyst requires different activation procedures, discuss the operation with your
supervisor and co-workers before proceeding to activate the catalyst

Normal operation of the equipment


• Make sure the right catalyst is loaded into the reactors. If needed, refer to the appropriate
procedure.
• Make sure there is enough H2 and N2 in the cylinders for the run.
• Check the level of NaClO solution (household bleach) in the scrubbers. Top it up if needed.
Replace it if the solution looks spent (brownish colour)
• Make sure the fumehood is in working order
• Check the fluid levels on both Julabo Heater and Fisher cooler. Top it if needed (Julabo uses
Duratherm fluid. Cooler can be topped up with distilled water).
• Make sure all the components are on, including
- Computer
- Opto 22
- GC
- GC’s computer
- Teledyne pumps
- Flowmeter controller
144
- Julabo heater
- Fisher Scientific cooler
• Set the Julabo heater at 60 oC. Make sure the fluid is circulating
• Set the Fisher cooler at 4 oC
• Load the file “MSHR_Main” on the computer.
• Activate the catalysts if needed. Refer to the specific instructions for the catalyst(s) and the
appropriate procedure
• Once the activation is finished, depressurize the feed tank.
• Remove the feed tank by disconnecting the gas line, liquid line, thermocouple and power
cable.
• Clean the tank if needed (cleaning agent depends on the liquid that was used previously on
the tank). Use the appropriate PPE that includes, but is not limited to: Lab coat, safety glasses,
appropriate gloves according to the substances to be used. Work inside the fumehood.
• Load the liquid feed in the tank. Quantify the amount of feed in the tank.
• Reconnect the tank in its place (gas line, liquid line, thermocouple and power to the heater)
• Start heating up the tank at 60% power. Once it is close to 60 oC, change the power to the one
listed on the computer (Power to heaters)
• Pressurize the tank to 30 psi using N2. Check for leaks
• Once the tank is leak-free, vent and pressurize it twice more to ensure an inert atmosphere.
• Pressurize it to 30 psi and keep the N2 valve open to maintain the pressure
• Attach a recipient to the purge between the H2/oil mixer and the pressure transducer to collect
the oil.
• Fill up both syringes in the ISCO pump. Fill only to 50 ml. Refer to the appropriate manual.
NOTE: in order to operate the pump, the low-H2-flow-alarm should not be active. If needed,
flow N2 at minimum 100 ml/min and acknowledge the alarm
• Open the purge before the pressure transducer. Run both pumps at 5 ml/min with a nitrogen
flow of 100 ml/min.
• Once the oil starts coming out the purge, stop the pumps and close the purge.
• Refill the pumps up to 50 ml. Do not flow more than 4 ml/min.

145
• Pressurize the whole setup with H2. Set the backpressure valve to the maximum operating
pressure of the run. The maximum operating pressure is 100 bar and is limited by the hydrogen
flowmeter, the pressure switch and the backpressure valve
• Once the operating pressure is achieved; check the entire setup for leaks using the combustible
gas detector (LeakatorTM)
• Set the H2 flow that will be used for the run
• Once the unit is leak-free, start heating the lines that will be used during the run.
- Increase the temperature in manual mode with a power percentage lower than 50%.
- Once the temperature is within 1 or 2 oC from the operating temperature, lower the power
to the one listed on the computer (power to heater) and switch to auto.
- Put the desired set point
- NOTE: Some of the heating zones might require more than 50% power to reach the desired
temperature.
- NOTE: The control of some heating zones is not tuned. Refer to the “power to heaters” file
on the control computer.
• Close one of the heavy product tank and one of the light product tank. Make sure they are
empty
• Check the parameter for the pump to run in continuous flow mode:
- Refill to 10%
- Pump up to 1%
- Refill at 2x the operating flow
- Start logging the opto22 data.
- Open the logging window.
- Click on “open file”
- Write a representative name for the file
- Click open
- Select the timing of each entry
- Start logging
• Increase the pressure of both syringes of the pump; follow the steps:
- Make sure there is no H2 flow alarm on the control system. Acknowledge the alarm if
needed.

146
- Start pumping at the desire operational flow (usually between 0.2 to 1 ml/min), press stop
when the pressure is close to the setpoint
- Repeat the same procedure until the pressure in the pumps is equal to the pressure in the
whole hydrotreater.
• Make sure that the pump is in continuous flow mode and that the right flow set-point is
entered. Press start. NOTE: the pump should be controlled from the control panel. The
computer is not setup to control the pump. Make sure the pump interlock is deactivated.
• Start counting the stabilization time
• Once the stabilization time is over, start a mass balance. For that:
- Double check that the tanks that are not-aligned to the process flow are empty
- Make sure the pressure of the heavy products tank that is not-aligned is roughly the same
as the pressure on the system. Use the N2 line to increase the pressure if needed
- Close the valves to the heavy products and the light products tank that is aligned to the
process flow. Each tank has a valve on top and on the bottom
- Align the lines to the other heavy product and light products tank. Put the indicator tag on
the handle of the valve to the heavy products tank than is aligned
- Slowly open the bottom valve of the heavy products tank that is not-aligned. Collect any
liquid present. NOTE: Double check the indicator tag before opening the bottom valve.
- Repeat with the light-products tank that is not-aligned
- Once no more liquid products are collected, wait 30 minutes and repeat.
- Once liquid products are collected, proceed with the next step
- Make sure all the temperatures, pressure and flow are stable.
- Align one of the heavy product tanks and one of the light product tanks to collect products.
Put the indicator ring on the heavy products tank that is aligned
- Write down the initial gas volume as indicated on the dry test meter.
- Start counting the time for the mass balance
• During the mass balance, analyze the gas products using the online GC.
- Align the 3-way valve so the gas products streams pass through the sampling loop on the
GC.
- Let the gas flow for 5 minutes
- Make sure there is pressure in the sampling line

147
- Start the method on the GC’s computer
- Once the injection is done, close the 3-way valve
• Before the time for the mass balance has elapsed, make sure the heavy -products tank that is
not-aligned is at the operating pressure
• Once the time for the mass balance has elapsed.
- Close both the heavy and the light products tanks that are collecting samples
- Align the heavy and the light products tanks that are empty, to collect sample
- Write down the final gas volume from the dry test meter
- Carefully empty both the heavy and the light products tanks. Quantify the amount
collected.
• Repeat the steps as needed.
• Once the experiment is finished, shut down the unit.

Unit shut down


• Empty both syringes of the pump
• Empty all the product tanks
• Stop the syringe pump and the gas flow to the unit
• Disconnect the feed tank including:
- Gas and venting line
- Liquid line
- Thermocouple
- Power cord
• Connect a spare thermocouple to bypass the interlock on the Opto22
• Prepare a solution of 50% isopropanol in Toluene. 1 or 2 liters are usually enough to clean up
the unit
• Load the solution in the clean-up tank. Connect the liquid and gas / venting line to the clean-
up tank
• Fill up both syringes of the pump with the cleanup solution up to 55 ml
• Align all the equipment. Make sure all the product tanks are aligned to the process stream

148
• Start feeding N2 at 100 ml/ min. Start pumping the cleaning solution at 1 ml/min. (Note:
Adjust the fill rate and the filling up point on the pump to allow for this flow in continuous.)
• After 2 hours, shut down all the heating zones.
• Continue pumping for at least 12 hours or until the solvent comes out clean.
• Once the solvent comes out clean, empty both syringes of the pump. Remove the cleanup tank
from the line.
• Empty all the product tanks. Depressurize the unit

Hazardous Waste Disposal Procedures


• All liquid waste should be properly labeled and sent to the appropriate disposal unit using
Chematix
• All solid waste (spent catalyst) should be collected, properly labeled and weighted. Verify
with your supervisor if the catalyst has to be returned to the supplier. Some of the spent catalyst
might be pyrophoric. Verify with your supervisor if the catalyst to be disposed is pyrophoric
and needs to be in water.
• All gas is passed though the scrubbers filled with household bleach to adsorb any H2S present.
The remaining gas is vented on the upper part of the fumehood.

Equipment Maintenance Procedures

• During normal operation, gas-tightness of the equipment should be checked periodically using
the “LeakatorTM” combustible gas detector
• Level of liquid in the scrubbers should be checked periodically. Liquid should be replaced
when spent (brown color)
• Syringe pump, GC and flowmeters need periodic maintenance. Refer to the appropriate user’s
manual.
• If the main needle of dry test gas meter “jumps”, the meter might have accumulated liquid
inside. Disconnect the meter, drain it and flush it with clean water.
• Clean the filter before the backpressure valve after each run. Replace the filtering element as
needed.
• Recalibrate the H2S sensors every 6 months

149
• The control computer has the automatic updates disabled to prevent it from restarting during
normal operation. Since it is connected to the internet, it is important to install any available
update before startup and after shutting down.
• The feed used in this reactor could be very corrosive. Before startup and after cleaning, check
lines and tanks for indications of corrosion. Replace components as needed.

Prepared By: Gonzalo Rocha Aguilera/Anderson Anibal Montanez Rincon

Date: August 31, 2017

Approved by: William McCaffrey Date: September 2017

Hazzard Assessment:

INSTRUCTIONS:

Tool must be completed by individual(s) who are doing the tasks identified.

Tool can be used to evaluate all tasks for a specific occupation or activity.
- List all work activities in the column 1. It is acceptable to group activities together when the
hazards encountered are the same for all of the activities
- List the existing and potential hazards associated with each task in the column 2, include both
health and safety hazards.
- List the type of hazard encountered in column 3. The hazards are Chemical, Biological,
Physical, Environmental, Ergonomic, Radiation, and Psychosocial.
- Complete the risk analysis and determine the overall risk level by assigning the Incident
Probability (column 4), Incident Severity (column 5) and enter the Risk Level in column 6.
- List the current or proposed controls for each hazard identified in column 7. The complexity
of the controls should be proportional to the overall risk level.
- Identify if the controls are already in place and complete in column 8. It is the responsibility
of the Supervisor to ensure controls are put in place in a reasonable timeframe

150
- Supervisors & Individuals completing the hazard assessment must review and sign off on the
document.
- The document must be kept on file by the supervisor or designate, usually in the training
records for the work site.
- The supervisor or designate must ensure that the results of the hazard assessment including
identified controls are communicated to any impacted employee.
- Supervisors must review relevant portions of the hazard assessment when there is an operating
or infrastructure change.
- The entire hazard assessment also must be reviewed at 3 years interval.

Incident Probability
4=Probable (may happen at least once a year)
3=Occasional (may happen once every 1-5 years)
2=Remote (not likely to happen, but possible once every 5-10 years)
1=Improbable (not likely to happen)

Potential Severity
4=Severe (death, serious injury/illness more than 2 days in the hospital, permanent disability,
property damage > $100,000, extensive environmental damage)
3=Substantial (lost time injury/illness, temporary disability, potential injury, damage >$50,000,
substantial environmental damage, significant adverse public reaction)
2=Minor (medical aid injury, minor illness, minor property damage <$50,000)
1=Minimal (first aid injury)
Risk Value = Incident Probability X Potential Severity
Risk Level
> 11, High Risk (take immediate action to eliminate the risk or implement appropriate controls to
lower the risk)
= 4 – 11, Medium Risk (take timely action to implement appropriate controls to lower or minimize
risk)
< 4 Low Risk (continued operation is permissible with minimal controls).

151
1. Task 2. Existing and 3. Hazard 7. Controls (such as immunizations, health 8. Are
Potential Classification screenings, safe work procedures, Controls in
Risk Value
Hazards (Include supervision, training, contingency plans, Place? If
(Include Chemical, emergency plans and kits, PPE, etc.) not, how
environmental Biological, and when?

5. Severity (1-4)
, health and Physical,

(Low, Medium
4. Probability

6. Risk Level
safety Ergonomic,
hazards) Radiation,
Psychosocial)

High)
(1-4)
Operation Use of high- Physical Medium Equipment is installed inside a fumehood to
2 4 minimize the risk.
pressure
hydrogen: Check for leaks before startup. First with N2 Yes
then with H2.
Danger of
Periodic checks of tightness during operation.
explosion due to Corrosion checks during down-times
leaks
Damage to the Physical 3 4 High Equipment has a rupture disk
equipment due to Equipment has a pressure switch that stops the
overpressure flow of H2 in case of overpressure
Yes
Pump has an overpressure switch that stops the
liquid flow.
Control system has an interlock that stops the
pump when no H2 is being fed.
Loss of airflow in Physical 1 4 Low Verify the operation of the fumehood
the fumehood can ventilation before startup
cause Add at least one of the operators to the call-list September,
to be warned after-hours if there is a loss in
accumulation of
ventilation 2017
highly
inflammable /
toxic gases
Condensation of Physical 2 3 Medium Keep electrical parts away of the condenser.
water in cold part
of the setup can
cause shortcircuits Yes
if the water gets in
contact with
electrical
components
High temperature Interlocks in the control system stops the
causing power to the heaters.
debilitation of the In critical zones, secondary thermocouple
2 4 Medium Yes
Physical monitors the temperature and shuts down the
components
heating if needed
leading to leaks
and explosion
Production of H2S Produced gas goes through a series of 4
during the scrubbers with household bleach to adsorb the
reaction causes H2S
H2S sensor inside the fumehood shuts down
high
the liquid feed if high concentration is detected
concentrations of 1 4 Low Yes
Chemical Personnel should wear personal H2S detectors
H2S in the while H2S production is suspected.
fumehood and/or Personal H2S detectors should be hanged in
lab space front of the door while not in use. These
detectors will provide a signal of high H2S
concentration before entering the lab
Sampling Sampling at high Lower the sash as much as possible during
pressure causing
exposure to sampling
4 3 High
chemicals and/or Physical Yes
If lowering the sash is not possible, use a facial
physical injuries
shield during the operation

152
1. Task 2. Existing and 3. Hazard 7. Controls (such as immunizations, health 8. Are
Potential Classification screenings, safe work procedures, Controls in
Risk Value
Hazards (Include supervision, training, contingency plans, Place? If
(Include Chemical, emergency plans and kits, PPE, etc.) not, how
environmental Biological, and when?

5. Severity (1-4)
, health and Physical,

(Low, Medium
4. Probability

6. Risk Level
safety Ergonomic,
hazards) Radiation,
Psychosocial)

High)
(1-4)
High temperature Use adequate gloves to operate the valves
in the lines can
transfer to the 4 1 Low
Physical Yes
valves’ handles
causing burns
during sampling
Shut down Feed can be Clean the unit with adequate solvent
corrosive leading
to premature 4 2 Medium immediately after shutting down. Wear an Yes
Physical
failure of
organic vapor mask.
components
Accumulation of Use the adequate PPE while removing the
liquid in the unit
can cause catalyst.
personnel
4 1 Low Be aware of the places where accumulation can Yes
exposure while Chemical
removing the occur. Be prepared to contain / clean any
catalyst
spillage. Wear an organic vapor mask.

Spent catalyst If the catalyst is known to be pyrophoric, the


2 2 operator should keep it in water at all times. Yes
may be Physical Low
pyrophoric

Start up Heating cords can Power to heaters is protected by a fuse


become damaged Heating cords should be inspected before start-
3 3 Yes
and short circuit Physical Medium up both visually and with the help of a
multimeter.
Replace the damaged cords as needed

During the feed Whenever possible, handle all the chemicals


preparation, inside fumehoods or use localized extraction
operators can be (elephant trunks)
exposed to September,
harmful chemicals 4 2 Use the adequate gas mask if the use of
Chemical Medium
fumehoods or extraction is not possible. 2017

Always use containers with lids when


transporting the chemicals outside the
fumehoods

Table B. 1. Hazzard Assessment form use by all University of Alberta Deparments and units.

Completed By:

- Gonzalo Rocha Aguilera (Postdoctoral fellow)


- Anderson Montanez (MSc Student)

153
Department
Chemical and Materials Engineering
Date: (year/month/day)
2017/08/31

154
Appendix C: Copyright

Copyright Authorization Letter

Library
University of Calgary
2500 University Dr. NW, Calgary, AB, T2N 3Y7
Canada
July 19, 2019

This letter hereby grants permission to Anderson Montanez to include in his master thesis entitled:
"Upgrading of HydrofactionTM Biocrude Through Catalytic Hydrotreating," the manuscript
described below. This permission letter is a requirement from the Copyright Office at the
University of Calgary.
Manuscript title: “Hydrotreating of Hydrofaction™ biocrude in the presence of presulfided
commercial catalysts”.
Authors: Parsa Haghighat, Anderson Montanez, Gonzalo Rocha Aguilera, Julie Katerine
Rodriguez Guerrero, Sergios Karatzos, Matthew A. Clarke, and William McCaffrey.
The manuscript was submitted at Sustainable Energy Fuels (Royal Society of Chemistry) on 30th
August 2018, accepted on 14th December 2018 and published on 24th December 2018.
DOI: 10.1039/c8se00439k

As a record, the authors mentioned signed below:

Dr. Parsa Haghighat

Dr. Gonzalo Rocha Aguilera

Dr. Julie Katerine Rodriguez Guerrero

Dr. Sergios Karatzos

Dr. Mathew A. Clarke

Dr. William McCaffrey

155
1

You might also like