111
University of Calgary PRISM: University of Calgary's Digital Repository Graduate Studies The Vault: Electronic Theses and Dissertations 2018-11-13 Assessment of Stage 1 in a Novel Bio-Oil Upgrading Process: Catalytic Hydrotreating Scheele Ferreira, Erika Maria Scheele Ferreira, E. M. (2018). Assessment of Stage 1 in a Novel Bio-Oil Upgrading Process: Catalytic Hydrotreating (Unpublished master's thesis). University of Calgary, Calgary, AB. doi:10.11575/PRISM/34509 http://hdl.handle.net/1880/109182 master thesis University of Calgary graduate students retain copyright ownership and moral rights for their thesis. You may use this material in any way that is permitted by the Copyright Act or through licensing that has been assigned to the document. For uses that are not allowable under copyright legislation or licensing, you are required to seek permission. Downloaded from PRISM: https://prism.ucalgary.ca

Assessment of Stage 1 in a Novel Bio-Oil Upgrading Process

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

University of Calgary

PRISM: University of Calgary's Digital Repository

Graduate Studies The Vault: Electronic Theses and Dissertations

2018-11-13

Assessment of Stage 1 in a Novel Bio-Oil Upgrading

Process: Catalytic Hydrotreating

Scheele Ferreira, Erika Maria

Scheele Ferreira, E. M. (2018). Assessment of Stage 1 in a Novel Bio-Oil Upgrading Process:

Catalytic Hydrotreating (Unpublished master's thesis). University of Calgary, Calgary, AB.

doi:10.11575/PRISM/34509

http://hdl.handle.net/1880/109182

master thesis

University of Calgary graduate students retain copyright ownership and moral rights for their

thesis. You may use this material in any way that is permitted by the Copyright Act or through

licensing that has been assigned to the document. For uses that are not allowable under

copyright legislation or licensing, you are required to seek permission.

Downloaded from PRISM: https://prism.ucalgary.ca

UNIVERSITY OF CALGARY

Assessment of Stage 1 in a Novel Bio-Oil Upgrading Process: Catalytic Hydrotreating

by

Erika Maria Scheele Ferreira

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE

DEGREE OF MASTER OF SCIENCE

GRADUATE PROGRAM IN CHEMICAL ENGINEERING

CALGARY, ALBERTA

NOVEMBER, 2018

Β© Erika Maria Scheele Ferreira 2018

ii

Abstract

The increasing awareness of global warming and depletion of conventional fossil fuel reserves

has motivated the study of alternative fuel sources to fulfill the increasing worldwide demand of

fuels. One promising alternative is the production of fuels using lignocellulose-derived bio-oils

that would not compete with the human food chain. However, this type of bio-oil remains a

challenge due to its high acidic oxygen content that results in corrosiveness and low energy density

compared with crude oils. Therefore, the present MSc. research focuses on the study of the first

stage of a novel catalytic upgrading approach that involves two different hydrogen-addition

processes. First, a mild hydrotreating process is carried out to reduce the oxygen content in the

bio-oil. Then, catalytic steam cracking (CSC), where hydrogen is produced by splitting water

molecules, is used to obtain lighter products from the hydro-treated oil. The main focus of this

research is to optimize the hydrotreating process.

The effect of the process variables such as operating pressure, temperature and space velocity,

on the product quality was evaluated, finding that the best quality hydrotreated product was

obtained at 345 Β°C, 0.2 h-1, 1400 psig, and increasing the temperature beyond 345 Β°C at these

conditions resulted in the appearance of fine solids dispersed in the synthetic product. Additionally,

a comparison of different catalyst formulations was done, finding that acidity is needed in the

catalyst to carry out hydrodeoxygenation reactions. It was also found that the two main compounds

contributing to the acidity of the bio-oil are carboxylic acids and phenols, the latest with a minor

contribution. By hydrotreating, it was possible to achieve a Total Acid Number (TAN) reduction

of 100 % and a maximum of 28 % reduction of the phenols content. A high quality hydrotreated

bio-oil with much reduced oxygen content, low viscosity and higher energy density was produced

in this work.

Keywords: Bio-oil, Total Acid Number, Hydrotreating, Catalytic Upgrading,

Hydrodeoxygenation

iii

Acknowledgments

First of all, I would like to give my most sincere appreciation to Dr. Pedro Pereira Almao for

the opportunity of being part of the Catalysis and Adsorption for Fuels and Energy (CAFE) group

at the University of Calgary. His guidance and advice during my years in Canada, especially, in

the development of this research work are greatly appreciated. I am very lucky to had him as my

supervisor during all these years.

I would like to thank Dr. Monica Bartolini, Mr. Lante Carbognani, Dr. Gerardo Vitale, Dr.

Carlos Scott, Dr. Josefina Scott and Dr. Azfar Hassan for all their insightful technical discussions

and helpful suggestions that guided this research work. Additionally, thanks to all the members of

CAFE group for their support, friendship and all those moments of joy that made this journey more

pleasant, especial thanks to Josune, Marianna, Eduardo, Victor, Christian and Jose Luis.

I would like to acknowledge the Department of Chemical and Petroleum Engineering at the

Schulich School of Engineering for offering me an outstanding formation. The financial support

provided by the Department and Steeper Energy Canada is also greatly appreciated.

I would also like to infinitely thank my parents for all the unconditional support they have

given me during my professional formation and life projects. Thanks for always believing in me,

for encouraging me, being the best examples and provide all the opportunities for me to grow as a

professional and as a person. In addition, I would like to thank my number one fans: my

grandparents Landys, Bertha and Maria. Thanks for the guidance and the advices and for always

cheering me up when I needed encouragement.

Finally, my infinite appreciation and love to my husband Fredy Cabrales. Thanks for being

the best company I could ever had during these past years. Thanks for your patience and

understanding during the hard times and for giving me the motivation and the extra push when I

needed it.

iv

Dedication

To my parents Sergio and Katiuska, my brother Stefan and my grandparents Pito, Tortu and

Nonnita, for all their support and motivation during this journey

To my thoughful and supporting hubby with all my love

v

Table of Contents

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

Acknowledgments.......................................................................................................................... iii

Dedication ...................................................................................................................................... iv

Table of Contents ............................................................................................................................ v

List of Tables ............................................................................................................................... viii

List of Figures and Illustrations ..................................................................................................... ix

List of Symbols, Abbreviations and Nomenclatures .................................................................... xii

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

1.1. Background and motivation ........................................................................................ 1

1.2. Novel Bio-oil Upgrading Approach............................................................................ 3

1.3. Scope of the Research ................................................................................................. 5

Chapter 2: Literature Review .............................................................................................. 6

2.1. Lignocellulosic biomass.............................................................................................. 6

2.2. Thermochemical processing of lignocellulosic biomass .......................................... 10

2.2.1. Hydrothermal Liquefaction ....................................................................................... 11

2.2.1.1. Hydrofaction Process ........................................................................................ 12

2.3. Bio-oil from lignocellulose ....................................................................................... 13

2.3.1. Chemical composition of bio-oil derived from lignocellulose via HTL ................... 14

2.3.2. Important properties of bio-oil .................................................................................. 16

2.4. Bio-oil upgrading ...................................................................................................... 18

2.4.1. Hydrotreating ............................................................................................................ 19

Chapter 3: Experimental Methods .................................................................................... 24

vi

3.1. Bio-oil feedstock ....................................................................................................... 24

3.2. Experimental Set Up ................................................................................................. 24

3.2.1. Feed Section .............................................................................................................. 26

3.2.2. Reaction Section ....................................................................................................... 26

3.2.3. Separation and Sampling Section ............................................................................. 27

3.3. Experimental Procedure ............................................................................................ 28

3.3.1. Reactor Filling .......................................................................................................... 29

3.3.2. Catalyst Activation.................................................................................................... 29

3.3.3. Hydrotreating Operation ........................................................................................... 30

3.4. Characterization Techniques ..................................................................................... 30

3.4.1. Total Acid Number ................................................................................................... 31

3.4.2. Water Content ........................................................................................................... 31

3.4.3. Product Distribution .................................................................................................. 32

3.4.4. Viscosity ................................................................................................................... 33

3.4.5. Thermogravimetric Analysis (TGA)......................................................................... 33

3.4.6. CHN Elemental Analysis .......................................................................................... 34

3.4.7. Microcarbon residue (MCR) ..................................................................................... 34

3.4.8. Fourier-transform Infrared spectroscopy (FTIR) ...................................................... 35

3.4.9. Pre-asphaltenes stability............................................................................................ 37

3.4.10. Gas Analysis ......................................................................................................... 37

Chapter 4: Results and Discussion .................................................................................... 39

4.1. Effect of the total operating pressure ........................................................................ 40

4.2. Temperature and space velocity screening ............................................................... 42

4.2.1. Temperature effect .................................................................................................... 43

vii

4.2.2. Space velocity effect ................................................................................................. 51

4.2.3. Correlation between TAN and Infrared absorptivity at 1710-1750 cm-1 .................. 53

4.2.4. Catalyst Lifetime ....................................................................................................... 56

4.3. Increased severity evaluation .................................................................................... 57

4.3.1. Product distribution of the HDT-bio-oils .................................................................. 62

4.4. Evaluation of a dual-catalyst bed reactor (Reaction #4) ........................................... 67

4.5. Catalyst performance comparison............................................................................. 69

Chapter 5: Conclusions and Future Work ......................................................................... 78

References ......................................................................................................................... 81

Appendix I: Modifications of RTU-1 ........................................................................................... 86

Appendix II: Operational Data and Experimental Results ........................................................... 88

Appendix III. Hydrogen consumption calculation........................................................................ 96

viii

List of Tables

Table 2.1. Typical biomass and waste compositions (wt. % dry mass) adapted from ENC25 ........ 7

Table 2.2. Thermochemical conversion technologies and products, adapted from Bridgewater32

....................................................................................................................................................... 10

Table 2.3. Typical properties of wood derived bio-oil and crude oil ........................................... 14

Table 2.4. Typical operating conditions for hydrotreating bio-oils12, 36, 53 ................................... 20

Table 3.1. Properties of the bio-oil provided by Steeper Energy.................................................. 24

Table 3.2. Relative error for gas chromatography ........................................................................ 38

Table 4.1. Characterization of HDT-bio-oil at 310 Β°C, 0.2 h-1 and different operating pressures 41

Table 4.2. Characterization of HDT-bio-oil at 1400 psig and 315 Β°C using CAT-M3. Cx describes

tested condition evaluated during R2 (See Figure 4.2) ................................................................. 52

Table 4.3. Characterization of HDT-bio-oil at 1400 psig and 320 Β°C using CAT-M3................. 53

Table 4.4. Relative error between the measured and calculated TAN for different samples of HDT-

bio-oil ............................................................................................................................................ 55

Table 4.5. Characterization of HDT-bio-oil at 325 Β°C, 1400 psig, 0.2 h-1 using CAT-M3 in different

reactions (R2 and R3) ................................................................................................................... 58

Table 4.6. Temperature range for the product cuts determined via SimDist or TGA .................. 64

Table 4.7. Characterization of HDT-bio-oil at different temperatures using CAT-M3+ ............. 69

Table 4.8. Characterization of HDT-bio-oil at different temperatures during different reactions 77

ix

List of Figures and Illustrations

Figure 1.1. Proposed novel bio-oil upgrading scheme combining HDT and CSC. ........................ 4

Figure 2.1. Structure of lignocellulosic biomass10 .......................................................................... 7

Figure 2.2. Chemical structure of cellulose14.................................................................................. 8

Figure 2.3. Main components of hemicellulose14 ........................................................................... 8

Figure 2.4. Partial structure of a hardwood lignin molecule from European beech14 .................... 9

Figure 2.5. Phase diagram of water for different operating regimes10.......................................... 12

Figure 2.6. Reaction scheme for the bio-oil formation proposed by Pedersen & Rosendahl42 .... 15

Figure 2.7. Chemical composition of bio-oils according to Milne et al.39 .................................... 16

Figure 2.8. Main reactions occurring in HDT process of bio-oil64, 65 ........................................... 21

Figure 2.9. Reactivity scale of oxygenated groups under hydrotreating conditions15 .................. 22

Figure 3.1. RTU-1 diagram, adapted from Cabrales Navarro69 .................................................... 25

Figure 3.2. Reactor and thermocouple profile probe schematic, adapted from Cabrales Navarro69

....................................................................................................................................................... 27

Figure 3.3. FTIR spectra for bio-oil with the most important bands ............................................ 36

Figure 4.1. H/C ratio and butane/butene ratio vs operating pressure for HDT-bio-oil. H/C ratios

determined over the liquid product; C4/C4= determined over the associated gas phase ............. 41

Figure 4.2. Temperature and space velocity changes during R2. In order to verify the stable

behavior of the catalyst the return of the initial condition was performed twice with the last one

being at the end of the whole test run. .......................................................................................... 43

Figure 4.3. Effect of the temperature on the TAN and viscosity reduction (1400 psig, 0.2 h-1).

Viscosities were determined at 40Β°C ............................................................................................ 44

Figure 4.4. FTIR spectra for the feedstock (bio-oil) and HDT-bio-oil at two temperatures. ....... 46

Figure 4.5. Carboxylic acids reduction vs DOD for HDT-bio-oil at different temperatures. Values

in parenthesis are set up experimental temperatures..................................................................... 47

Figure 4.6. H/C ratio and O/C ratio vs DOD for HDT-bio-oil at 1400 psig, 0.2 h-1 using CAT-M3

....................................................................................................................................................... 48

Figure 4.7. Microscope images at 40 X for the bio-oil feed (left) and HDT-bio-oil at 325 Β°C (right)

....................................................................................................................................................... 48

Figure 4.8. Hydrogen consumption and water yield vs DOD for HDT-bio-oil in R2 .................. 49

x

Figure 4.9. Gas yield distribution for different temperatures in R2.............................................. 50

Figure 4.10. TAN Reduction vs Time on Stream for CSC processing adapted from Trujillo.21 .. 51

Figure 4.11. Graphic given by the TAN equipment for bio-oil feed (left) and HDT-bio-oil 315 Β°C

(right) ............................................................................................................................................ 54

Figure 4.12. FTIR spectra of the bio-oil feed and HDT-bio-oil at 315 Β°C ................................... 54

Figure 4.13. Correlation between transmittance obtained by FTIR for the bands 1710 and 1740

cm-1 between 17 and ln (TAN) ..................................................................................................... 55

Figure 4.14. TAN Conversion of the HDT-bio-oil vs time on stream in R2 using CAT-M3 ...... 56

Figure 4.15. Carboxylic acids and phenols reduction for different temperatures in R3 ............... 58

Figure 4.16. Carboxylic acids and phenols reduction vs DOD in R3 ........................................... 59

Figure 4.17. Microscope images at 40 X for the HDT-bio-oil at different temperatures in R3 ... 60

Figure 4.18. Hydrogen consumption and water yield vs DOD for HDT-bio-oil in R3 ................ 61

Figure 4.19. Gas yield distribution for different temperatures in R3 ............................................ 62

Figure 4.20. Weight percentage of CS2 insoluble material at different temperatures in R2 & R3 63

Figure 4.21. TGA in N2 of the CS2 insoluble material from the HDT-bio-oil at 340 Β°C in R3 ... 65

Figure 4.22. Product distribution and conversion at 343 Β°C+ at different temperatures in R2 & R3

....................................................................................................................................................... 66

Figure 4.23. Product yields vs conversion at 343 Β°C+ for different conditions in R2 & R3 ........ 66

Figure 4.24. Carboxylic acids and phenols reduction for different temperatures using CAT-M3+

....................................................................................................................................................... 67

Figure 4.25. Microscopic images at 40 X for the HDT-bio-oil at different temperatures in R4 .. 68

Figure 4.26. Carboxylic acids and phenols reduction at different T and catalysts ....................... 70

Figure 4.27. Natural logarithm of viscosity vs conversion at 343 Β°C+ for R2, R3, R4 & R5 ...... 72

Figure 4.28. Natural logarithm of viscosity vs conversion at 343 Β°C+ for the same conditions tested

in R2 & R3 .................................................................................................................................... 72

Figure 4.29. MCR vs conversion at 343 Β°C+ for R2, R3, R4 & R5 ............................................. 73

Figure 4.30. Product distribution and conversion at 343 Β°C+ for different T and catalysts ......... 74

Figure 4.31. Comparison of CS2 insoluble material in the liquid product obtained using different

catalysts ......................................................................................................................................... 75

Figure 4.32. Microscope images for the HDT-bio-oil for different T and catalysts ..................... 76

xi

Figure 4.33. H2 consumption vs DOD for different catalysts ....................................................... 77

xii

List of Symbols, Abbreviations and Nomenclatures

Symbol Description Units

ASTM American Society for Testing and Materials

BPV Back Pressure Valve

CAFE Catalysis and Adsorption for Fuels and Energy

CHN Carbon, Hydrogen and Nitrogen

CS2 Carbon disulfide

CSC Catalytic Steam Cracking

CSR Catalytic Steam Reforming

Cx Condition β€œx” tested

DAO Deasphalted Oil

DOD Degree of Deoxygenation %

EU European Union

FID Flame Ionization Detector

FPD Flame Photometric Detector

FTIR Fourier Transform Infrared

GC Gas Chromatography

H/C Hydrogen to Carbon ratio

HDM Hydrodemetallization

HDN Hydrodenitrogenation

HDO Hydrodeoxygenation

HDS Hydrodesulphurization

HDT Hydrotreating

HDT-bio-oil Hydrotreated Bio-oil

HHV High Heating Value MJ/kg

HTL Hydrothermal liquefaction

IBP Initial Boiling Point Β°C

K-F Karl Fischer

KOH Potassium hydroxide

xiii

MCR Micro Carbon Residue

O.D. Outside Diameter

O/C Oxygen to Carbon Ratio

PTV Programmable Temperature Vaporizing

R1 Reaction #1

R2 Reaction #2

R3 Reaction #3

R4 Reaction #4

R5 Reaction #5

RTU-1 Reactivity Test Unit 1

SimDist Simulated Distillation

TAN Total Acid Number mg KOH/g

TCD Thermal Conductivity Detector

TGA Thermogravimetric Analysis

THF Tetrahydrofuran

tv Valves cycle time min

VGO Vacuum Gas Oil

WGS Water Gas Shift

WHSV Weight Hourly Space Velocity h-1

WGM Wet Gas Meter

X343Β°C+ Conversion at 343 Β°C+

XTAN TAN Conversion %

1

Chapter 1: Introduction

1.1. Background and motivation

In the mid-1800s, biomass supplied more than 90% of U.S. energy and fuel demands. But in

the late 1800s to early 1900s, fossil fuels became the preferred energy resource. The discovery of

crude oil helped to industrialize the world and improved living standards by creating an

inexpensive fuel source.1 In the past few years, transport has been almost totally dependent on

petroleum-based fuels such as gasoline, diesel, liquefied petroleum gas and compressed natural

gas; nevertheless, depletion of conventional fossil fuel reserves mainly used for transportation

purposes has motivated the exploration of alternative fuel sources to fulfill the increasing

worldwide demand.2 Additionally, the increasing awareness of global warming have led to strict

regulations for releasing greenhouse gases.3 Usage of biologically derived fuels, may play an

important role for blending with crude oil fractions to supply part of the global demand and to

meet the end-product specifications as they come from a cleaner, CO2 neutral, feedstock.

Investigations in this area are becoming more relevant, as these bio-oils have the advantage of

having reduced contents of contaminants such as sulphur and nitrogen.4 Therefore, they produce

lower amounts and less harmful gas emissions compared to conventional fossil fuels on a life cycle

basis.5

Traditional oil and chemical companies such as Shell, Conoco-Phillips, Dupont and BP are

already transitioning to the carbohydrate economy by developing the technology and infrastructure

for biofuels and biochemicals production.6 Government leaders are also recognizing the

importance of this growing industry by providing tax breaks, grants, incentives and mandates. For

example, in 2006 the U.S. government started giving $0.14/L for ethanol production as a subsidy;

a number of European Union (EU) countries give full tax exemption for biotransportation fuels;

and EU is promoting the growing of crops used for biodiesel and bioethanol production by

providing a carbon credit of $54/ha.6

However, there are some political, economic and technical disadvantages associated with

using biofuels. First, most of the natural material used for producing bio-oils like corn, wheat,

sugar beet and oil seeds, can interfere with the human food chain and can lead to exacerbate current

global food shortage issues.4, 7 Second, because the biofuels industry is starting to grow, some

2

current biomass technologies have low overall thermal conversion efficiencies, making the process

highly expensive and inefficient.6 Finally, bio-oils face some technical challenges regarding some

of their properties such as poor volatility, high viscosity and acidity, thermal instability (gum-

polymers formation) and high content of oxygenated compounds that reduce their miscibility with

petroleum-based fuels.7

The biofuels industry is in its early stage with many novel biomass conversion technologies

being developed to improve overall energy and economic efficiency.8 It is foreseeable that, as

petroleum reserves decline, the price of fossil fuel products will increase and biofuels will

eventually be cost-competitive with petroleum-derived fuels.6 One promising alternative is the

production of biofuels using low-value feedstock, such as waste wood from the pulp and paper

industry, which would not interfere with the human food chain. Additionally, the greatest

advantage of using biological-derived fuels is that, unlike fossil fuels, biomass takes carbon out of

the atmosphere while it is growing, and returns it as it is burned. This maintains a closed carbon

cycle with almost no net increase in atmospheric CO2 levels.9

Steeper Energy Ltd, a Danish-Canadian company, is in the process of commercializing a

hydrothermal liquefaction technology called Hydrofactionβ„’ for production of added-value liquid

fuel from lignocellulosic biomass via water supercritical chemistry.7 This process has been proven

to yield between 45-50 wt.% of liquid product with a low oxygen content when compared with

regular lignocellulose processing technologies.10 The bio-oil produced by Hydrofractionβ„’ still

presents a high acidity, viscosity and oxygen content compared with petroleum-derived fuels.

Hence, an upgrading process is required to convert the bio-oil in bio-fuels or a product miscible

with crude oil.

In the last two decades, literature related to the catalytic removal of oxygen from bio-oil

derived from lignocellulose has been rapidly growing. In a hydrotreating process,

hydrodeoxygenation (HDO) reactions are used to remove oxygen from bio-oils in the form of

water, CO and CO2 by adding H2 to the process.11 Catalytic HDO has been investigated as a

feasible route for the production of fuels from bio-oils. Hydrotreating processes address the

instability of the bio-oil and it is carried out in order to prevent catalyst deactivation in further

processing, minimize coke formation and improve the properties of the oil.6

3

In 2000, the literature of kinetics and reaction networks of HDO was reviewed by Furimsky.12

Four years later, Czernik and Bridgwater investigated developments in the applications of bio-oils

in the industry13 and in 2006 Mohan et al. discussed the process of converting wood into bio-oils

via pyrolysis.14 In 2007, Elliot summarized the historical perspective on developments in catalytic

hydroprocessing of bio-oils.15 Properties and applications of bio-oils produced via pyrolysis or

hydrothermal liquefaction (HTL) has been reviewed by several authors.16, 17 Additionally, different

standard hydrotreating catalysts have been tested for HDO of bio-oils derived from

lignocellulose.18, 19

The early work demonstrated that hydroprocessing of bio-oils was feasible, although not

economical yet due to the severe reaction conditions, hydrogen consumption and the necessity of

further upgrading the hydrotreated product to obtain the commercial products.11 A promising

approach proposed in the present research is to combine hydrotreating with a catalytic steam

cracking (CSC) process, which produces hydrogen, in order to reduce or eventually eliminate

external hydrogen production making the upgrading economically viable.

1.2. Novel Bio-oil Upgrading Approach

The novel bio-oil upgrading approach aims to combine a mild hydrotreating process to mainly

reduce the oxygen content of the feedstock with a catalytic steam cracking process to reach a

deeper conversion of the bio-oil into light fuels. Catalytic steam cracking (CSC) is a process that

uses water as a hydrogen supply by catalytically splitting water molecules while cracking large

and heavy molecules in the feedstock.2 This process is configured in a single catalyst bed and it

uses a dual-function catalyst. The catalyst has a rare earth metal that cleaves the water molecule to

form hydrogen radicals and a hydrogenating metal combined with an acid support to promote

hydrocracking.20 The produced hydrogen radicals are involved in the saturation of hydrocarbon

radicals generated by molecules cracking and act as scavengers to prevent condensation reactions,

thus coke formation.2

The proposed scheme has two main features that look promising for future of upgrading of

bio-oil. First, CSC produces its own hydrogen by cleaving water molecules; hence, there is no

need to feed hydrogen to the CSC process. Second, and most importantly, the unconsumed

hydrogen produced in CSC may be recovered and recycled back to the HDT unit, reducing the

4

fresh hydrogen make-up needed for this unit. According to a study of this process made by Trujillo

in her MSc. thesis,21 the recycle of the unconsumed hydrogen per se from the CSC stage would

meet 8.6% of the hydrogen requirement for the HDT stage. However, the theoretical hydrogen

available from the hydrocarbons gaseous stream from CSC to be recovered considering a catalytic

steam reforming step was calculated and the yield exceeded the hydrogen requirements for the

HDT stage. This means, that with proper treatment for the gas stream produced in CSC, there is

enough hydrogen produced in this process to guarantee all the hydrogen make-up for HDT.

Figure 1.1 shows the proposed bio-oil upgrading diagram via HDT and CSC. First, the HDT

process consists of an up-flow fixed bed reactor containing an in-house formulated catalyst. This

catalyst has been found to be active for hydrogenation reactions in bio-oils by Trujillo in her MSc.

thesis.21 Next, the hydrotreated oil is to be further upgraded via CSC in an up-flow fixed bed

reactor containing a catalyst also assessed by Trujillo. As seen in Figure 1.1, product gases from

both reactors are going to be submitted to catalytic steam reforming (CSR) to process hydrocarbon

gases and recover hydrogen. The hydrogen produced is to be recycled to the HDT process to

minimize or eliminate the make-up hydrogen needed.

Figure 1.1. Proposed novel bio-oil upgrading scheme combining HDT and CSC.

5

In order to optimize the conditions and catalysts for each stage, the processes were evaluated

separately as a first step of the general study. Screening of conditions and catalysts for HDT were

made by Trujillo in her MSc. thesis that produced the starting point for this research. Trujillo also

studied the CSC process including some operating conditions, catalyst types and their deactivation

as well as the hydrogen balance for the whole process.21 The present work in this thesis is going

to have its main focus in optimizing the HDT process in terms of finding the best pressure, catalyst

configuration, temperature and space velocity to produce a high-quality bio-oil that can be

processed by CSC without deactivating the catalyst used in this second processing step (CSC)

while allowing it to fulfill the highest yield of distillate and naphtha products for the integrated

process.

1.3. Scope of the Research

The general objective of this research is to study the novel proposed hydrogenation process

of bio-oil derived from lignocellulosic biomass material provided by Steeper Energy. In order to

accomplish the general goal, some specific objectives were established as follows:

1. Conduct a systematic study of the variables effect such as total operating pressure,

temperature and space velocity on the reactivity of bio-oil via HDT.

2. Evaluate the best conditions and catalysts to minimize the acidity and oxygen content

on the feedstock.

3. Understand the catalyst lifetime during long term evaluation runs.

4. Develop a correlation between two analytical characterization techniques: Total Acid

Number (TAN) and Fourier-transform Infrared (FTIR) to simplify the acid content

analysis.

5. Compare the effect of different catalyst formulations and reactor configurations on the

product quality.

6. Produce a high-quality hydrotreated product for further processing via CSC to prevent

rapid CSC catalyst deactivation due to coke formation.

6

Chapter 2: Literature Review

2.1. Lignocellulosic biomass

Biomass is an abundant renewable source to produce energy efficient fuels such as bioethanol

and bio-diesel in an eco-friendly manner. These types of fuels mainly utilize plants rich in

carbohydrates like sugar cane, wheat, maize, potato, barley, corn or sugar beet as feedstock and

makes up the first generation of bio-fuels.22 The first generation of biofuels is based on well-known

and established technologies, whereas the production of bio-fuels from wood mass is still in the

early stages of research and development and is considered the second-generation bio-fuels.23

Nowadays, a large volume of wood and forest biomass is readily and commercially accessible.

The components of the biomass are obtained from wood harvest and processing residues and

include: tree branches, bark, leaves and limbs, non-merchantable wood, wood pulp wastes and

sawdust.24 Also, biomass from waste wood does not interfere with the human feed chain, which is

one of the main disadvantages of first-generation biofuels.4

Wood-based biomass is essentially a composite material constructed from oxygen-containing

organic polymers and is usually called lignocellulosic biomass. Figure 2.1 shows the main

structure of lignocellulosic biomass. Lignocellulose can be found in the cell walls of plants and

wood and is composed by three major components: cellulose, hemicellulose and lignin. Some

organic extractives such as proteins, resins and waxes and inorganic minerals can be found in

minor concentrations.10, 14 The weight percent of the components varies depending on the wood

species. Table 2.1 shows the typical composition of cellulose, hemicellulose and lignin for

different lignocellulosic materials. As a general trend, it can be observed that the major component

in all different lignocellulosic biomass is cellulose, followed by lignin that in some forest residues

represent the major component.

7

Figure 2.1. Structure of lignocellulosic biomass10

Table 2.1. Typical biomass and waste compositions (wt. % dry mass) adapted from ENC25

Lignocellulosic materials Cellulose Hemicellulose Lignin

Hard woods

Poplar 46.2 24.4 24.5

Birch 40.6 29.6 20.2

Willow 60.5 29.9 25.6

Soft woods

Spruce 44.1 21.2 26.9

Pine 43.6 24.9 25.6

Coniferous wood 57.5 22.5 30.0

Forest residues

Bark, pine 23.7 24.9 50.0

Wood stems 42.6 22.3 37.7

General residues 45.5 21.0 27.3

Other lignocellulosics 38.3 25.2 14.8

Corn stover 37.3 35.8 20.1

Sugarcane bagasse 37.9 26.8 18.3

Wheat straw 37.1 31.2 8.5

Switch grass 46.2 24.4 24.5

8

Cellulose is a high molecular weight linear polymer that consists of D-glucose molecules

bound together by -1,4-glycoside linkages.14 Cellulose fibers comprise between 40-50 % of dry

wood providing the strength of the wood.22 A large portion of cellulose is crystalline and it has a

high tendency to form intermolecular and intramolecular hydrogen bonds.10 In Figure 2.2, the

structure of the cellulose can be observed. The crystalline structure of the cellulose makes it very

resistant to thermal or biological decomposition. However, when exposed at water at supercritical

conditions , cellulose transforms from a crystalline to an amorphous structure allowing cellulose

degradation.26 When cellulose is decomposed by a complete acid hydrolysis, it breaks down to

form glucose.6, 27

Figure 2.2. Chemical structure of cellulose14

Hemicellulose is composed by amorphous and heterogeneous groups of branched

polysaccharides (copolymer of glucose, mannose, galactose, xylose and arabinose) shown in

Figure 2.3. Hemicellulose exhibits a lower average molecular weight than cellulose.14 Cellulose

fibers are surrounded by hemicellulose that acts as a linkage between cellulose and lignin as seen

in Figure 2.1.22 Hemicellulose contains short side-chain branches pending along the main

polymeric chain that makes its decomposition easier. It decomposes at lower temperature (200-

260 C) and forms less chars than cellulose.28 When hemicellulose is decomposed via hydrolysis,

it breaks down to form its 5 monomer sugars (glucose, galactose, mannose, xylose and arabinose).6

Figure 2.3. Main components of hemicellulose14

9

Finally, lignin is a highly complex three-dimensional macromolecule resulting from the

polymerization of different phenylpropane units bound together by ether and carbon-carbon

bonds.22 Figure 2.4 shows a partial structure of a lignin molecule.29 The phenyl propanoid units

that comprised lignin are not linked in a simple, repeating way due to electron delocalization in

the aromatic ring, the double bond-containing chain and the oxygen functionalities.6 Lignin is

markedly different in structure and composition from cellulose and hemicellulose because of its

high aromaticity.11 Thus, it is more difficult to dehydrate than cellulose or hemicellulose and its

maximum rate of decomposition occurs between 350 and 450 C.30 The main products from lignin

decomposition are phenols due to the cleavage of ether and carbon-carbon bonds.14

Figure 2.4. Partial structure of a hardwood lignin molecule from European beech14

10

As mentioned before, cellulose, hemicellulose and lignin interact at the plant cell wall

structural level. Cellulose and hemicellulose adhere to each other due to hydrogen bonding and

van der Waals forces. Additionally, lignin and hemicellulose form ether and ester bonds with each

other.31

In general, lignocellulosic biomass is comprised of carbon (50 wt.%), hydrogen (6 wt.%) and

oxygen (43 wt.%).10 Nitrogen and small traces of chloride account for the remaining 1%. Sulphur

is not present in this type of biomass. The high oxygen content present in this biomass is the main

disadvantage to produce transportation biofuels that are compatible and expectantly competitive

with fossil fuels. Therefore, processing lignocellulosic biomass is needed to decrease the oxygen-

to-carbon (O/C) ratio while increasing the hydrogen-to-carbon (H/C) ratio.

2.2. Thermochemical processing of lignocellulosic biomass

There are three methods of converting biomass into valuable products: gasification, pyrolysis

and liquefaction. Each one of the methods gives different range of products and employs different

equipment and operating conditions. As seen in Table 2.2, gasification is mainly used to produce

synthesis gas and fuel gas; pyrolysis is used to produce liquid fuels or chemicals, charcoal or solid

char and fuel gas; finally, the liquefaction process produces directly bio-oil or liquid fuels.32 Since

the purpose of this work is to upgrade converted biomass into liquid products that can be used as

biofuels, either pyrolysis or liquefaction must be employed.

Table 2.2. Thermochemical conversion technologies and products, adapted from Bridgewater32

Technology Primary Product Application

Gasification Gas Synthesis gas, fuel gas

Pyrolysis

Fast or flash pyrolysis Liquid Liquid fuel substitution, chemicals

Carbonization Charcoal Solid fuel or slurry fuel

Slow pyrolysis Gas, liquid char, solid char Fuel gas, solid fuel, liquid fuel

Liquefaction Liquid Oil or liquid fuel substitution

Combustion Heat Heating

11

Pyrolysis is the process where organics are thermally decomposed to solid, liquid or gas by

heating in absence of oxygen.14 Depending on the operating conditions, solid, liquid or gas

products can be produced. For example, slow pyrolysis produces large amounts of coke that can

be used as solid fuel, whereas fast pyrolysis has proven to maximize the liquid products by using

temperatures of 500 C and very short residence time (less than 1 s).32 Fast pyrolysis has the

advantage of lower capital cost compared with liquefaction processes.6 However, this process

requires a dry biomass, high heating rates and high temperatures and produces a highly-oxygenated

bio-oil because the process does not reduce the oxygen content.14

On the other hand, liquefaction or Hydrothermal Liquefaction (HTL) is considered a

promising technology for bio-oil production because of its high biomass conversion, high bio-oil

yield and low O/C ratio products.33 Additionally, HTL has no limitation to input biomass with high

water content.34

The focus of this research is to upgrade a bio-oil produced with lignocellulosic biomass via

hydrothermal liquefaction. Thus, more details including the operating conditions and catalysts are

given in Section 2.2.1.

2.2.1. Hydrothermal Liquefaction

Hydrothermal liquefaction is a biomass to bio-oil conversion route carried out in water at

moderate temperature between 250 and 400 C and high pressures (up to 30 MPa) with or without

the presence of a catalyst.33 HTL is less developed than fast pyrolysis due to the high cost and

technical difficulties associated with high-pressure processing. Many complex reactions take place

during the transformation of biomass into bio-oil where macromolecular compounds are degraded

into unstable and reactive small molecules that can repolymerize into products with a wide range

of molecular weight distribution.35 The general objective of the process is to control the reaction

rate and reaction mechanisms to minimize the oxygen content of the liquid product and maximize

the yield of the liquid product.32

The presence of different catalysts have been studied by several authors,33, 35 finding that alkali

(alkaline oxides, carbonates and bicarbonates), metals (zinc, copper, iodine, cobalt sulphide, ferric

hydroxide) and Ni and Ru heterogeneous catalysts (which aid preferential hydrogenation) have

been used for liquefaction.

12

Hydrofactionβ„’ is a hydrothermal liquefaction process developed by Steeper Energy that

combines super-critical water chemistry and homogenous catalysts to convert biomass residues to

a high-energy bio-oil.10 More details about this technology will be given in Section 2.2.1.1.

2.2.1.1. Hydrofaction Process

Steeper Energy is commercializing a hydrothermal liquefaction technology called

Hydrofactionβ„’ as a promising path to convert lignocellulosic biomass to bio-oil. This technology

has been proven successfully in a continuous pilot facility.10 Hydrofactionβ„’ includes the use of

supercritical water chemistry, higher pressure and temperatures than other HTL processes reported

in literature.10 The operating conditions are above the critical point of water at pressures between

300-350 bar and temperatures of 390-420 Β°C. Figure 2.5 shows a phase diagram of water to

visualize the different operating regimes.10 In Hydrofactionβ„’, a homogenous catalyst is used in

the form of potassium carbonate (K2CO3) for desired catalytic effects; recirculation of the oil and

aqueous products is also used to improve feed characteristics, energy balance, oil yields and

desired kinetics.10

Figure 2.5. Phase diagram of water for different operating regimes10

The polarity and dielectric constant decrease significantly when water gets closer to its

supercritical state allowing water to dissolve biomass molecules that are hydrophobic at ambient

conditions including phenolics and polyaromatic hydrocarbons derived from lignin.36 Also, at

supercritical conditions, mass and heat transfer rates are enhanced and interphase mass and heat

transfer resistances are significantly diminished.10 Finally, it was proven that water at supercritical

conditions, sustains a high-density at a high-pressure range compared to most HTL processes

operating near the critical point of water.10

13

Jensen, et al. proposed a scheme for the major reactions taking place at Hydrofactionβ„’

conditions that includes: water dissociation, solvolysis, hydrolysis, dehydration, decarboxylation,

steam and CO2 reforming, water gas shift (WGS), aldol condensation and retro aldol, among

others.10 The high-density, alkaline supercritical water promotes depolymerization of

macromolecules through hydrolysis and solvolysis reactions. Some radical reactions may occur as

well due to the high temperatures; however, radical scavengers are used to participate in chain-

terminating reactions.37

Organic solvents and alkaline conditions favor the degradation of the lignocellulose to its

major macromolecules: cellulose, hemicellulose and lignin. First, the cellulose and hemicellulose

depolymerize to oligomers and eventually monomers through hydrolysis and solvolysis. The

oligomers and monomers further dehydrate and isomerize to carboxylic acids, aldehydes and

enols. Depolymerization of lignin can take two different pathways: an ionic pathway where

hydrolysis and solvolysis reactions take place, which is favored because of the conditions of the

process; or a radical pathway through the thermolytical cleavage of both ether and C-C bonds.

From the ionic pathway, low molecular weight phenols are formed.10

The organic compounds contained in the bio-oil resulting from this technology along with

some reaction pathways for cellulose, hemicellulose and lignin are presented in Section 2.3.1.

2.3. Bio-oil from lignocellulose

Bio-oils are physically very similar to crude oil as they are dark brown flowing liquids;

however, they have a very distinctive smoky and acid odor that distinguish them from petroleum-

derived oils.38 Bio-oils are a complex mixture of compounds derived from the depolymerization

of cellulose, hemicellulose and lignin. This complex mixture include water, solid particles and

hundreds of organic compounds such as acids, alcohols, ketones, aldehydes, phenols and ethers,

among others.39 Some of these compounds are directly related to the undesired properties of bio-

oil like high acidity, oxygen content, viscosity, low heating value and instability.

When comparing the properties of bio-oil and crude oil a significant difference is noticed.

Table 2.3 presents the typical properties of a bio-oil and a crude oil.6, 16 It can be seen that the two

properties that differ the most between a bio-oil and crude oil are the moisture content and the

elemental composition, where it can be observed that bio-oils have higher oxygen content than

14

crude oils. However, a low content of contaminants such as nitrogen9 and sulphur has been found

in bio-oils derived from lignocellulose.6

Table 2.3. Typical properties of wood derived bio-oil and crude oil

Bio-oil6, 16 Alberta Bitumen40

Moisture content [wt. %] 1-30 <1

Elemental Composition [wt. %]

C 65-75 82-83

H 5-8 10-11

O 10-40 <1

N <0.5 <1

S <0.05 4.5-6.0

High heating value (HHV) [MJ/kg] 20-30 40

Viscosity at 40 Β°C [cP] 6,000-30,000 12,000

2.3.1. Chemical composition of bio-oil derived from lignocellulose via HTL

The chemical composition of bio-oils may vary depending on different factors, such as

biomass type, feedstock composition, feedstock pretreatment, process for converting biomass and

operating conditions of the process.6 In general, bio-oils are a blend of more than 400 important

organic compounds at different compositions.41 Oxygenated aromatics, heterocyclic compounds

and long chain aliphatic backbones can be found on this renewable oil.42

Carrier et al. investigated the conversion of hemicellulose, cellulose and lignin at supercritical

water conditions and found that products can be grouped into two main pools: oxygenated and

substituted 5-membered ring structures, such as ketonic cyclopentanes and cyclopentanes; and

oxygenated and substituted aromatics.43, 44 Quitain et al. performed a qualitative evaluation on

hydrothermal treatment of a type of bark and identified furfural, benzenes, phenols and acids such

as stearic and palmitic as the main compounds found in the produced bio-oil.27

The reaction mechanism to produce bio-oil is complex and consists of multiple chemical

reactions. It was found that cellulose and hemicellulose (carbohydrates) present similarities in

15

terms of yield, composition and chemical mechanism,42 which reduces the complexity of the

mechanism. Figure 2.6 shows a proposed reaction scheme for the formation of bio-oil.42

Figure 2.6. Reaction scheme for the bio-oil formation proposed by Pedersen & Rosendahl42

Quitain et al.27 found that carbohydrates mainly yield oxygenated 5-membered ring structures

such as furfural and 5-hydroxymethyl furfural whereas lignin yields oxygenated aromatic

compounds such as catechol, phenols and cresols.42, 45 The composition of the different elements

was found to be directly related with the content of lignin, cellulose and hemicellulose in the

biomass. Feedstocks with higher content of lignin yielded more content of aromatics than those

with more cellulose or hemicellulose.42 Milne et al. summarized the chemical composition of bio-

oils derived from lignocellulosic biomass and it is presented in Figure 2.7. It can be observed that

bio-oil contains a numerous variety of compounds such as acids, esters, ketones, aldehydes, sugars,

miscellaneous oxygenates, furans, phenols, guaiacols and syringols.6, 39. In this Figure, the black

column corresponds to the minimum composition found in bio-oils of this compound while the

16

gray column corresponds to the maximum composition found in lignocellulosic-derived bio-oils

of the same compound.

Figure 2.7. Chemical composition of bio-oils according to Milne et al.39

2.3.2. Important properties of bio-oil

The main physicochemical properties resulting from the chemical composition of bio-oils will

be discussed in this section. By following the changes of these properties, it can be determined if

a bio-oil was successfully upgraded to be use as a petroleum-derived fuel.

Water in bio-oils result from the original moisture of the feedstock and from dehydration

reactions during biomass processing. Water content can vary from 15 to 30 % and although water

reduces the viscosity of the oil and enhances the fluidity, it is hard to remove from bio-oils. Its

presence lowers the heating value and flame temperature, reducing the combustion rates of the

oil.16, 46 Bio-oil produced via Hydrofactionβ„’ has only 1-3 % of water which is another advantage

of this process.10

Due to its chemical composition, bio-oils usually have a pH of 2-4 and a total acid number of

50-100 mgKOH/g.38 As mentioned before, they comprise a substantial amount of carboxylic acids

in the form of acetic and formic acids that leads to a high level of acidity. For this reason, bio-oils

17

are corrosive to common construction materials such as carbon steel and aluminum.47 The

corrosiveness is extremely severe at high temperatures, which imposes more requirements on

construction materials and operating conditions for the upgrading process before using bio-oil as

transportation fuels.38

The oxygen content of bio-oils may vary between 10-40 %,46 distributed in more than 300

identified organic compounds. These oxygenated compounds make bio-oils polar, and therefore

immiscible with non-polar petroleum fuels. The presence of oxygen leads to a low heating value,

corrosiveness and instability.13, 16 Also, polymerization of oxygenated compounds in the form of

phenols has been reported.6 One of the primary reasons for differences in the properties and

behavior between hydrocarbon fuels and bio-oils is the high oxygen content. As seen in Table 2.3,

oxygen content for petroleum-derived hydrocarbons is between 10-40 times lower than for bio-

oils.

Viscosity plays an important role in the design and operation of the fuel injection because it

is a measure of the fluid resistance to shearing forces.16 The viscosity for bio-oils can vary between

6000-40000 cP at 40Β°C, depending on the feedstock and processing of the biomass. Also, the

chemical structure of the bio-oil may be related to this property. Studies have found that alcohols,

acid groups and intermolecular interactions have a strong effect on viscosity; hydrogenated

compounds are more viscous than aromatic compounds and branched hydrocarbons have lower

viscosities than straight chains.48, 49

The heating value is the amount of heat produced by a complete combustion of fuel and it is

measured as a unit of energy per unit mass or volume of substance.50 It is a quantitative

representation of the energy content of an oil because it dictates the amount of energy produced

for each volume of burned fuel. Usually bio-oils produced from plants have a higher heating value

than those produced from straw, wood or agricultural residues. The heating value of a bio-oil (20-

30 MJ/kg) is lower than the one of crude oil (40 MJ.kg). This could be related to the high amount

of oxygenated compounds found in bio-oils, since studies have found that the heating value is

proportional to the elemental composition of an oil being negatively affect by the oxygen content.9

These undesired properties have limited the range of bio-oil applications. They cannot be

directly used as transportation fuels due to bio-oils high viscosity, acidity, oxygen content and low

18

heating value. Therefore, upgrading of bio-oil is needed to improve its properties for liquid fuel,

starting with the removal of the oxygen content that will affect directly the other properties

mentioned above.

2.4. Bio-oil upgrading

In order to unlock the potential commercialization of bio-oils, upgrading of the converted

biomass is needed. Properties that negatively discern the quality of bio-oil from crude oil such as

high viscosity, acidity and high oxygen content can be improved by different upgrading routes.

The three different routes described for upgrading bio-oil to liquid transportation fuels are:

hydrotreating, hydrocracking and emulsification.6

Hydrotreating (HDT) is a simple hydrogenation process that is used to improve the product

quality without significantly altering the boiling range of an oil.16 This process is the most

commonly applied because it reduces the oxygen content of the bio-oil while increasing the H/C

ratio of heavy molecules.51 In general, depending on the targeted molecules, reactions can be

classified as hydrodesulphurization (HDS), hydrodenitrogenation (HDN), hydrodemetallization

(HDM) or hydrodeoxygenation (HDO).12 For bio-oil, the main reaction taking place is HDO

because, in contrast with crude oil, it does not have a significant amount of sulphur, nitrogen or

metals for the other reactions to take place. One of the advantages of HDO is that during the

process, oxygen in the feed is mainly converted to water, which is environmentally friendly.52

Hydrotreating involves processing bio-oil at moderate temperatures to avoid coke formation.53

It serves as a pre-treatment step to hydrogenate unsaturated hydrocarbons and remove oxygen from

the feedstock. Hence, further upgrading is needed to have a high-quality oil.

Hydrocracking (HDC) is a high-temperature process (>350 ΒΊC) where hydrogenation

accompanies cracking to produce a large amount of light product while increasing the H/C ratio

of the feedstock.6 The products from this reaction include hydrocarbons, water-soluble organics,

oil-soluble organics, gases and coke. The wide range of products is the result of combining

catalytic cracking reactions with hydrogenation reactions.6 A dual-function catalyst containing a

cracking function (silica-alumina or zeolite) and a hydrogenating function (Pt, W and Ni) is used

for catalyzing the reactions.54 Although HDC combines hydrogenation with further upgrading of

19

the feedstock, the high costs due to the severe conditions required such as high temperature and

high hydrogen pressure to deal with acids makes this route not as common as hydrotreating.54

Finally, one of the simplest methods for using bio-oil as a transportation fuel is

emulsification. This process has been investigated by many researchers55-58 and consists of

blending bio-oils with diesel using surfactants.6 Overall, upgrading bio-oil through emulsification

provides a short-term approach to the use of this type of oil in diesel engines due to the promising

ignition characteristics showed by the emulsion. However, most fuels properties like heating value,

cetane number and acidity did not meet the requirements which is why other alternatives such as

HDT and HDC are being favored.6, 54

As mentioned before, a better alternative for upgrading the bio-oil is to combine

hydrogenation with cracking reactions in order to first pre-treat the feedstock by increasing the

H/C ratio and decreasing the O/C ratio; and then reach a deeper conversion with a cracking process.

The main disadvantage of this approach is the high amount of hydrogen needed for processing the

bio-oil in a regular hydrotreating-hydrocracking configuration. Nevertheless, the novel bio-oil

upgrading scheme proposed in this research that combines HDT with CSC, where hydrogen can

be produced and recycled, could be a promising upgrading approach.

2.4.1. Hydrotreating

A main goal of upgrading bio-oil is to convert the oxygen-rich, high-molecular-weight

compounds into hydrocarbons that are compatible with petroleum-derived fuels.11 A potentially

valuable process for pre-treating the feedstock is hydrotreating or hydrodeoxygenation, which has

been proven to significantly improve the quality of bio-oils in terms of oxygen content, viscosity,

acidity and stability.12 Without the HDO step, direct high-temperature catalytic processing, needed

to obtain the commercial products like gasoline and diesel, resulted in high levels of coke

production that plugged the catalyst bed.53

For HDT reactions to take place, the presence of hydrogen and a catalyst with a hydrogenating

function is needed. Conventional hydroprocessing catalysts, such as CoMo and NiMo supported

in alumina were useful for HDO in the sulphided form.59, 60 However, the alumina supports were

found to be instable in the presence of high levels of water. Also, a significant amount of coke was

observed when using alumina as the catalyst support.53 Other catalysts, containing Pt, Ni, Pd or

20

other metallic group, are currently being tested for this type of feed. These catalysts were assessed

to be more active at lower temperatures than the sulphided molybdenum-based ones. Metallic

phases can be easily supported on non-alumina supports like carbon or titania to avoid the water

instability of alumina. The main concern for the metallic catalysts is the high cost associated with

most hydrogenating metals like Pt or Pd.53

Regarding the operating conditions for hydrotreating bio-oils, Table 2.4 summarizes typical

conditions found in the literature. Generally, the temperature for HDT is in the low range to remove

oxygen primarily in the form of water, without severely reducing the chain length of the molecules

in the feed. Also, high-temperatures when treating bio-oils promote coke formation resulting from

the original oxygenated compounds.53 High pressure range, as seen in Table 2.4, is generally used

for HDT because hydroprocessing catalysts usually require high pressures to enable H2 and

reagents to reach all the active sites of the catalyst and perform the hydrogenation reactions.15, 61

Additionally, high pressures ensure a higher solubility of hydrogen in the oil, thus a higher

availability of hydrogen in the catalyst surrounding area. By favoring hydrogenation, the reaction

rate increases and the coke formation in the reactor decreases.62

Table 2.4. Typical operating conditions for hydrotreating bio-oils12, 36, 53

Parameter Common values

Temperature [Β°C] 250-400

Pressure [MPa] 3-18

Liquid hourly space velocity [h-1] 0.1-0.8

H2 feed rate [L H2 STP/ L oil] 100-700

Some of the O-compounds in the feed tend to polymerize from undesirable reactions between

aldehydes and organic acids. This leads to an increase of the molecular weight and is the main

cause for bio-fuels instability.12 Nevertheless, studies have proven that HDT is an effective way to

convert aldehydes and unsaturated compounds into more stable molecules by removing oxygen

atoms.63 The main reactions expected to take place during the HDO of bio-oils are presented in

Figure 2.8. Additionally, undesired reactions such as reverse water gas shift, methanation and coke

formation are expected to occur.64 Hydro-decarboxylation and hydro-decarbonylation remove

21

oxygen in the form of carbon dioxide and carbon monoxide, respectively. Hydro-deoxygenation

removes oxygen in the form of water without cleaving the molecules chain length.65

Figure 2.8. Main reactions occurring in HDT process of bio-oil64, 65

As reported by Milne et al.,39 bio-oil comprise many functional groups that are expected to

react at different temperatures. Grange et al. studied the activation energies and the reactivity

temperatures of different compounds found in bio-oils, finding that molecules with a bound or

sterically hindered oxygen (furans or ortho substituted phenols) required a significantly high

temperature for the reaction of hydrodeoxygenation to take place.66 Furimsky12 summarized the

apparent reactivity for different compounds as:

alcohol > ketone > alkylether > carboxylic acid β‰ˆ M-/p-phenol β‰ˆ naphtol > phenol > diarylether β‰ˆ

O-phenol β‰ˆ alkylfuran > benzofuran > dibenzofuran

A study made by Weisser et al. is in agreement with Furimsky’s reactivity proposal67. In

Figure 2.9 it can be observed that at low temperatures (<200 Β°C), olefins, aldehydes and ketones

are the components reduced by hydrogen. Removing these components have a positive impact on

the stability of the bio-oil.15 Alcohols are reacted at 250-300 Β°C by catalytic hydrogenation but

also by thermal dehydration to form olefins. Carboxylic and phenolic ethers are reacted at 300 Β°C

while phenols and dibenzofurans need temperatures higher than 350 Β°C to react with hydrogen.

3 +

22

Figure 2.9. Reactivity scale of oxygenated groups under hydrotreating conditions15

Finally, Elliot et al. studied the effect of temperature for HDO of wood-based oil using a Pd/C

catalyst in a fixed bed reactor. The operating pressure was 14 MPa and the temperature range was

between 310-340 Β°C. It was found that above 340 Β°C the degree of deoxygenation (DOD) did not

increase further, but instead extensive cracking took place accompanied by a decrease in the oil

yield.68

Although HDT is considered a very effective technology to process and improve the

properties of bio-oil, it is important to consider the amount of hydrogen needed to achieve high

HDO conditions and its impact on the profitability of the process. Venderbosch et al. investigated

the hydrogen consumption for bio-oil upgrading as a function of the DOD, finding that the

hydrogen consumption increases sharply when the DOD reaches more than 50%.61 This could be

related with the reactivity of different compounds, e.g. highly reactive oxygenates like ketones can

be easily converted with low hydrogen consumption because oxygen is available for reaction,

whereas complex molecules like furans, need to be hydrogenated/saturated first which increases

the hydrogen consumption notably.12

23

The product after HDT is usually a bio-oil with a reduced oxygen content, viscosity and

acidity. Nevertheless, it still comprises non-polar high-molecular-weight organic compounds51

that, in order to obtain commercial products, require further processing of the oil in a hydrogen

rich environment. To allow further processing without consuming more hydrogen in a

hydrocracking process, a new alternative is proposed. Catalytic Steam Cracking (CSC) is a

moderate-conversion process that produces hydrogen through steam dissociation and cracks heavy

molecules both thermally and catalytically.20 The unconsumed hydrogen in CSC can be recycle to

the HDT process in order to make bio-oil upgrading more economically viable.

24

Chapter 3: Experimental Methods

3.1. Bio-oil feedstock

The experiments performed in this research project were done using a bio-oil feedstock

provided by Steeper Energy. This feedstock is produced using lignocellulosic biomass via a

patented process named Hydrofractionβ„’, a supercritical hydrothermal liquefaction technology

explained in detail in Section 2.2.1.1. Properties of the feedstock used in this research are presented

in Table 3.1.

Table 3.1. Properties of the bio-oil provided by Steeper Energy

Property Value

Viscosity @40 Β°C [cP] 31172

TAN [mg KOH/g] 48.32

Microcarbon [wt. %] 22.01

Oxygen content [wt. %] 10.82

H/C molar ratio 1.36

O/C molar ratio 0.10

Water content [wt %] 1.08

Distillation Cuts [wt. %]

Naphtha (IBP - 190 Β°C) 2.1

Jet Fuel (190 - 260 Β°C) 5.2

Diesel (260 - 343 Β°C) 10.5

VGO (343 - 545 Β°C) 26.5

Residue (545 Β°C +) 55.8

3.2. Experimental Set Up

The reactivity tests for upgrading the bio-oil in this research were carried out in a Reactivity

Test Unit (RTU-1) bench-scale pilot plant designed and constructed by Cabrales Navarro.69 RTU-

1 is equipped with an up-flow tubular reactor that can be used to emulate the performance of

industrial processes such as hydrotreating, thermal cracking or visbreaking. Catalytic Steam

25

Cracking (CSC) and thermal cracking reactions of De-Asphalted Oil (DAO) were performed in

this unit by Cabrales Navarro69 prior to the beginning of this research. Also, a detailed description

of the design, construction and operation of RTU-1 is reported.69

The unit can be divided in three main sections: Feed section, Reaction section and Separation

and Sampling section. For the purposes of this thesis, the Separation and Sampling section was

modified to accommodate the unit for the bio-oil feedstock and hydrotreating conditions. A

summary of the modifications is found in Appendix I.

A whole schematic of RTU-1 including the modifications done to the unit is presented in

Figure 3.1.

Figure 3.1. RTU-1 diagram, adapted from Cabrales Navarro69

26

3.2.1. Feed Section

The feed section is equipped with two steel tanks to supply feedstock to the pump. The main

tank is a 10 L custom made stainless steel vessel of 6.7” and 15” height, built in the Engineering

Machine Shop at the University of Calgary. This tank is equipped with a spring-type relief valve

that opens at 100 psig in case the vessel over pressurizes. The main feed tank is heated between

80-100 Β°C to ensure mobility of the feedstock and it is pressurized up to 100 psig for

homogenization purposes and to provide head pressure to refill the pumps. The second feed tank

is an auxiliary 1 L Swagelok vessel operated at room temperature where vacuum gas oil (VGO) or

dichloromethane is stored for cleaning purposes. A Teledyne ISCO series 500D dual-pump

continuous flow system with dual pneumatic valves and controlled by a Series D Controller is used

to pump the feedstock to the system. The continuous flow mode allows refilling one pump while

the other one delivers fluid to the system. In case there is a reactor or lines plugging downstream,

a spring-type relief valve is placed in the pump outlet line that directs the feed flow to an auxiliary

500 mL Swagelok tank depending on the set pressure. The feed section is also equipped with a

heated water and gas outlet line (TC-201) for CSC processing. In this thesis, only hydrotreating

reactions were performed, thus only hydrogen was injected through this line. A Brooks Instrument

5850 EM hydrogen mass flow controller is installed for hydrogen injection. Lines made of ¼’’

O.D. 316 stainless steel tubing provided by Swagelok connect all the parts of the feedstock

pumping. Additionally, heating tapes are used for heating the lines at temperatures up to 140 Β°C

(TC-101 to TC-108) due to limitations in temperature of the pneumatic valves of the ISCO pumps.

Finally, every heated piece is insulated with Superwool insulation (Ref. 6# SW 607 supplied by

Improheat-Edmonton) to reduce heat losses.

3.2.2. Reaction Section

RTU-1 is equipped with a tubular reactor operated in up-flow mode. In this case, the reactor

was operated as a fixed-bed, filled with different types of supported catalyst for different test runs.

Volume of the reactor for most of the test runs was 29.1 mL. The reactor was assembled with 35.5

cm length 316 stainless steel Swagelok tubing, ½” O.D. and 0.049’’ wall thickness. An Omega

thermocouple with 7 sensing points is installed inside the reactor for temperature monitoring as

27

presented in Figure 3.2. As seen in Figure 3.2, 6 sensing points are distributed inside the reaction

zone and the other point indicates the temperature before the inlet of the reactor.

Figure 3.2. Reactor and thermocouple profile probe schematic, adapted from Cabrales Navarro69

To heat the reaction section, three individually controlled heating tapes (TC-204, TC-205 &

TC-206) are wrapped around the reactor to have versatility to adjust any of the sections output

independently to obtain a homogenous temperature profile. There is a ¼” O.D. pre-heating line

before the reactor entrance in order to increase the temperature of the feed and reduce the heat load

at the reactor entrance.

3.2.3. Separation and Sampling Section

The hot separation system is equipped with the following double ended 304L stainless tanks

supplied by Swagelok: one 1 gallon stability tank (not used in this research), two 1 L mass balance

vessels and one 40 mL intermediate tank.

The mixture of gases, water and liquid hydrocarbons exiting the reactor go to a collector tank

(MB Tank 1) operated at low temperature (90 Β°C) because of restrictions of the tank at high

28

pressure (1400 psig). In this collector tank, gases are separated from the liquid products and passed

through a back pressure valve (BPV) that maintains the operating pressure at the given set point.

Then, the gas stream is sent to the gas release and depressurization section. This section consists

of two KOH traps for gas sweeting in case of having hydrogen sulphide as a product of the reaction,

a Gas Chromatograph (GC) for gas analysis and a Shinagawa W-NK-0.5-18 Wet Gas Meter

(WGM) for gas flow measurements.

The first tank (MB Tank 1), as mentioned before, is used to separate the gases and collect the

liquid product. The second tank (MB Tank 2) works as a hot separator at 110 Β°C and atmospheric

pressure to ensure water separation. To collect mass balances and separate the water from the

liquid product, the last one must pass from MB Tank 1 to the hot separator (MB Tank 2) without

a high drop in the pressure of the system. For this purpose, an automated sampling system equipped

with two computer-controlled pneumatic valves are set to control the valve between MB Tank 1

and the 40 mL vessel (V-301) and the valve between the 40 mL vessel and MB Tank 2 (V-302).

The pneumatic valves are timed in such a way that V-301 opens and approximately 90% of the

small vessel is filled with liquid. This causes a small pressure drop in the unit, less than 3% of the

operating pressure. V-301 is left open for 300 s and then it closes automatically. After 10 s, V-302

opens 300 s and the product is released to the hot separator. After this point, the cycle set in the

computer starts again. To ensure water separation, a constant amount of nitrogen is injected at the

bottom of the hot separator and the residence time should not be less than one hour before

collecting the sample. Water and light products that distill at 110 Β°C are sent to a 304L Swagelok

stainless steel 75 mL mass balance tank (MB Tank 3) operated at room temperature and

atmospheric pressure. Nitrogen is used to flush the samples from MB Tank 2 and MB Tank 3.

3.3. Experimental Procedure

RTU-1 was used to test different catalysts and conditions through this thesis. Prior to the start-

up of the unit, it is necessary to fill the reactor with the catalyst and to treat the catalyst to activate

the metals on it. These two steps will be explained in section.3.3.1 and 3.3.2. Also, the operation

and start-up of RTU-1 for hydrotreating will be discussed in section 3.3.3.

29

3.3.1. Reactor Filling

Figure 3.2 shows a schematic of the packed bed reactor used on the experiments. First, the

reactor inlet fittings, containing the thermocouple, were closed and attached to the empty tubing

(reactor). Next, quartz wool was introduced through the tube exit end to reach the inlet of reactor.

Using a funnel, carborundum previously washed with hydrogen chloride was added until reaching

the isothermal zone (point #2 of the thermocouple). To separate the carborundum from the catalyst,

more quartz wool was incorporated to the tube. Then, catalyst was loaded until reaching point #7

applying vibration to ensure a well-packing. Afterward, more quartz wool, carborundum and

quartz wool again, were incorporated to the reactor until reaching the outlet.

The filled reactor was assembled into RTU-1 and leak test was performed at 1650 psig of

Nitrogen for at least 24 hours. Leaks were detected using Snoop Liquid Leak Detector from

Swagelok. The maximum acceptable pressure drop per hour was 0.5%.

The amount of catalyst loaded in the reactor and the Weight Hourly Space Velocity (WHSV)

selected for each experiment allowed the determination of the oil mass flowrate to be used. It can

be determined following Eq. 3.1.

π‘Šπ»π‘†π‘‰ [β„Žβˆ’1] = 𝑂𝑖𝑙 π‘šπ‘Žπ‘ π‘  π‘“π‘™π‘œπ‘€π‘Ÿπ‘Žπ‘‘π‘’ [

π‘”β„Ž

]

π‘€π‘Žπ‘ π‘  π‘œπ‘“ π‘π‘Žπ‘‘π‘Žπ‘™π‘¦π‘ π‘‘ [𝑔]

Eq. 3.1

3.3.2. Catalyst Activation

The oxidation states of the metals added to the catalyst at the start of the test run are very

important to obtain maximum performance. In order to reach the desired oxidation state for the

catalysts tested in this research, a reduction under nitrogen and hydrogen was required for

activation of the catalyst. Previous studies done by Vitale, et. al70 showed that all the metals used

in the catalysts were reduced at 500 Β°C. To start the reduction, nitrogen was flowed through the

reactor filled with catalyst at a rate of 60 mL/min. The temperature was ramped to 500 Β°C at a rate

of 10 Β°C/min. After reaching the set point, external temperatures were adjusted to ensure a

homogenous profile. The temperature was maintained at 500 Β°C for 6 hours and the nitrogen flow

was set to zero after reaching room temperature. The same procedure was repeated using hydrogen

instead of nitrogen. However, after reaching the set point (500 Β°C), the temperature was maintained

30

for 8 hours instead of 6 hours. Hydrogen flow was set to 10 mL/min after reaching room

temperature to keep the unit under a hydrogen environment for the start-up. The catalyst activation

was done at atmospheric pressure.

3.3.3. Hydrotreating Operation

In order to bring all the process variables to reaction conditions in a smooth manner after the

catalyst treatment, there is a procedure that needs to be followed. First, the temperatures in the feed

and separation and sampling section were increased to the regular operation set-point (between

90 Β°C and 110 Β°C). Next, the pressure of the system was increased and hydrogen flow was set to

reaction conditions. The back pressure valve (BPV) was manually closed to constrain the gas flow

at the exit of the system until reaching the set-point value. Then, the reactor temperatures were

increased at a rate of 10 Β°C/min until reaching the set-point. External temperatures were adjusted

to obtain a homogeneous temperature profile. Afterward, oil was flowed through the system at a

rate of 2 mL/min for 60 minutes to fill the lines leading to the reactor, and the reactor as well.

Finally, oil mass flow rate was fixed to the set-point determined using Eq. 3.1 and the time to reach

stability for the reaction was started when the internal temperatures achieved the reaction set-point.

Each condition tested was considered stable after oil passed through the reactor three times its

volume at the corresponding set-points of temperature, pressure and oil and hydrogen flowrate.

As explained in Section 3.2.3, an automated sampling system was used to transfer the liquid

product to the separation tank. The valves cycle time (tv) in the automated system was calculated

from Eq. 3.2, where 40 mL is the volume of the vessel between MB Tank 1 and MB Tank 2.

𝑑𝑣 [π‘šπ‘–π‘›] = 40 π‘šπΏ

𝑂𝑖𝑙 πΉπ‘™π‘œπ‘€ π‘Ÿπ‘Žπ‘‘π‘’ [π‘šπΏπ‘šπ‘–π‘›]

Eq. 3.2

3.4. Characterization Techniques

The liquid and gas products obtained from hydrotreating the bio-oil were analyzed to

understand the effect of varying different operation conditions. Most of the techniques described

in this section were modified from the ASTM norms used for heavy oil for a feasible

implementation with the resources available at the CAFE group at the University of Calgary. Also,

some modifications due to the nature of the feedstock (bio-oil) were done.

31

3.4.1. Total Acid Number

Total Acid Number (TAN) is defined as the naphthenic acid content on a crude oil and it is

commonly expressed as the milligrams of potassium hydroxide (KOH) needed to neutralize a gram

of crude oil.71 TAN number or the acidity of the samples was measured following the ASTM D644

norm.72 Although this method is well-established and accepted, with the available Mettler T70

titrator this method has disadvantages. It does not differentiate strong acids from weak acids, thus

it does not distinguish the type of molecules in the sample such as naphthenic acids, phenols,

mercaptans or other acidic components present in the sample.71 From the preceding drawbacks,

this technique was coupled with other characterization methods in order to obtain more

information about the samples.

A Mettler Toledo T70 Titration Excellence was used for measuring the acidity of the samples

using a titrant solution of 0.05 M KOH in 2-propanol and water. In this method, between 0.5 and

0.7 g of sample is diluted with 60 ml of solvent composed by 50 % toluene, 49.5 % 2-propanol

and 0.5 % of water. The vessel with the solution was placed in the auto-sampler tray and the

electrode, titrant dispenser and mixer were placed inside the vessel. The neutralization reaction

was monitored by potentiometry until reaching completion. Total Acid Number was reported by

the equipment depending on the amount of titrant used, its respective concentration and the amount

of sample used. Finally, Eq. 3.3 was used to calculate TAN conversion (XTAN). The relative error

for the TAN measurement is 3% for TAN>50 mgKOH/g, 10% for TAN ranging between 1-

5 mgKOH/g and 20% for TAN<1 mgKOH/g.

𝑋𝑇𝐴𝑁[%] = (1 βˆ’π‘‡π΄π‘ π‘œπ‘“ π‘ π‘Žπ‘šπ‘π‘™π‘’ [

π‘šπ‘”πΎπ‘‚π»π‘” ]

𝑇𝐴𝑁 π‘œπ‘“ 𝑓𝑒𝑒𝑑 [π‘šπ‘”πΎπ‘‚π»

𝑔 ]) βˆ— 100 Eq. 3.3

3.4.2. Water Content

In order to corroborate the correct separation of the water from the hydrocarbon product and

to quantify the water content on the heavy product for further calculations, the water content in the

bio-oil sample was determined by coulometric Karl-Fischer (K-F) following the procedure

described by Carbognani, et. al.73 This procedure is a modification of the ASTM D4928 method

where tetrahydrofuran (THF) is used as a solvent for homogenization purposes.74

32

In this method, approximately 0.5 g of sample diluted in 10 mL of THF were agitated until

the bio-oil was completely dissolved. A known mass of this solution was injected into the Karl-

Fischer Mettler Toledo Model DL-32. Later on, the water content was determined from the current

generated by titration of the sample with the K-F reagent applying the calibration for the

equipment.

3.4.3. Product Distribution

Product Distribution for oils provides the quantity of the weight fractions that can be distilled

at different temperatures. This is directly related to the economic value of the oil because the

fractions that can be distilled at lower temperature are easier to process and transform into valuable

products. This property is very important when defining the upgrading scheme required to process

the feedstock.75

Simulated Distillation (SimDist) was used to obtain the liquid product distribution following

the ASTM D-7169-05 norm76 with an in-house modification by Carbognani et al.77 In this method,

1 ΞΌL of a solution prepared with 150 mg of sample diluted in 20 mL of CS2 and previously filtered

with a 0.45 ΞΌm membrane was injected in an Agilent 6890N chromatograph instead of 0.2 ΞΌL as

established in the ASTM norm. The chromatograph is equipped with an automatic injector, a PTV

injection port and a 5 m x 0.53 mm metallic capillary column with a 0.1 ΞΌm film methyl silicone

stationary phase (Ref. P/N SS 112-102-01 from Separation Systems Inc). The in-house norm

modification was made to reduce from 25% to 5% the volumetric error from the injection of the

sample. Also, to diminish the volumetric error due to the potential presence of nanoparticles. For

example, the presence of 100 nm particles can result in a 20% error when 0.2 ΞΌL of sample are

injected.77 Using SimDist, the liquid distribution of an oil can be determined. Also, the conversion

for a product at 343Β°C+ can be determined using Eq. 3.4 where VGO is the oil fraction that boils

above 343 Β°C and Residue is the oil fraction with boiling point above 550 Β°C. The error of this

characterization technique is 1% for the light fractions (<550 Β°C) and 4% for heavier fractions

(>550 Β°C).78

𝑋343Β°C+ [%] = ( 1 βˆ’π‘‰πΊπ‘‚ + 𝑅𝑒𝑠𝑖𝑑𝑒𝑒 π‘œπ‘“ π‘ π‘Žπ‘šπ‘π‘™π‘’

𝑉𝐺𝑂 + 𝑅𝑒𝑠𝑖𝑑𝑒𝑒 π‘œπ‘“ 𝑓𝑒𝑒𝑑) βˆ— 100 Eq. 3.4

33

The feedstock and products obtained from lignocellulose biomass contain a great quantity of

polar molecules.14 These molecules are not completely soluble in nonpolar solvents such as CS2.

In this way, to account for the CS2-insolubles left out, samples were filtered and a quantification

of the insolubles was done. For that purpose, approximately 1 g of sample was diluted in 100 mL

of CS2. The solution was passed through a 0.45 ΞΌm membrane (previously weighted) using a

vacuum pump to accelerate filtration. When all the solution was filtrated, the membrane plus the

solids were dried in a VWR oven at 80 ΒΊC. Finally, the membrane and solids were weighted and

the CS2-insolubles were calculated as a percentage of the initial sample following Eq. 3.5.

πΌπ‘›π‘ π‘œπ‘™π‘’π‘π‘™π‘’π‘  𝑖𝑛 𝐢𝑆2 [%] = π‘€π‘Žπ‘ π‘  π‘œπ‘“ π‘“π‘–π‘™π‘‘π‘’π‘Ÿπ‘’π‘‘ π‘ π‘œπ‘™π‘–π‘‘π‘  [𝑔]

π‘€π‘Žπ‘ π‘  π‘œπ‘“ π‘ π‘Žπ‘šπ‘π‘™π‘’ [𝑔]βˆ— 100 Eq. 3.5

3.4.4. Viscosity

A Brookfield viscometer model DV-II+ Pro coupled with a water recirculation system model

TC-502 was used to determine dynamic viscosity. The temperature range for the equipment is

between 0 and 100 ΒΊC. The measurement starts by setting up the temperature controller at 40 ΒΊC.

Next, the spindle or the measuring device was screwed to the bottom of the motor and the

viscometer was closed with the sample cell. Once the temperature was reached, the gap between

the bottom of the sample cell and the spindle was adjusted to a value of 0.1 mm. Then, an amount

of sample enough to cover the surface of the spindle was placed in the sample cell. The viscometer

was closed and the rotation engine was started. The rotation speed was adjusted until reaching a

torque of 50-70%. The shear is generated by the cohesive forces between the fluid and the metal

plates. Rotation was continued until the spindle completed at least 5 full rotations to guarantee

proper formation of the fluid film. Finally, the dynamic viscosity (in cP) was reported by the

equipment. Viscosity reduction can be determined using Eq. 3.6. The relative error for the viscosity

measurement is Β± 5%.

π‘‰π‘–π‘ π‘π‘œπ‘ π‘–π‘‘π‘¦ π‘Ÿπ‘’π‘‘π‘’π‘π‘‘π‘–π‘œπ‘› [%] = (1 βˆ’π‘‰π‘–π‘ π‘π‘œπ‘ π‘–π‘‘π‘¦ π‘œπ‘“ π‘ π‘Žπ‘šπ‘π‘™π‘’ [𝑐𝑃]

π‘‰π‘–π‘ π‘π‘œπ‘ π‘–π‘‘π‘¦ π‘œπ‘“ 𝑓𝑒𝑒𝑑 [𝑐𝑃]) βˆ— 100 Eq. 3.6

3.4.5. Thermogravimetric Analysis (TGA)

Thermogravimetric analysis consists on analyzing the heat and weight changes experienced

by a liquid or solid sample when submitted to an increase of temperature under the flow of a gas.

34

For this method, a DT Q 600 system from β€œThermal Analysis Instruments Company” was used.

The oil sample (approx. 10 mg) is heated at 10 ΒΊC/min up to 1000 ΒΊC under a nitrogen flow of 100

Std. mL/min. The equipment produced data for weight loss, differential mass loss, heat flow and

differential heat flow for different temperatures.

3.4.6. CHN Elemental Analysis

A Perkin Elmer 2400 CHN Analyzer was used to determine the elemental composition of

carbon, hydrogen and nitrogen of the samples. The composition of these elements was used to

determine the H/C ratio and the percentage of oxygen removed of the samples, the later determined

by difference since sulphur contents are negligible. This characterization technique was conducted

in the Department of Chemistry Instrumentation Facility at the University of Calgary following

the ASTM D5291 norm.79 In this method, the combustion of the sample to form CO2, H2O and

NOx was reached at very high temperature (1000 Β°C) in a combustion tube loaded with an oxidation

catalyst. Next, NOx was reduced to N2 in a reduction tube and passed through a separation column

to be detected by a thermal conductivity detector (TCD). Finally, the composition of carbon,

hydrogen and nitrogen were determined using a standard chemical as reference and oxygen was

calculated by difference of the other elements. To evaluate the conversion of oxygen, the degree

of deoxygenation (DOD) was defined following Eq. 3.7. The relative error for each component is

different, for carbon is 0.5%, for hydrogen is 2% and for oxygen is 3%.

The high heating value (HHV) of a bio-oil is defined by the formula presented in Eq. 3.8.9

𝐷𝑂𝐷 [%] = (1 βˆ’π‘‚π‘₯𝑦𝑔𝑒𝑛 π‘π‘œπ‘›π‘‘π‘’π‘›π‘‘ 𝑖𝑛 π‘π‘Ÿπ‘œπ‘‘π‘’π‘π‘‘ [𝑀𝑑. %]

𝑂π‘₯𝑦𝑔𝑒𝑛 π‘π‘œπ‘›π‘‘π‘’π‘›π‘‘ 𝑖𝑛 𝑓𝑒𝑒𝑑 [𝑀𝑑. %]) βˆ— 100 Eq. 3.7

𝐻𝐻𝑉 [𝑀𝐽

π‘˜π‘”] = 0.335(𝐢) + 1.423(𝐻) βˆ’ 0.154(𝑂) βˆ’ 0.145(𝑁) Eq. 3.8

3.4.7. Microcarbon residue (MCR)

Microcarbon residue (MCR) determines the carbon residue remaining after evaporation and

pyrolysis of an oil under given conditions. This property is an indicative of the coke forming

tendency of an oil under thermal degradation conditions.

35

MCR was determined using the muffle furnace method developed by Hassan, Carbognani and

Pereira-Almao.80 This method is an in-house modification of the ASTM norms (ASTM D-189, D-

524 and D-4530) which reduces the analysis time and increases the samples turnaround. For this

analysis, between 10 and 20 mg of sample were weighted in a 2 mL glass vial on a 5-digits Mettler

Toledo XS 205 balance. Each sample was weighted twice. Vials were placed on the sample

platform assembly inside the muffle furnace. The platform is equipped with 26 nitrogen injection

tubes (1 per sample) to create an oxygen free environment. Next, vials were covered with a glass

cover with a 1/8” orifice in the middle and the nitrogen flush was started with a flow of 900 mL/min

for 45 min to purge the air from the furnace. Then, the temperature was increased at a rate of 10

Β°C/min until reaching a temperature of 520 Β°C that was maintained for 20 min. Finally, the furnace

was let to cool down, vials were weighted and MCR weight percentage was calculated following

Eq. 3.9. The relative error for this analytical technique is 2%.

𝑀𝐢𝑅 [%] = π‘€π‘Žπ‘ π‘  π‘œπ‘“ π‘ π‘Žπ‘šπ‘π‘™π‘’ π‘Žπ‘“π‘‘π‘’π‘Ÿ β„Žπ‘’π‘Žπ‘‘π‘–π‘›π‘” [𝑔]

π‘€π‘Žπ‘ π‘  π‘œπ‘“ π‘ π‘Žπ‘šπ‘π‘™π‘’ [𝑔]βˆ— 100 Eq. 3.9

3.4.8. Fourier-transform Infrared spectroscopy (FTIR)

Infrared spectroscopy is one of the most sensitive techniques for studying the functional

groups in solid and liquid samples.81 It is used to study the chemical footprint and main functional

groups present in oils. In this research, FTIR was vital to determine the acid reduction of the bio-

oil, more specifically, the phenols and carboxylic acids present in the feedstock.

FTIR spectra were recorded using an IRAffinity-1S spectrometer from Shimadzu. Samples

for FTIR were prepared by weighting 150 mg of oil in 10 mL of carbon tetrachloride (CCl4). This

solvent was selected because it is transparent in the 1000 to 4000 cm-1 wavelength range to avoid

its interference with the peaks of interest. The background was measured before each analysis with

a CaF2 cell containing CCl4 (dichloromethane was used for cleaning). Next, each sample was

injected in the same CaF2 cell and immediately put inside the chamber to acquire the spectrum that

ranged between 1000 to 4000 cm-1. All spectra were baseline corrected in a systematic way to

avoid subjective influence. Also, spectra were normalized by bringing the strongest absorbing

spectra signal to 10 % of transmittance. In this way, spectra for all the samples can be compared

36

between them and their relative intensities can provide information concerning chemical changes

during hydroprocessing.

The most important bands taken into consideration in this work are shown in Figure 3.3. The

band assignment is based on Silverstein’s work and it will allow the understanding of the behavior

of compounds of interest during the processing of the bio-oil.82 Region #1 is assigned to the ethers

and they are specifically located between 1350 and 1150 cm-1. Following the ethers band, there is

region #2 that represents the C=O acids, more specifically, the carboxylic acids. In this region,

there is two bands, the one at 1710 cm-1 is assigned to intermolecular bonded carboxylic acids

while the one at 1740 cm-1 corresponds to the free C=O acids. Lastly, region #3 is allocated to

phenol OH groups. The first band near 3600 cm-1 depicts phenols with no intermolecular hydrogen

bonding and the second band near 3550 cm-1 corresponds to vibrations for phenol groups forming

intermolecular hydrogen bonds with other molecules.

Figure 3.3. FTIR spectra for bio-oil with the most important bands

In this research, to calculate the carboxylic acids and phenols reduction, the transmittance

given by FTIR was used. Since each compound is assigned to two bands, the first step was to

obtain the transmittance for the bands of interest with the computer software from the FTIR. Then,

37

the two bands were averaged for the products obtained by hydrotreating and for the feedstock.

Once we had the average of the transmittance for each compound, Eq. 3.10 was used to calculate

the reduction of carboxylic acids and phenols separately.

𝐴𝑐𝑖𝑑 π‘Ÿπ‘’π‘‘π‘’π‘π‘‘π‘–π‘œπ‘› [%] = π‘‡π‘Ÿπ‘Žπ‘›π‘ π‘šπ‘–π‘‘π‘‘π‘Žπ‘›π‘π‘’π‘“π‘’π‘’π‘‘ βˆ’ π‘‡π‘Ÿπ‘Žπ‘›π‘ π‘šπ‘–π‘‘π‘‘π‘Žπ‘›π‘π‘’π‘π‘Ÿπ‘œπ‘‘π‘’π‘π‘‘

π‘‡π‘Ÿπ‘Žπ‘›π‘ π‘šπ‘–π‘‘π‘‘π‘Žπ‘›π‘π‘’π‘“π‘’π‘’π‘‘βˆ— 100 Eq. 3.10

3.4.9. Pre-asphaltenes stability

One similarity of crude oil and bio-oil is that both have polyaromatic oxygentated compounds,

deriving from lignin decomposition, suspended or peptized in the media by smaller similar

molecules. These large molecules could tend to precipitate when submitted to a high severity

process that could change the peptizing medium. To determine the stability of these molecules

after reaction in the products, a small drop of sample was transferred with a clip to a microscope

slide and covered with its corresponding cover slip. Then, the sample was examined in an optical

digital microscope model DC3-163 supplied by National Instruments with a magnification of 40X

for visual discard or confirmation of precipitated solids. It is important that the microscope slide

is heated to guarantee that paraffins, if existing, are melted and are not confused with stable solid

precipitates.

3.4.10. Gas Analysis

In order to determine the molar composition of the gases generated in the hydrotreating

process it is necessary to perform Gas Analysis. A SRI Instruments chromatograph model 8610C

was used, equipped with 4 columns, 2 TCD detectors, 2 switching valves, 1 flame photometric

detector (FPD) and 1 flame ionization detector (FID). The first TCD contains a set of 183 cm

molecular sieve column model MS13X and a silica gel column of the same length. The first TCD

is operated with helium as a carrier gas to detect hydrocarbons within the C1-C5 range, hydrogen

at high concentrations and permanent gasses such as CO2 and CO. Hydrogen and helium have

similar conductivity, therefore it is difficult to detect low concentrations of H2 using He as carrier

gas. In this way, the second TCD uses argon as carrier gas to detect hydrogen at low concentrations

using a 3 feet molecular sieve column model MSX13X. Sulphur compounds can be detected with

the FPD detector. Finally, to quantify the amount of light hydrocarbons within the C1-C5 range,

the FID detector can be used. Both FPD and FID detectors operate with a mixture of hydrogen and

38

air as carrier and use a 60 m capillary column model MXT1. The calibration of the system was

done using standard calibration gases with known and certified compositions provided by Praxair.

The relative errors between the gas composition from the standards and the values determined by

the equipment are presented in Table 3.2. Standard 1 was used for the first 4 gases presented in

Table 3.2 (H2, CO, CO2 and CH4) while Standard 2 was used to calibrate the rest of the gases. It

is important to mention that the error is higher when the real composition of the gas is considerably

higher or lower than the values from the standard.

Table 3.2. Relative error for gas chromatography

Component Standard 1 (% mol) Standard 2 (% mol) Relative Error (%)

Hydrogen (H2) 13.60 10.00 3.4

Carbon monoxide (CO) 0.00 15.00 0.6

Carbon dioxide (CO2) 3.99 10.00 0.9

Methane (CH4) 3.99 5.00 0.5

Ethane 15.90 5.00 0.8

Ethylene 1.50 0.00 1.4

Propane 8.21 0.00 0.9

Propylene 3.10 0.00 1.0

n-Butane 2.60 0.00 1.3

i-Butane 1.00 0.00 1.4

1-Butane 2.39 0.00 1.5

i-Pentane 0.10 0.00 6.5

39

Chapter 4: Results and Discussion

In this chapter, the results obtained from hydroprocessing the bio-oil are presented. Through

this research, test run or reaction (R1) was performed to study the effect of the total operating

pressure in the reaction and products quality. Once the operating pressure was selected, reaction #2

(R2) was carried out to assess the temperature and space velocity effect in the products quality and

more specifically, in the acids reduction. Next, reaction #3 (R3) was used to evaluate the effect of

increasing the temperature on the properties of the products. Then, a dual-catalyst bed

configuration was tested in reaction #4 (R4) to test the catalyst effect on the acids reduction.

Finally, one last catalyst was tested in reaction #5 (R5) to compare the effect of different catalysts

on the quality of the products. A summary of the pilot plant information for the reactions carried

out is presented in Appendix II. Moreover, the results presented for each condition tested in the

present research were for 3 or more mass balances collected after a stability time, i.e. at steady

state.

The 5 catalytic experiments were performed using 4 different in-house formulated

hydroprocessing catalysts identified as CAT-M2, CAT-M3, CAT-M3+ and CAT-M4. The

catalysts were synthesized and provided by Dr. Gerardo Vitale from the CAFE research group and

details about their preparation and formulation cannot be provided to secure patentability.

CAT-M2 was used in the first reaction and contains a highly-dispersed hydrogenating metal

supported on a Lewis-type ceramic acidic phase. CAT-M3 was used in R2 and R3 and consisted

of a modified version of CAT-M2 that contained a higher amount of the hydrogenating metal

(~30% more). CAT-M3+ was the name given to the dual-catalyst bed configuration used in R4.

In this reaction, two catalysts were introduced inside the reactor. The first bed occupied 70% of

the reactor bed and contained pure CAT-M3, the same catalyst used in R2 and R3. The second bed

occupied 30% of the reactor bed and contained a solid herein named β€œ+” that consisted of a

bifunctional catalyst having a BrΓΈnsted-type acidic phase and another hydrogenating metallic

phase capable of hydrogenating and cracking aromatic compounds. Finally, CAT-M4 consisted

of a modified version of CAT-M3 that involved the addition of a second active hydrogenating

component, where the first metal composition remained the same as in CAT-M3.

40

4.1. Effect of the total operating pressure

As a first step of the research, the effect of the total operating pressure on the quality of the

products was assessed. Trujillo in her MSc. Thesis evaluated different operating conditions for the

hydrogenation of Steeper bio-oil and found that by using CAT-M2, the lowest TAN value was

achieved at 310 Β°C, 0.2 h-1 and 1400 psig.21 This was the starting point for the study of pressure.

It is reported in the literature that a high pressure range, between 430 and 2600 psig, is used for

HDT reactions.61 However, the higher the hydrogen pressure used for this process, the higher the

safety issues and costs associated due to expenses for high-pressure vessels among other

requirements.8 Hence, pressures lower than 1400 psig were evaluated to see if the same or better

qualities could been obtained at different conditions. The experiment started with a pressure of

1400 psig, followed by 1200 psig, 900 psig and the last pressure evaluated was back to 1100 psig

at constant temperature of 310 Β°C, space velocity 0.2 h-1 and using CAT-M2.

Table 4.1 shows the results obtained for some properties of the hydrotreated-bio-oil (HDT-

bio-oil). From this table, it is observed that the TAN conversion is similar for 1400 psig, 1200 psig

and 1100 psig, while for 900 psig there is a significant difference. In general, the 900 psig condition

showed a lower quality HDT-bio-oil as noticed when comparing the viscosity, conversion, MCR

and oxygen content reduction. Comparing the viscosities, the lower value was obtained at 1100

psig followed by the one obtained at 1400 psig. Nevertheless, the values for conversion at 343 Β°C+,

MCR and DOD for the 1100 psig condition were poorer than for 1200 psig and 1400 psig. As a

general conclusion from the results presented in Table 4.1, it seems that the TAN conversion,

MCR, conversion and DOD reached a plateau when submitted to pressures higher than 1200 psig.

Another aspect that is important to highlight is that pressure 1100 psig was tested after 900

psig. The DOD obtained for 900 psig is higher than for 1100 psig, which could mean that the

catalyst could have deactivated when submitted to a pressure of 900 psig. Other possible

explanation could be the error associated to the calculation of the DOD, which carries the error of

the equipment plus the error of the oxygen calculation by difference from the other elements.

41

Table 4.1. Characterization of HDT-bio-oil at 310 Β°C, 0.2 h-1 and different operating pressures

Property

Pressure [psig]

Feed 1400 1200 1100 900

TAN Conversion [%] - 80.8 81.9 80.9 76.9

Viscosity [cP] @ 40 Β°C 31172 9622 11513 7586 17148

Conversion 343 Β°C+ [wt. %] - 19.6 20.9 17.1 15.3

MCR [wt. %] 22.01 17.41 17.57 18.31 18.31

DOD [%] - 38.3 43.3 30.8 34.0

Even though 1200 psig appears to be the best operating pressure in terms of most of the

properties and will pose as a more economical option compared with 1400 psig, some other

parameters need to be taken into consideration before making a conclusion about the best pressure

for the hydrotreating process. Two parameters that indicate the hydrogenation level of a reaction

are the H/C ratio of an oil and the paraffin to olefin ratio of a hydrotreated product. Figure 4.1

shows the H/C ratio and butane/butene ratio of the HDT products at different pressures.

Figure 4.1. H/C ratio and butane/butene ratio vs operating pressure for HDT-bio-oil. H/C ratios

determined over the liquid product; C4/C4= determined over the associated gas phase

0

2

4

6

8

10

12

Feed 1400 1200 1100 900

1.30

1.32

1.34

1.36

1.38

1.40

1.42

1.44

Bu

tan

e/B

ute

ne

rati

o [

mo

l%]

Pressure [psig]

H/C

rat

io [

mo

lar]

H/C C4/C4=

42

As seen in Figure 4.1, the H/C ratio increased as the operating pressure increased. At 900 psig

the H/C ratio was improved by 0.02 points while at 1400 psig it was enhanced by 0.06 points.

Furthermore, the butane/butene (C4/C4=) ratio also incremented as the pressure increased. The

slope for the C4/C4= ratio sharply increased from the HDT-bio-oil at 1200 psig to 1400 psig

meaning that the hydrogenation at 1400 psig was greatly improved.

Finally, the pressure selected for continuing this research was 1400 psig due to its

improvement on the hydrogenation performance reflected both in the high H/C ratio and the lowest

MCR. The next objective was to temperature and space velocity as these parameters also have

great influence on the reaction. The 1400 psig pressure was selected to ensure a higher solubility

of hydrogen in the oil and a higher availability of hydrogen near the catalyst62 to avoid coke

formation due to an increase in the the severity of the reaction.

4.2. Temperature and space velocity screening

After the pressure was set to 1400 psig, a screening of temperature and space velocity was

done in reaction #2. Figure 4.2 presents a diagram with the changes of temperature and space

velocity during R2. For this reaction, a new catalyst formulation (CAT-M3) with a higher

composition of the hydrogenation metal was tested. R2 started with the best conditions tested

before: 310 Β°C, 0.2 h-1 and 1400 psig.

One of the goals of this screening was to obtain a product with zero TAN for further processing

in CSC. Nevertheless, some limitations were encountered with the equipment available in the

laboratory when trying to analyze samples with a TAN value lower than 2 for bio-oils. Thus, to

analyze the acidity of the HDT-bio-oil, FTIR was used. A correlation between these two

characterization techniques was performed and is presented in Section 4.2.3. Secondly, the catalyst

lifetime was also tested in R2, which was a long run test that lasted 55 days. The results obtained

for the catalyst lifetime are presented in Section 4.2.4. For this study, the same condition was

evaluated at the beginning of the reaction, in the middle of the reaction and at the end of the

reaction as it can be seen in Figure 4.2.

From Figure 4.2, it can be observed that the first stage of R2 was the temperature screening,

where temperatures between 295-325 Β°C were tested at 1400 psig, 0.2 h-1 using CAT-M3. The

second stage was to evaluate the space velocity between the 0.2-0.5 h-1 range at 315 Β°C. An

43

additional condition (320 Β°C, 0.3 h-1) was also studied in this stage. Finally, as aforementioned, the

last condition tested (310 Β°C, 0.2 h-1) was carried out with the objective of evaluating the catalyst

lifetime.

Figure 4.2. Temperature and space velocity changes during R2. In order to verify the stable

behavior of the catalyst the return of the initial condition was performed twice with the last one

being at the end of the whole test run.

4.2.1. Temperature effect

A temperature screening between 295-325 Β°C was performed to study the effect of the

temperature on TAN conversion of bio-oils, while monitoring viscosity and solids stability to

avoid extensive cracking of the feedstock.

Figure 4.3 presents the results obtained for TAN conversion and viscosity reduction at

different temperatures. It can be observed that the TAN conversion for CAT-M3 started at 80 %

when submitted to a temperature of 295 Β°C and increased as the temperature increased to the point

44

of almost reaching 100 % TAN conversion. It is important to mention that the TAN value for

temperatures higher than 310 Β°C were calculated using the correlation developed in this research

because of the limitations encountered trying to detect a TAN value lower than 2 mgKOH/g with

the standard titration procedure.

When comparing the TAN conversion (81 %) obtained for R1 at 310 Β°C using CAT-M2 with

the one obtained for R2 (95 %) it is noticeable that the hydrogenating metal composition did affect

positively the catalyst performance. Moreover, the viscosity reduction obtained in R1 was 70 % at

310 Β°C, 1400 psig and 0.2 h-1 whereas the one achieved in R2 for the same operating conditions

was 80 %, supporting the fact that CAT-M3 performed better for hydrotreating bio-oil than CAT-

M2.

Figure 4.3. Effect of the temperature on the TAN and viscosity reduction (1400 psig, 0.2 h-1).

Viscosities were determined at 40Β°C

Regarding the effect of the temperature on the viscosity reduction, Figure 4.3 shows that the

viscosity reduction followed the same trend as the TAN conversion, as the temperature increased,

the viscosity reduction increased. The maximum viscosity reduction reached in this reaction was

30

40

50

60

70

80

90

100

110

50

55

60

65

70

75

80

85

90

95

100

290 300 310 320 330

Vis

cosi

ty R

edu

ctio

n [

%]

TAN

Co

nve

rsio

n [

%]

Temperature [Β°C]TAN Reduction Viscosity Reduction

45

95 %. The differential of the viscosity reduction between 295 Β°C and 300 Β°C seems to be more

severe than for the other temperatures. One explanation for this could be that there are

intermolecular bonds between some molecules in the feedstock such as carboxylic acids and

phenols that are broken when they are expose to a thermal process. These bonds could make two

light molecules pose as one heavy molecule or aggregate when measuring the viscosity. This

hypothesis can be supported by FTIR where the vibrations of each molecule are detected

separately.

Figure 4.4 shows the FTIR spectra for bio-oil and HDT-bio-oil at the lower (295 Β°C) and

higher (325 Β°C) temperatures tested. In both bio-oil and HDT-bio-oil, there are free and bonded

OH stretching signals as well as free and bonded C=O acid bands. The bonded peaks, as indicated

by the arrows in Figure 4.4, correspond to molecules that are forming intermolecular hydrogen

bonds.82 From this figure, it seems that the intermolecular bonds are mainly formed between

phenols and C=O acids, i.e. when acids were reduced, bonded phenolic –OH decreased despite the

total phenolic –OH remained constant. In the case of the HDT-bio-oil, the phenol bonded peak

was severely reduced for both temperatures and it seems that was converted to free phenols,

meaning that intermolecular bonds were being broken while maintaining a similar phenol

composition to the original feedstock. As for the C=O acid peaks, both free and bonded bands

were significantly reduced for both temperatures which supports the proposition that the hydrogen

intermolecular bonds were being cleaved by thermal or catalytic effects.

In Figure 4.4, it can be noticed a difference in the intensity of the C=O acid peaks for the 2

temperatures evaluated. HDT-bio-oil at 295 Β°C has a higher C=O acid peaks compared to HDT-

bio-oil at 325 Β°C. This is also supported by the difference in TAN conversion achieved for each

condition. However, the phenol peak at 325 Β°C remained similar as the one at 295 Β°C. When

calculating the phenol reduction for each condition evaluated, the maximum phenol reduction

achieved was 11 % of the original composition of the feedstock. Hence, in the conditions studied

in this reaction, the phenols were not significantly reduced as it was expected, because it is reported

by several authors that phenols are not hydrotreated at temperatures lower than 350 Β°C.15, 17

46

Figure 4.4. FTIR spectra for the feedstock (bio-oil) and HDT-bio-oil at two temperatures.

Figure 4.5 shows the carboxylic acids reduction versus the degree of deoxygenation for HDT-

bio-oil at the temperatures evaluated in R2. It can be observed that the C=O acids reduction follows

a linear trend with the DOD, presenting a correlation factor of 0.9713. Since the data plotted in

Figure 4.5 was obtained from experimental work, the correlation factor shows that the linear trend

is a good fit.

The linearity found between the C=O acids reduction and DOD supports the results obtained

by FTIR for acid compositions. The DOD is a parameter that takes into consideration the total

oxygen removal from the feedstock. From literature review, it is known that the main compounds

found in bio-oils are acids and phenols,39, 42 which means that oxygen atoms are primarily

converted from the carboxylic acids and some phenols. The results found by FTIR showed that the

maximum phenols reduction was 11 % while the maximum C=O acids reduction was 98 %.

Because the C=O acids reduction was around 9 times higher than the phenols reduction, poor

phenols removal did not affect significantly the linearity between C=O acid reduction and DOD.

47

Figure 4.5. Carboxylic acids reduction vs DOD for HDT-bio-oil at different temperatures.

Values in parenthesis are set up experimental temperatures

The changes of the H/C ratio and O/C ratio with the DOD are presented in Figure 4.6. As it

was expected, the O/C ratio follows a linear trend with the DOD and with a high correlation factor

of 0.9997. As the DOD increases, the O/C ratio of the HDT-bio-oil decreases, whereas the H/C

ratio increases. Both trends are indications of the deep hydrogenation achieved in the process using

CAT-M3. The maximum H/C ratio obtained was 1.47, which is closer to the value of an H/C ratio

from a typical crude oil. Nevertheless, the minimum O/C ratio obtained was 0.04, still far above

from the ones usually found for crude oils (0.01).

In order to check the samples stability and solids presence of the HDT-bio-oil, microscopic

images (see description of method, equipment and augmentation to obtain this images in

experimental part), shown in Figure 4.7, were taken for the feed and the product of the higher

temperature studied in R2 (325 Β°C). Comparing the microscope images for the feed and the HDT-

bio-oil at 325 Β°C, it can be noticed that there is no solid precipitation in the HDT-bio-oil. This

means that up to 325 Β°C the solids in the bio-oil are still stable and there is no risk of plugging the

reactor.

48

Figure 4.6. H/C ratio and O/C ratio vs DOD for HDT-bio-oil at 1400 psig, 0.2 h-1 using CAT-M3

Figure 4.7. Microscope images at 40 X for the bio-oil feed (left) and HDT-bio-oil at 325 Β°C

(right)

The hydrogen consumption and the water yield as a function of the degree of deoxygenation

in R2 are presented in Figure 4.8. As it was expected, both parameters increased as the DOD

y = -0.0011x + 0.1016RΒ² = 0.9997

y = 0.0016x + 1.3562RΒ² = 0.9289

1.34

1.36

1.38

1.40

1.42

1.44

1.46

1.48

0.000

0.020

0.040

0.060

0.080

0.100

0.120

0.0 10.0 20.0 30.0 40.0 50.0 60.0 70.0

H/C

[m

ola

r]

O/C

[m

ola

r]

DOD [%]

O/C H/C

49

increased and have a linear relation with the oxygen reduction. However, the slope for the H2

consumed is more than 2 times higher than the slope for the water yield. This means that more

than half of the H2 consumed in the reaction was not used for the hydrodeoxygenation reaction

that removes oxygen and produces water as a secondary product without breaking the hydrocarbon

chain.52 From the preceding, more than one half of the hydrogen was then consumed in the other

reactions regarding hydrotreating such as hydrodecarbonylation, hydrodecarboxylation and

secondary reactions like methanation or saturation of olefins. The secondary products from these

reactions are carbon monoxide, carbon dioxide, methane and water. CO, CO2 and methane

compositions can be followed by analyzing the gases produced in the hydrotreating process.

Details about how hydrogen consumption was calculated are presented in Appendix III.

Figure 4.8. Hydrogen consumption and water yield vs DOD for HDT-bio-oil in R2

Figure 4.9 shows the gas yield distribution for different temperatures in R2. In this graph, the

CO2 yield increased from 300 Β°C to 310 Β°C and stayed in a similar composition for the other

temperatures. The methane yield remained between 0.6 and 1.0 % throughout all the reaction,

which led to think that methanation reaction was not promoted by increasing the temperature up

y = 0.2659x + 0.5951RΒ² = 0.9528

y = 0.1054x + 0.4109RΒ² = 0.932

0

2

4

6

8

10

12

0

2

4

6

8

10

12

14

16

18

20

0 20 40 60 80

Wat

er y

ield

[%

]

H2

con

sum

pti

on

[m

g H2/g

oil]

DOD [%]

Hydrogen consumed Water yield

50

to 325 Β°C. The ethane and propane yield increased as the temperature was increased possibly due

to hydrocracking reactions promoted by higher temperatures. Conversely, the 1-butene yield

decreased as the temperature increased, meaning that the hydrogenation of olefins was enhanced

at higher temperatures. Finally, the butane yield increased from 1.7 to 2.8 % from the 300 Β°C to

the 310 Β°C temperature. Butane production could result from the dealkylation of alkyl structures

derived from lignin-produced bio-crudes.7 The increment of n-butane after reaching 310 Β°C is in

agreement with studies from the literature that have stated that alkylethers and phenolic ethers start

reacting after 300 Β°C under hydrotreatment conditions.17, 42

Figure 4.9. Gas yield distribution for different temperatures in R2

As a last result, Trujillo21 processed a HDT-bio-oil obtained in this step of the research, more

specifically, the one obtained at 320 Β°C, 0.2 h-1, 1400 psig using CAT-M3 and she found Figure

4.10. Trujillo processed 4 different feedstocks via CSC, using the same catalyst in all the runs, to

study the process and the catalyst deactivation: a bio-oil feed without hydrotreating, with a TAN

value of 48.2 mgKOH/goil, HDT-bio-oil-A, with a TAN value of 21.3 mgKOH/goil, HDT-bio-oil-B

with a TAN value of 11.6 mgKOH/goil and HDT-bio-oil-C with a TAN value of 1.2 mgKOH/goil. As

0.0

0.5

1.0

1.5

2.0

2.5

3.0

295 Β°C 300 Β°C 310 Β°C 315 Β°C 320 Β°C 325 Β°C

Gas

yie

ld [

%]

CO2 Methane Ethane Propane 1-Butene n-ButaneCO2

51

it can be seen in Figure 4.10, the catalyst used in bio-oil feed presented a rapid deactivation,

followed by the HDT-bio-oil-A and HDT-bio-oil-B. The HDT-bio-oil-C did not show any sign of

deactivation. When the TAN values trend is analyzed, it is noticeable that this parameter has an

effect on the CSC catalyst deactivation. Feedstocks with higher TAN values, promotes a faster

deactivation of the CSC catalyst.

It is important to highlight that the CSC catalyst must be stable to ensure a continuous

hydrogen production for recycling to the HDT process. Therefore, producing a high-quality HDT-

bio-oil (C) that did not deactivate the CSC catalyst was an achievement of this research.

Figure 4.10. TAN Reduction vs Time on Stream for CSC processing adapted from Trujillo.21

4.2.2. Space velocity effect

The space velocity screening was made at 315 Β°C, 1400 psig using CAT-M3 and a space

velocity range of 0.2-0.5 h-1. Also, an additional condition (320 Β°C, 1400 psig, 0.3 h-1) was tested

in reaction #2.

Space velocity is defined as the inverse of residence time and it is an important parameter that

could define the severity of a reaction. If the space velocity decreases, the residence time increases,

52

extending the contact time between the feed and the catalyst. This enhances the possibility of

hydrogenating reactions like hydrodeoxygenation occurring during hydroprocessing. Although

increasing the residence time of the bio-oil in the reactor could result in better hydrogenation of

bio-oil, a long residence time means less production per unit of time of the HDT-bio-oil when

using the same reactor size. This could be avoid by using a bigger reactor, however, it would mean

more capital cost. Therefore, it is important to study the effect of the space velocity on the

properties of the bio-oil to select the condition that uses the highest space velocity possible, i.e.

more production, while also maintaining a high HDT-bio-oil quality.

Table 4.2 shows the results obtained for the HDT-bio-oil at 1400 psig, 315 Β°C and different

space velocities using CAT-M3. The additional condition tested is presented in Table 4.3

compared with the feed and the condition with a lower space velocity (320 Β°, 0.2 h-1).

Table 4.2. Characterization of HDT-bio-oil at 1400 psig and 315 Β°C using CAT-M3. Cx

describes tested condition evaluated during R2 (See Figure 4.2)

WHSV [h-1] - 0.2 0.3 0.4 0.5

Temperature [Β°C] - 315 315 315 315

Property Feed R2C3 R2C8 R2C9 R2C10

TAN Conversion [%] - 96.6 91.5 86.1 82.7

Viscosity [cP] @ 40 Β°C 31172 4369 5989 7424 8462

Conversion 343 Β°C+ [wt. %] - 26.0 18.5 17.4 15.7

MCR [wt. %] 22.01 17.16 16.55 17.96 17.77

DOD [%] - 53.8 36.8 30.7 28.4

In both Table 4.2 and Table 4.3, it can be seen that increasing the space velocity diminishes

the quality of the HDT-bio-oil. TAN conversion, the conversion at 343 Β°C+ and DOD decrease as

the space velocity was increased. This was expected because when the space velocity is increased,

the residence time is decreased, and thus, if the contact time between the catalyst and the bio-oil

is shorter, the upgrading of the bio-oil will be less effective. The only property that did not showed

significant variations was the MCR value, which for both cases, remained approximately constant

when the space velocity was changed.

53

Table 4.3. Characterization of HDT-bio-oil at 1400 psig and 320 Β°C using CAT-M3

WHSV [h-1] - 0.2 0.3

Temperature [Β°C] - 320 320

Property Feed R2C2 R2C11

TAN Conversion [%] - 97.6 93.9

Viscosity [cP] @ 40 Β°C 31172 2566 5106

Conversion 343 Β°C+ [wt. %] - 29.6 19.6

MCR [wt. %] 22.01 16.90 16.64

DOD [%] - 55.0 39.6

The results obtained for the assessment of the space velocity effect on the product quality are

in agreement with previous studies found in the literature. For example, an investigation made by

Elliot et al. in a continuous flow reactor showed that the oxygen content was reduced from 21% to

10% when changing the WHSV from 0.70 to 0.25 h-1 using a Pd/C catalyst at 340 Β°C and 2000

psig.68

4.2.3. Correlation between TAN and Infrared absorptivity at 1710-1750 cm-1

Although TAN analysis is deemed to be an acceptable method to determine the acidity of bio-

oil, it is important to stress that the standard method (i.e.: ASTM D664) was originally developed

for measuring the acidity of petroleum-derived products, fundamentally the naphthenic acids.

Other chemical species present in bio-oil (e.g., sugars, furans, ketones, aldehydes, phenols) could

affect/contribute to the TAN values measured for these types of samples. Therefore, a second

method to detect the acids in the bio-oil samples such as FTIR should be used to corroborate the

TAN measurement.

When trying to determine the TAN value for some of the HDT-bio-oil samples obtained in

R2, some limitations were encountered using the equipment available in the laboratory. For

samples with low TAN value (<2 mgKOH/g), the equipment response was β€œNot-a-Number” and

it could be observed that the inflection point in the titration curves was undistinguishable by the

apparatus (see Figure 4.11, for instance). However, when the samples were analyzed by FTIR, the

54

band corresponding to the C=O acids was observed (Figure 4.12), meaning that the samples still

had some acids that the TAN equipment could not detect.

Figure 4.11. Graphic given by the TAN equipment for bio-oil feed (left) and HDT-bio-oil 315 Β°C

(right)

Figure 4.12. FTIR spectra of the bio-oil feed and HDT-bio-oil at 315 Β°C

To overcome the limitations presented by the TAN method and to minimize the possible error

of the TAN values for bio-oil samples, a correlation between FTIR analysis and TAN was carried

out. Figure 4.13 shows the linear correlation between transmittance obtained by FTIR and the

natural logarithm of TAN. The graph was made with results from the samples that were detected

by the TAN method (those with a TAN value higher than 2 mgKOH/g). The correlation factor for

55

the linear trend between the two parameters is 0.9740, which is within the error for correlating two

experimental values obtained by two different techniques.

Figure 4.13. Correlation between transmittance obtained by FTIR for the bands 1710 and

1740 cm-1 between 17 and ln (TAN)

The relative error between the TAN values measured by the equipment and the ones calculated

using the correlation for the HDT-bio-oil samples are presented in Table 4.4. It can be noticed that

the higher relative error obtained was 10.9 %, which is within the range of the error of the

equipment presented in Section 3.4.1. Hence, the correlation between FTIR and TAN was used to

calculate the TAN values for the samples that the standard equipment could not measure. Also, it

was implemented to obtain the TAN values for the next sets of reaction products of the present

research.

Table 4.4. Relative error between the measured and calculated TAN for different samples of

HDT-bio-oil

Sample Measured TAN

[mgKOH/g]

Calculated TAN

[mgKOH/g]

Relative

Error [%]

300 Β°C 0.2 h-1 4.8 5.2 8.4

310 Β°C 0.2 h-1 2.4 2.5 3.8

315 Β°C 0.3 h-1 4.1 3.7 10.9

315 Β°C 0.4 h-1 6.7 6.5 4.0

315 Β°C 0.5 h-1 8.4 8.7 3.9

y = -6.0432x + 97.9516RΒ² = 0.9740

80

85

90

95

100

0.5 1.0 1.5 2.0 2.5

Tra

nsm

itta

nce

(%)

Ln (TAN[mgKOH/g])

56

4.2.4. Catalyst Lifetime

One of the challenges faced by the process of hydrotreating bio-oil is the catalyst deactivation.

Conventional alumina-supported CoMo and NiMo catalysts are commonly used for HDO in the

sulfided form. However, because of the lack of sulphur compounds in the bio-oil feed, a decrease

in activity during time has been observed possibly due to transformation of the catalyst from

sulfided to an oxide form.17 Moreover, selective catalytic hydrogenation can also be carried out

with transition metal catalysts such as Pt, Rh and Pd supported in alumina. In this case, the catalysts

were reported to suffer from significant deactivation due to carbon formation.53

In order to evaluate the deactivation of the catalyst used for R2 (CAT-M3) that has the same

fundamental nature of CAT-M3+ and CAT-M4, a study of the catalyst lifetime was performed

throughout R2. As seen in Figure 4.2, the temperature 310 ΒΊC was tested three times along the

length of the reaction to monitor eventual catalyst deactivation. Figure 4.14 presents the variation

of TAN Conversion with the time on stream during reaction #2 using CAT-M3.

Figure 4.14. TAN Conversion of the HDT-bio-oil vs time on stream in R2 using CAT-M3

70

75

80

85

90

95

100

0 200 400 600 800 1000 1200 1400

TAN

Co

nve

rsio

n [

%]

Time on stream [h]

310 Β°C310 Β°C 310 Β°C

320 Β°C315 Β°C

325 Β°C

300 Β°C

295 Β°C

57

It can be observed in Figure 4.14 that CAT-M3 did not deactivate within the 55 days of the

reaction. The first condition tested in R2 was 310 Β°C and resulted in a TAN conversion of 95 %,

after 810 hours on stream and increasing the temperature to reach 325 Β°C stepwise, the temperature

was lowered from this final high temperature to 310 Β°C to check the activity of the catalyst. The

second time 310 Β°C was tested, a TAN conversion of 95 % was reached, confirming that the

catalyst was still active after almost 34 days. Finally, the temperature of 310 Β°C was tested again

to check the activity of the catalyst at the end of the reaction. In this last condition, a TAN

conversion of 95 % was achieved for the third time, confirming that the catalyst did not deactivate

throughout all the reaction with modified conditions that lasted for a total of 55 days.

4.3. Increased severity evaluation

In R2, a TAN conversion of 98 % and an important upgrading of the bio-oil in terms of

properties like viscosity and DOD was reached. However, the maximum phenols reduction

achieved was 11%. Thus, reaction #3 was carried out to evaluate the effect of a higher increase in

temperature on the properties of the bio-oil and more specifically, the phenols reduction.

It is reported in the literature that the start-up temperature for hydrotreating phenols is

350 Β°C.15, 17 For this reason, the targeted temperature of R3 was set at 350 Β°C. Nevertheless, to

avoid possible reactor plugging due to coke formation, the temperature was increased gradually.

The operating conditions for this reaction were similar to the ones used in R2 (1400 psig, 0.2 h-1

using CAT-M3) with a range of temperature screening between 310-345 Β°C. The temperatures of

310 and 325 Β°C were tested to compare the catalyst performance in R2 and R3 because the catalyst

used for R3 corresponds to a second batch of the CAT-M3. Finally, it was not possible to reach

the targeted temperature of 350 Β°C due to solid formation in the HDT-bio-oil as will be explained

below.

Table 4.5 shows the results for the HDT-bio-oil at 325 Β°C, 1400 psig and 0.2 h-1 using CAT-

M3 in different reactions (R2 and R3). Comparing the results obtained in R2 with the ones obtained

in R3 at the same conditions, it is noticeable that a similar level of upgrading was reached in both

reactions using the same catalyst. The difference in all the properties presented in Table 4.5 are

within the error of the characterization techniques used to determine them. Another possible reason

58

for the minor differences between the results obtained in R2 and R3could be the error associated

to the pilot plant or the packing of the reactor between runs.

Table 4.5. Characterization of HDT-bio-oil at 325 Β°C, 1400 psig, 0.2 h-1 using CAT-M3 in

different reactions (R2 and R3)

Property Feed R2 R3

TAN Conversion [%] - 98.3 96.8

Viscosity [cP] @ 40 Β°C 31172 1561 2040

Conversion 343 Β°C+ [wt. %] - 28.73 28.27

MCR [wt. %] 22.01 15.35 15.09

DOD [%] - 59.8 62.1

The reduction of carboxylic acids and phenols for the temperatures tested in R3 are presented

in Figure 4.15. It can be observed that the carboxylic acids reduction increased as the temperature

increased whereas the phenols reduction remained constant in around 10 % from 325 to 335 Β°C

and it increased to 21 % at 340 and 345 Β°C, as it was expected because, as mentioned before, the

phenols conversion under hydrotreating conditions start at 350 Β°C.17 In this reaction, it was

possible to achieve a 100 % C=O acids reduction at 345 Β°C.

Figure 4.15. Carboxylic acids and phenols reduction for different temperatures in R3

0

10

20

30

40

50

60

70

80

90

100

325 330 335 340 345

Red

uct

ion

Per

cen

tage

[%

]

Temperature [Β°C]

Carboxylic acids reduction Phenols reduction

59

Figure 4.16 shows the carboxylic acids reduction versus the degree of deoxygenation for

HDT-bio-oil at the temperatures evaluated in R3. It can be noticed that the C=O acids reduction

follows a linear trend with the DOD with a correlation factor of 0.9126. This factor is not as good

as the one obtained for R2 (0.9713) and it is an expected result. In this reaction, the phenols

reduction was double than in R2, and since the DOD takes into consideration all the oxygen

removed from the feed, the DOD is impacted by both carboxylic acids and phenols. Thus, a

faultless linear relation between the C=O acids and the DOD should not be expected.

Figure 4.16. Carboxylic acids and phenols reduction vs DOD in R3

The idea of this reaction was to reach at least 350 Β°C in temperature being the other conditions

the same as those in R2. The increase of temperature in this reaction was made gradually because

at higher temperatures there is more risk of plugging the reactor due to solids formation. In order

to check the stability of the HDT-bio-oil, optical microscopy images were taken for the product

obtained at different temperatures.

Figure 4.17 shows the microscopic images taken for the temperatures tested in R3. It can be

observed that there is no solid formation at 325 Β°C. However, at 330 Β°C small solids precipitation

can be observed, meaning this is the start-up temperature for solid formation using CAT-M3. It

y = 1.2206x + 13.738RΒ² = 0.9126

80

82

84

86

88

90

92

94

96

98

100

60 62 64 66 68 70 72 74

Car

bo

xylic

aci

ds

red

uct

ion

[%

]

DOD [%]

60

can also be noticed that at 340 and 345 Β°C the solid precipitation is more severe. At both

temperatures, the solids tend to agglomerate near the edges of the sample following a similar

behavior as asphaltenes.83 These solids could deposit in the surface and pores of the catalyst and

either deactivate the catalyst or plug the reactor. The solids formed from bio-oils are similar to the

coke formed from petroleum-derived oils since they consist of aromatic hydrocarbons with boiling

points between 350 and 650 Β°C.84

Figure 4.17. Microscope images at 40 X for the HDT-bio-oil at different temperatures in R3

Figure 4.18 presents the hydrogen consumption and water yield vs the DOD in R3. It can be

seen in the graph that both parameters seem to follow an exponential trend with respect the DOD.

Also, it is important to highlight the parallelism for both variables but in different ranges of values.

This could mean that as the severity of the reaction is increased, the hydrodeoxygenation reaction

is promoted over hydrodecarboxylation and hydrodecarbonylation. Additionally, it shows that the

milligrams of H2 consumed per gram of oil rise at increasingly higher rates as the DOD increases.

61

Figure 4.18. Hydrogen consumption and water yield vs DOD for HDT-bio-oil in R3

The gas yield distribution for different temperatures in R3 is presented in Figure 4.19. It can

be observed in this graph that methane, ethane and propane yields increased as the temperature

increased whereas the butane yield increased from 325 to 335 Β°C and then decreased when 345 Β°C

temperature was reached. Additionally, a sharp decrease in the 1-butene yield was observed when

the temperature was increased, reaching a concentration of zero for this olefin at 345 Β°C. When

comparing the gas distribution obtained from R2 with the one obtained from R3, it was noticed

that some CO production is found in R3 that could be formed by hydrodecarbonylation reactions

or by the undesired reverse water gas shift reaction. This reaction takes place by reacting CO2 with

hydrogen following Eq. 4.1.

𝐢𝑂2 + 𝐻2 ↔ 𝐻2𝑂 + 𝐢𝑂 Eq. 4.1

The reverse water gas shift reaction is favored with increase of temperature and because the

carbon dioxide yield remains constant when increasing the temperature, it is more plausible that

CO is formed by hydrocarbonylation reactions. Regarding other undesired reaction like

methanation, it is noticeable that the methane yield is increasing while the CO2 yield is decreasing

when the temperature is increased. In this case, methane could be resulting from the methanation

reaction presented in Eq. 4.2. Therefore, some of the oxygen was removed in the form of carbon

0

2

4

6

8

10

12

0

5

10

15

20

25

20 40 60 80

Wat

er y

ield

[%

]

H2

con

sum

pti

on

[m

g H2/g

oil]

DOD [%]Hydrogen consumed Water yield

62

dioxide from the bio-oil via the hydrodecarboxylation reaction, is consumed in-situ producing

methane via the methanation reaction shown in Eq. 4.2.

𝐢𝑂2 + 4𝐻2 ↔ 2𝐻2𝑂 + 𝐢𝐻4 Eq. 4.2

The presence of the methanation reaction at high temperatures is in agreement with the

literature and with the trend found from Figure 4.18.52 For methanation to occur, four molecules

of hydrogen must react with a molecule of carbon dioxide. Hence, there will be a sharp increase

on the miligrams of hydrogen consumed per gram of oil that will be used to produce methane and

water. Methane is not a desired product from the hydrotreating reaction point of view since it

consumes high quantities of hydrogen; its production must be avoided as much as possible.

Figure 4.19. Gas yield distribution for different temperatures in R3

4.3.1. Product distribution of the HDT-bio-oils

The liquid product samples obtained from hydrotreating bio-oil were analyzed via SimDist.

From the distillation profiles, the product distribution and the conversion at 343 Β°C were

calculated. Appendix II shows the distillation profiles for the analyzed samples.

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

325 Β°C 330 Β°C 335 Β°C 340 Β°C 345 Β°C

Gas

yie

ld [

%]

CO2 CO Methane Ethane Propane 1-Butene n-ButaneCO2

63

Due to the high polarity presented by some components of bio-oils, the feed and the HDT-

bio-oil were not fully miscible in the solvent used for the SimDist analysis (CS2). Thus, as

explained in Section 3.4.3, the samples were filtered to calculate the quantity of insolubles in CS2

per condition and to analyze the distribution of the insoluble by TGA. This insoluble material is

believed to contain most of the polar molecules originally present in the feed, more specifically,

polyphenols. The weight percentage of insoluble portion in CS2 obtained for the feed and HDT-

bio-oil in R2 and R3 are presented in Figure 4.20. It can be observed that the portion of insoluble

material does not follow a trend with respect to the temperature of the reaction and it ranges

between 3.5-7.8 wt. % in R2 and R3. Moreover, the insoluble material reduction achieved from

the feed to the HDT-bio-oil was between 60 and 80%.

Figure 4.20. Weight percentage of CS2 insoluble material at different temperatures in R2 & R3

To evaluate the insoluble material product distribution, a TGA in nitrogen was performed.

Figure 4.21 shows the TGA graph obtained for the CS2 insoluble material from the HDT-bio-oil

at 340 Β°C in R3. In order to compare the product distribution calculated from SimDist with the one

64

calculated by TGA, Carbognani et al. recently published a correlation between the two of them for

petroleum-derived oils. In this work, Eq. 4.3 was used to calculate the corresponding TGA

temperature for SimDist analysis. Thus, by having the TGA corresponding temperature for

SimDist, the product distribution or conversion can be calculated and compared with SimDist data.

The relative error derived from this conversion was determined to be Β± 10 %.85

π‘†π‘–π‘šπ·π‘–π‘ π‘‘ π‘‡π‘’π‘šπ‘π‘’π‘Ÿπ‘Žπ‘‘π‘’π‘Ÿπ‘’ [°𝐢] = 1.36 𝑇𝐺𝐴 π‘‡π‘’π‘šπ‘π‘’π‘Ÿπ‘Žπ‘‘π‘’π‘Ÿπ‘’ [°𝐢] + 90 Eq. 4.3

The product distribution range of temperature for SimDist and the one calculated for TGA

using Eq. 4.3 are showed in Table 4.6. It can be noticed that the starting temperature for each cut

is lower for the TGA than for the SimDist. According to Carbognani et al.85 the TGA instruments

have an existing thermal cracking phenomena that breaks the molecules at lower temperature than

the SimDist.

Table 4.6. Temperature range for the product cuts determined via SimDist or TGA

Product Cut

Temperature range [Β°C]

SimDist TGA

Naphtha IBP-190 IBP*-73.5

Jet Fuel 190-260 73.5-125

Diesel 260-343 125-186

VGO 343-550 186-338

Residue 550+ 338+

IBP* is the IBP calculated with Eq. 4.3

Knowing that the starting temperature for the residue fraction in the TGA is 338 Β°C, from

Figure 4.21 it can be seen that at least 95 % of the insoluble material does not convert up to this

temperature. This result confirms that the CS2 insolubles contain mostly heavy polar aggregates

that are in the range of the residue fraction. Consequently, in all the products distribution and

conversions reported in the present research, the percentage of insoluble in the liquid sample was

added to the portion of residue obtained by SimDist.

65

Figure 4.21. TGA in N2 of the CS2 insoluble material from the HDT-bio-oil at 340 Β°C in R3

Figure 4.22 shows the product distribution and conversion at 343 Β°C+ calculated for different

temperatures in R2 and R3. Comparing the feed with the HDT-bio-oil products, it is observed that

between 3.5 and 5.8 % of the original bio-oil was converted to gases and between 5.2 and 8.8 %

was converted to water. The water yield increased as the temperature increased while the gas yield

did not follow a specific trend. Regarding the naphtha and VGO fractions, they were converted up

to 2 % compared with the feed. Nevertheless, the residue yield decreased significantly when the

temperature increased and the diesel fraction showed a meaningful increased at higher

temperatures. The conversion at 343 Β°C + increased around 11% from 300 Β°C to 320 Β°C and

remained constant until reaching 38.1 % at 340 Β°C.

From Figure 4.22, it seems that the residue fraction is mainly being converted to gases, water

and diesel, since the other product cuts did not show a significant variation when residue fraction

presented an important reduction. To support this idea, the product yields were plotted versus the

conversion of the product at 343 Β°C+ in Figure 4.23. The amount of residue (550 Β°C+) tend to

sharply decreases as the conversion increases, while VGO remained practically unchanged,

possibly because its rate of consumption is similar to its rate of generation. On the other hand, light

cuts such as naphtha and jet fuel moderately increased while the diesel fraction tended to severely

increase as the conversion increased.

66

Figure 4.22. Product distribution and conversion at 343 Β°C+ at different temperatures in R2 & R3

Figure 4.23. Product yields vs conversion at 343 Β°C+ for different conditions in R2 & R3

0

10

20

30

40

50

60

0 10 20 30 40

Yiel

d [

wt.

%]

Conversion 343Β°C+ [wt. %] Gas Water Naphtha Jet Fuel Diesel VGO Residue

67

4.4. Evaluation of a dual-catalyst bed reactor (Reaction #4)

In order to achieve a further conversion of the phenols, a dual-catalyst bed configuration

herein named CAT-M3+ was tested. As explained before, the first bed of catalyst (CAT-M3) has

a strong hydrogenation-dehydrogenation activity while the second bed (+) it was intended to

benefit the adsorption of phenols and the activation of their C=C bonds to be hydrogenated and

cracked (ring opening), to selectively eliminate the phenols. The conditions used in this reaction

were 1400 psig, 0.2 h-1 with respect CAT-M3, 0.5 h-1 with respect to catalyst + and a range of

temperature between 330 and 340 Β°C.

Figure 4.24 shows the reduction of carboxylic acids and phenols for the temperatures tested

in R4 using CAT-M3+. Reasonable trends were observed for the conversion of C=O acids and

phenols with the increase of temperature. Both reduction percentages increased as the severity of

the reaction was increased. Regarding the phenols conversion, it can be observed that at 330 Β°C, a

reduction of phenols of almost 20 % was achieved and when the temperature reached 340 Β°C, the

phenols were reduced 28 % from the original feed content. CAT-M3+ showed a better catalytic

activity towards phenols conversion but it did not achieve the same C=O acids reduction as CAT-

M3.

Figure 4.24. Carboxylic acids and phenols reduction for different temperatures using CAT-M3+

0

10

20

30

40

50

60

70

80

90

100

330 335 340

Red

uct

ion

Per

cen

tage

[%

]

Temperature [Β°C]

Carboxylic acids reduction Phenols reduction

68

As the severity of the reaction increased, the stability of the liquid product was monitored as

in the other reactions. Microscope images taken for the HDT-bio-oil at different temperatures in

R4 are presented in Figure 4.25. Analyzing the stability of the liquid samples, it can be seen that

at 330 Β°C the HDT-bio-oil was still stable but at 335 Β°C, some solids started to precipitate. At

340 Β°C, the solids remained relatively constant compared with 335 Β°C. However, to avoid further

deterioration of the HDT-bio-oil and possible reactor plugging, the reaction was ended at 340 Β°C.

Figure 4.25. Microscopic images at 40 X for the HDT-bio-oil at different temperatures in R4

Some relevant properties of the HDT-bio-oil at different temperature conditions using CAT-

M3+ are summarized in Table 4.7. A reduction in the viscosity and MCR values was found as the

temperature increased. On the other hand, the DOD increased sharply from 335 Β°C to 340 Β°C,

reaching 69 % of oxygen removal from the bio-oil feed. The conversion of the product at 343 Β°C+

did not follow a definite trend, it increased 9 % from 330 Β°C to 335 Β°C and decreased 3.5 % when

the 340 Β°C temperature was reached. It is known that the conversion of an oil usually increases

with temperature rise, hence more conditions need to be evaluated to conclude if one of the points

(335 or 340 Β°C) is an outlier or if the catalyst mixture affected the usual trend.

69

Table 4.7. Characterization of HDT-bio-oil at different temperatures using CAT-M3+

Property

Temperature [Β°C]

Feed 330 335 340

Viscosity [cP] @ 40 Β°C 31172 3136 576 492

Conversion 343 Β°C+ [wt. %] - 23.5 32.6 28.9

MCR [wt. %] 22.01 16.99 14.84 14.76

DOD [%] - 56.8 58.3 69.0

4.5. Catalyst performance comparison

A fifth reaction was carried out to evaluate the incorporation of a second hydrogenating

component to the CAT-M3 formulation and it was named as CAT-M4. This catalyst was tested to

see if it was possible to achieve a deeper conversion of phenols by improving the reactivity of

oxygenated groups with this second hydrogenating metal. The operating conditions used in this

reaction were 1400 psig, 0.2 h-1 and a range of temperature between 330 and 340 Β°C using CAT-

M4. The results of this reaction (R5) were compared with the ones obtained in R3 and R4 and are

presented in this section with the purpose of comparing the catalysts performance.

Figure 4.26 presents the acidity reduction (for both carboxylic acids and phenols) estimated

by FTIR quantitative analysis method at different temperatures and using different catalysts. As a

general trend, both C=O acids and phenols reduction increased as the severity of the reaction

increased. The highest carboxylic acids conversion was achieved by CAT-M3, followed by CAT-

M+ and CAT-M4. In theory, CAT-M3+ should have reached the same carboxylic acid reduction

as CAT-M3 at the same conditions because this dual-catalyst bed is the combination of CAT-M3

at the same space velocity tested in R3 (using CAT-M3) and a second bed composed by an acidic

catalyst. However, in R4 the weight of CAT-M3 was reduced to maintain a similar total weight of

catalyst in the reactor as in R3, i.e. in R3 29 g of CAT-M3 were used whereas in R4 21g of CAT-

M3 plus 8 g of catalyst + was employed. The difference in the weight of the portion corresponding

to CAT-M3 influences the linear velocity of this reactor bed. A lower linear velocity could have

70

negatively affected the performance of CAT-M3 in this experiment, explaining the difference in

the carboxylic acids reduction between R3 and R4. It can also be established that the catalyst + did

not contribute with the reduction of carboxylic acids. Analyzing CAT-M4 activity with respect to

the carboxylic acids conversion, it was noticed that the catalyst did not perform as it was expected.

This catalyst was supposed to improve the hydrogenation reactions due to its dual-hydrogenation

metal composition. However, one possible explanation for its poor performance might be that the

second hydrogenating metal phase could have covered the first hydrogenating phase, reducing the

quantity of active phases in CAT-M4 instead of increasing them or forming some sort of solid

solution that produced active sites of much less activity than the original ones or sites favoring

undesired reactions.

Figure 4.26. Carboxylic acids and phenols reduction at different T and catalysts

0

10

20

30

40

50

60

70

80

90

100

330 335 340

Red

uct

ion

Per

cen

tage

[%

]

Temperature [Β°C]

CAT-M3 Carboxylic CAT-M3+ Carboxylic CAT-M4 Carboxylic

CAT-M3 Phenols CAT-M3+ Phenols CAT-M4 Phenols

71

In Figure 4.26 it can be observed that the highest activity toward phenols conversion was

found for CAT-M3+, followed by CAT-M3 and CAT-M4. The acidic phase (+) used in the dual-

catalyst bed during R4 seem to play an important role in the conversion of phenols, reaching a

reduction of 28 % at 340 Β°C. The reason why CAT-M4 presented a poor phenols conversion could

be that adding two hydrogenating metals covered more the acidic support, reducing the acidic sites

of the catalyst needed for cracking and ring opening.

On the other hand, there is not a significant difference between the phenols reduction obtained

at 335 and 340 Β°C using CAT-M3+. The phenols present in the bio-oil have different sizes and

complexities. The molecules originated mostly from lignin depolimerization are known to be small

molecules such as cathecol and cresol or more complex structures with different side chain groups

such as methoxy, alkyl ester, alkyl methoxy and ether linkages.27, 39, 42 One of the reasons for

reaching a stagnant point for the phenols conversion could be that the catalyst pore diameter is not

big enough for the bulky molecules to reach the active sites confined in the catalyst crystals, thus

only the small molecules are being converted by this second catalytic bed.

Figure 4.27 illustrates the viscosity profile as a function of the conversion at 343 Β°C+ for the

HDT-bio-oil in R2, R3, R4 and R5. In this figure, it can be observed that the natural logarithm of

the viscosity follows a linear trend with the conversion of the HDT-bio-oil and the viscosity of the

liquid product is reduced several orders of magnitude at the maximum conversions for all cases.

Comparing the slopes for the reactions and catalysts tested, HDT-bio-oil using CAT-M4 presents

the poorer slope compared with the products using the other catalysts. The viscosity reduction in

R3 using CAT-M3 and in R4 using CAT-M3+ has similar slopes, meaning the addition of the β€œ+”

bed in R4 did not affect the viscosity reduction in any way. It can also be seen that the slopes for

R2 and R3 are not similar although both reactions were performed with CAT-M3. The conditions

tested in R3 were more severe than the ones carried out in R2, hence some thermal contribution

can be attributed to the conversion of products to the slope obtained for R3. To confirm this idea,

the natural logarithm of the viscosity was plotted versus the conversion at 343 Β°C+ in Figure 4.28

for the same conditions evaluated in both R2 and R3 (310 and 325 Β°C) to compare the slopes

obtained.

72

Figure 4.27. Natural logarithm of viscosity vs conversion at 343 Β°C+ for R2, R3, R4 & R5

Figure 4.28. Natural logarithm of viscosity vs conversion at 343 Β°C+ for the same conditions

tested in R2 & R3

y = -0.0807x + 10.576RΒ² = 0.9099

y = -0.1314x + 10.967RΒ² = 0.9015

y = -0.1277x + 10.451RΒ² = 0.9377

y = -0.0605x + 10.336RΒ² = 0.9568

4

5

6

7

8

9

10

11

12

0 5 10 15 20 25 30 35 40 45

ln (

Vis

cosi

ty a

t 40

Β°C

[cP

])

Conversion 343Β°C+ [wt. %]

R2-CAT-M3 R3-CAT-M3 R4-CAT-M3+ R5-CAT-M4

y = -0.0807x + 10.576RΒ² = 0.9099

y = -0.0865x + 10.493RΒ² = 0.8581

6

7

8

9

10

11

12

0 10 20 30 40

ln (

Vis

cosi

ty a

t 40

Β°C

[cP

])

Conversion 343Β°C+ [wt. %]

R2-CAT-M3 R3-CAT-M3

73

In Figure 4.28, it can be confirmed that the difference in the slopes found in Figure 4.27 for

R2 and R3 is due to the contribution of the thermal effect to the conversion when submitting the

bio-oil to high temperatures (>325 Β°C). The slopes found in Figure 4.28 by plotting the same

conditions for both reactions (R2 and R3) are almost exact, agreeing with the results presented in

Table 4.5 that established the satisfactory performance of CAT-M3 in both reactions.

As observed in Figure 4.29, the obtained HDT-bio-oil using CAT-M4 has a higher tendency

to form insoluble materials (MCR) when compared with the products from the other catalysts at

similar conversion. Some of the active sites of CAT-M4 could be covered by the second

hydrogenating phase, reducing the total active sites in this catalyst. Therefore, the hydrotreating

process relies more on thermal effects than on catalytic activity, increasing the tendency to form

insoluble materials. Similarly to the viscosity profile, the slopes for the products obtained from R3

and R4 present a parallelism, which means that the acidity of catalyst + did not affect this property

either. Again, the difference found between the slopes form the HDT-bio-oil in R2 and R3 can be

attributed to the thermal contribution on the conversion after reaching a high severity.

Figure 4.29. MCR vs conversion at 343 Β°C+ for R2, R3, R4 & R5

y = -0.1905x + 21.989RΒ² = 0.9057

y = -0.2513x + 22.048RΒ² = 0.9702

y = -0.2303x + 22.038RΒ² = 0.982

y = -0.1742x + 22.07RΒ² = 0.8938

8.50

10.50

12.50

14.50

16.50

18.50

20.50

22.50

24.50

0 10 20 30 40 50

MC

R [

%]

Conversion 343Β°C+ [wt. %]

R2-CAT-M3 R3-CAT-M3 R4-CAT-M3+ R5-CAT-M4

74

The product distribution and conversion at 343 Β°C+ for two temperatures evaluated using

different catalysts are showed in Figure 4.30. As general trends, the highest conversion was

reached using CAT-M3, followed by CAT-M4 and CAT-M3+; the conversion at 343 Β°C+

increased as the temperature was increased for all catalysts; and the gas, naphtha, jet fuel and VGO

fractions remained almost constant when the temperature was increased for all the hydrotreated

products obtained using different catalysts. The residue fraction was higher for the HDT-bio-oil

using CAT-M3+ at both temperatures and that could be because, as explained before, the bulky

molecules conforming the residue fraction could not reach the active sites confined inside the

catalyst. Even if the size of the pores is larger than the molecules reacting and products, slow mass

transport of these molecules through the pore system can also reduce considerably the reaction

rate. A conversion of 38.1 % was reached at 340 Β°C using CAT-M3, while the maximum

conversion achieved using CAT-M4 and CAT-M3+ was around 29 %.

Figure 4.30. Product distribution and conversion at 343 Β°C+ for different T and catalysts

To compare the CS2 insoluble material obtained for each catalyst at different reaction

severities, the insolubles were plotted versus the conversion at 343 Β°C+ (see Figure 4.31). In the

75

graph, there is three main regions of insoluble, a low region between 3 and 6 wt. %, a medium

region between 6 and 8 wt. % and a high region around 10 wt. %. The insoluble material formed

by using CAT-M3 for hydrotreating the bio-oil is within the low region, the one formed by using

CAT-M3+ is within the medium region, whereas the one produced using CAT-M4 is in the high

region. This agrees with the findings regarding the acidity content of the products. The HDT-bio-

oil produced using CAT-M4 presented the higher polar compounds (C=O acids and phenols)

content compared with the other catalysts. Additionally, CAT-M4 showed less catalytic activity

and more thermal degradation due to the coverage of some active sites of the first hydrogenating

metal. Promoting thermal reactions could lead to polymerization of some high-molecular-weight

compounds, resulting in more CS2 insoluble material. On the other hand, the insolubles obtained

using CAT-M3+ could be in the medium region due to a higher quantity of carboxylic acids

compared with CAT-M3. Also, some cracking due to the acidic phase added when using this

catalyst could also result in some polymerization reactions, increasing the high-molecular-weight

compounds.

Figure 4.31. Comparison of CS2 insoluble material in the liquid product obtained using different

catalysts

76

Figure 4.32 shows the microscope images taken for the HDT-bio-oil produced at two different

temperatures (330, 340 Β°C) using three different catalysts (CAT-M3, CAT-M3+, CAT-M4). As

seen in Figure 4.32, the more stable HDT-bio-oil was produced using CAT-M4, followed by CAT-

M3+ and CAT-M3. The nature of this precipitated material is still unknown but it has a similar

behavior to the asphaltenes formed when processing petroleum-derived oil. However, due to these

images, it was possible to conclude that the CS2 insoluble material filtered from the liquid samples

was not the same type of material that naturally precipitates and was observed in the microscope

images. From Figure 4.31 it was found that the CS2 insoluble material formed using CAT-M4 is

in the high region, while in Figure 4.32 there is no intrinsic solid precipitation up to 340 Β°C.

Table 4.8 shows the conversion reached at different temperatures during different reactions

and if solid precipitations were observed. It can be seen that the solids start to precipitate after a

conversion around 29 % is reached no matter the temperature of the reaction.

Figure 4.32. Microscope images for the HDT-bio-oil for different T and catalysts

77

Table 4.8. Characterization of HDT-bio-oil at different temperatures during different reactions

Reaction Catalyst Temperature [Β°C] Conversion at

343 Β°C+ [wt. %] Solid precipitation

R2 CAT-M3 325 28.7 No

R3 CAT-M3 325 28.3 No

R3 CAT-M3 330 29.9 Yes

R4 CAT-M3+ 330 23.5 No

R4 CAT-M3+ 340 28.9 Yes

R5 CAT-M4 340 28.3 No

Finally, the hydrogen consumed per gram of oil vs the DOD for different catalysts is plotted

in Figure 4.33. The hydrogen consumption shows a linear trend with the DOD for all the catalysts.

Additionally, it seems that the hydrogen consumed to process the bio-oil using the three different

catalysts follows the same pathway because the slopes from the graph are very similar. Therefore,

the amount of hydrogen consumption to reach the same degree of deoxygenation will not vary

depending on the catalyst used. However, the severity of the reaction and the products obtained at

different conditions are, as proved by the experimental results obtained, dependent on the catalyst

used.

Figure 4.33. H2 consumption vs DOD for different catalysts

y = 0.2897x - 0.0176RΒ² = 0.9993

y = 0.3041x - 0.0164RΒ² = 0.9997

y = 0.2842x - 0.1491RΒ² = 0.9879

0

5

10

15

20

25

0 20 40 60 80

H2

con

sum

pti

on

[m

g H2/g

oil]

DOD [%]

CAT-M3

CAT-M3+

CAT-M4

78

Chapter 5: Conclusions and Future Work

A summary of the key concluding remarks is presented in this section along with some

recommendations for directing the next stage of investigations in this area.

Conclusions

General trends

Some general trends were found when hydrotreating bio-oil at different conditions. First, it

was observed that increasing the total operating pressure secures a higher solubility of hydrogen

in the oil and thereby a higher availability of hydrogen in the vicinity of the catalyst, resulting in a

deeper hydrogenation of the feed. A more upgraded HDT-bio-oil was obtained at higher

temperatures and lower space velocities, i.e. a better quality HDT-bio-oil in terms of viscosity,

DOD, TAN conversion and MCR was obtained at the highest severity of reaction.

Moreover, a linear relation between the O/C and H/C ratio and the degree of deoxygenation

of the HDT-bio-oil was found, meaning that the removal of oxygen atoms from the bio-oil is linked

with the addition of hydrogen to it. Additionally, the hydrogen consumption was found to follow

a linear trend with the DOD when submitted to temperatures up to 330 Β°C. After this temperature,

it seems to follow a much faster trend with the DOD because the methanation reaction that

consumes a high amount of hydrogen is promoted at high temperatures. In addition, the hydrogen

consumption was found to be directly dependent on the severity of the reaction. These

inconveniences plus the in-stabilization of the largest molecules in the media mark the necessity

of introducing cracking reactions from the severity limit obtained for T, P and WHSV 330 Β°C,

1400 psig and 0.2 h-1, respectively, making hydro-cracking or steam-cracking the options.

Finally, it was observed that the conversion started to increase sharply after achieving 29-

30 wt. %. In addition, the solids observed in the microscope started to precipitate after reaching

the same level of conversion. Therefore, an important thermal contribution was observed in the

results of the HDT-bio-oil, specifically in the conversion at 343 Β°C+, after reaching 29-30 wt. %.

Upgrading of bio-oil via HDT

One of the main goals achieved in the present research was that it was possible to produce a

high quality HDT-bio-oil that was further processed via CSC (out of the scope of this thesis) and

79

did not deactivate the CSC catalyst.21 A viscosity reduction of 99 % and MCR reduction of 50 %

was reached by using CAT-M3, a temperature of 345 Β°C, 0.2 h-1 of space velocity and 1400 psig.

Moreover, a conversion at 343 Β°C+ of 39.5 wt. % was achieved at the same conditions. It was also

found that the residue was mainly converted to water, gases and diesel possibly because the VGO’s

rate of consumption was similar to its rate of generation. Finally, a reduction of 80 % of the CS2

insoluble material was achieved for the HDT-bio-oil.

Acidity of HDT-bio-oil

It was possible to measure the acidity of the HDT-bio-oil in terms of C=O acids and phenols

using the FTIR characterization technique. Moreover, a correlation between TAN and infrared

absorptivity at 1710-1750 cm-1 was successfully carried out. The relative error found for the results

measured with both analytical methods was 10.9 %, which is within the range of the error for the

TAN measurement. Therefore, the limitations faced by trying to measure bio-oils with low TAN

value (< 2 mgKOH/g) were overcome. The correlation developed can be used for determining the

acidity content in an inexpensive, more accurate and easier way by using only FTIR.

A C=O acids reduction of 100 %, thus a TAN conversion of 100 %, was achieved when

hydrotreating the bio-oil at 345 Β°C and using CAT-M3 (1400 psig, 0.2 h-1). On the other hand, a

phenols reduction of 28 % was reached at 340 Β°C, 0.2 h-1 and 1400 psig using CAT-M3+.

Catalysts performance comparison

The lifetime of CAT-M3 was tested in a long term run that lasted 55 days. It was found that

CAT-M3 remained constantly active throughout all the long term run, meaning that CAT-M3,

under the hydrotreating conditions tested in this research, did not observably deactivate for a period

of time 55 days.

Comparing CAT-M3 with CAT-M2, it was observed that at the same conditions, CAT-M3

performed better in terms of TAN conversion, MCR, DOD and viscosity reduction. In this way, a

higher amount of hydrogenating active phase in the catalyst enhances the bio-oil upgrading, as

long as the hydrogenating phase is well dispersed.

Moreover, higher temperatures were tested using CAT-M3, CAT-M3+ and CAT-M4. The

best catalyst found for C=O acids reduction was CAT-M3, whereas the best catalyst observed for

80

the phenols reduction was CAT-M3+. Regarding properties such as conversion, viscosity, MCR

and DOD, CAT-M3 was found to perform the best, followed by CAT-M3+ and CAT-M4.

In general, the more upgraded HDT-bio-oil was obtained using CAT-M3 with the exception

of the phenols reduction. CAT-M4 did not perform as good as the other two catalyst tested in the

present work and it might be related to the reduction of the active phases due to coverage of the

first hydrogenation phase by the second phase added or the formation of some solids that produced

active sites of much less activity than the ones formed in CAT-M3.

Future work

The novelty of the proposed upgrading approach relies on the fact that the unconsumed

hydrogen produced in CSC may be recycled back to the HDT unit, reducing or eliminating the

hydrogen make-up for the HDT unit. Thus, CSC and HDT units are supposed to be merged into

one single continuous operation. However, until now, the studies regarding upgrading

lignocellulose-derived bio-oils in the CAFE group have been carried out in separate units for HDT

and CSC. Therefore, merging the HDT and CSC units is being done, out of the scope of this thesis,

to study the whole process and its possible complications. Moreover, the CSR of the gases

produced in HDT and CSC will be explored in order to determine the maximum quantity of

hydrogen that could be produced to cope with the needs of the HDT unit. Additionally, an

economic study must be done to analyze the feasibility of the novel approach.

Regarding the optimization of the HDT process, a kinetic model should be developed to have

a better understanding of the process and to be able to predict the product quality at different

operating conditions. Finally, a catalyst that could reach the product quality as CAT-M3 should be

improved or modified to achieve a higher phenols reduction. The catalyst named as β€œ+’ in the

present research could be modified to have larger pore sizes so the bulky molecules in the bio-oil

could reach the active sites.

81

References

1. Klass, D.L., Biomass for renewable energy, fuels, and chemicals. Elsevier: 1998.

2. Pereira, P., et al., Aquaconversion technology offers added value to E. Venezuela synthetic

crude oil production. Oil & Gas Journal 2001. 99(20), 79-85.

3. Mikulec, J., KleinovÑ, A., Cvengroő, J., and Banič, M., Catalytic transformation of tall oil

into biocomponent of diesel fuel. International Journal of Chemical Engineering 2012.

4. Anthonykutty, J.M., et al., Value added hydrocarbons from distilled tall oil via

hydrotreating over a commercial NiMo catalyst. Industrial & Engineering Chemistry

Research 2013. 52(30), 10114-10125.

5. Helwani, Z., Othman, M., Aziz, N., Fernando, W., and Kim, J., Technologies for

production of biodiesel focusing on green catalytic techniques: a review. Fuel Processing

Technology 2009. 90(12), 1502-1514.

6. Huber, G.W., Iborra, S., and Corma, A., Synthesis of transportation fuels from biomass:

chemistry, catalysts, and engineering. Chemical reviews 2006. 106(9), 4044-4098.

7. Egeberg, R., Michaelsen, N., Skyum, L., and Topsoe, H., Novel hydrotreating technology

for production of green diesel. ERTC, Nov 2009, 9-11.

8. Huber, G.W. and Dumesic, J.A., An overview of aqueous-phase catalytic processes for

production of hydrogen and alkanes in a biorefinery. Catalysis Today 2006. 111(1-2), 119-

132.

9. Demirbaş, A., Calculation of higher heating values of biomass fuels. Fuel 1997. 76(5),

431-434.

10. Jensen, C.U., Guerrero, J.K.R., Karatzos, S., Olofsson, G., and Iversen, S.B., Fundamentals

of Hydrofactionβ„’: Renewable crude oil from woody biomass. Biomass Conversion and

Biorefinery 2017. 7(4), 495-509.

11. Saidi, M., et al., Upgrading of lignin-derived bio-oils by catalytic hydrodeoxygenation.

Energy & Environmental Science 2014. 7(1), 103-129.

12. Furimsky, E., Catalytic hydrodeoxygenation. Applied Catalysis A: General 2000. 199(2),

147-190.

13. Czernik, S. and Bridgwater, A., Overview of applications of biomass fast pyrolysis oil.

Energy & fuels 2004. 18(2), 590-598.

14. Mohan, D., Pittman, C.U., and Steele, P.H., Pyrolysis of wood/biomass for bio-oil: a

critical review. Energy & fuels 2006. 20(3), 848-889.

15. Elliott, D.C., Historical developments in hydroprocessing bio-oils. Energy & Fuels 2007.

21(3), 1792-1815.

16. Zhang, Q., Chang, J., Wang, T., and Xu, Y., Review of biomass pyrolysis oil properties

and upgrading research. Energy conversion and management 2007. 48(1), 87-92.

17. Mortensen, P.M., Grunwaldt, J.-D., Jensen, P.A., Knudsen, K., and Jensen, A.D., A review

of catalytic upgrading of bio-oil to engine fuels. Applied Catalysis A: General 2011. 407(1-

2), 1-19.

18. Bu, Q., et al., A review of catalytic hydrodeoxygenation of lignin-derived phenols from

biomass pyrolysis. Bioresource Technology 2012. 124, 470-477.

19. Elliott, D.C. and Baker, E., Upgrading biomass liquefaction products through

hydrodeoxygenation. 1984, Pacific Northwest Lab., Richland, WA (USA).

82

20. Trujillo-Ferrer, G. Thermal and Catalytic Steam Reactivity Evaluation of Athabasca

Vacuum Gasoil. University of Calgary, Calgary, AB, Canada, 2008

21. Trujillo, M., Catalytic Upgrading Process of Ligno-cellulose Derived Heavy Crude Oil.

2018, University of Calgary: Calgary.

22. StΓΆcker, M., Biofuels and biomass‐to‐liquid fuels in the biorefinery: Catalytic conversion

of lignocellulosic biomass using porous materials. Angewandte Chemie International

Edition 2008. 47(48), 9200-9211.

23. Kanaujia, P.K., Sharma, Y., Garg, M., Tripathi, D., and Singh, R., Review of analytical

strategies in the production and upgrading of bio-oils derived from lignocellulosic biomass.

Journal of Analytical and Applied Pyrolysis 2014. 105, 55-74.

24. Eisentraut, A., Sustainable production of second-generation biofuels. 2010.

25. ENC, P. Database for biomass and waste. 2012 18 July 2018]; Available from:

https://www.ecn.nl/phyllis2/.

26. Deguchi, S., Tsujii, K., and Horikoshi, K., Cooking cellulose in hot and compressed water.

Chemical Communications 2006(31), 3293-3295.

27. Quitain, A.T., Sato, N., Daimon, H., and Fujie, K., Qualitative investigation on

hydrothermal treatment of hinoki (Chamaecyparis obtusa) bark for production of useful

chemicals. Journal of agricultural and food chemistry 2003. 51(27), 7926-7929.

28. Soltes, E.J., Elder, T., and Goldstein, I., Organic Chemicals from Biomass. 1981, CRC

Press, Pyrolysis.

29. Nimz, H., Chemistry of potential chromophoric groups in beech lignin. Tappi Tech Ass

Pulp Pap Indus 1973.

30. Bauen, A., in Encyclopedia of Energy, Elsevier, Editor. 2004: Amsterdam.

31. SjΓΆstrΓΆm, E., Wood Chemistryβ€”Fundamentals and Applications Academic. New York

1993.

32. Bridgwater, A., Catalysis in thermal biomass conversion. Applied Catalysis A: General

1994. 116(1-2), 5-47.

33. Zhu, Z., Rosendahl, L., Toor, S.S., Yu, D., and Chen, G., Hydrothermal liquefaction of

barley straw to bio-crude oil: effects of reaction temperature and aqueous phase

recirculation. Applied Energy 2015. 137, 183-192.

34. Brand, S. and Kim, J., Liquefaction of major lignocellulosic biomass constituents in

supercritical ethanol. Energy 2015. 80, 64-74.

35. DEMIRBAŞ, A., Balat, M., and BOZBAŞ, K., Direct and catalytic liquefaction of wood

species in aqueous solution. Energy Sources 2005. 27(3), 271-277.

36. Peterson, A.A., et al., Thermochemical biofuel production in hydrothermal media: a review

of sub-and supercritical water technologies. Energy & Environmental Science 2008. 1(1),

32-65.

37. Kruse, A. and Dinjus, E., Hot compressed water as reaction medium and reactant:

properties and synthesis reactions. The Journal of supercritical fluids 2007. 39(3), 362-

380.

38. Lu, Q., Li, W.-Z., and Zhu, X.-F., Overview of fuel properties of biomass fast pyrolysis

oils. Energy Conversion and Management 2009. 50(5), 1376-1383.

39. Milne, T., Agblevor, F., Davis, M., Deutch, S., and Johnson, D., A review of the chemical

composition of fast-pyrolysis oils from biomass, in Developments in thermochemical

biomass conversion. 1997, Springer. p. 409-424.

83

40. Banerjee, D.K., Oil sands, heavy oil & bitumen: from recovery to refinery. PennWell

Books: 2012.

41. Branca, C., Giudicianni, P., and Di Blasi, C., GC/MS characterization of liquids generated

from low-temperature pyrolysis of wood. Industrial & Engineering Chemistry Research

2003. 42(14), 3190-3202.

42. Pedersen, T.H. and Rosendahl, L.A., Production of fuel range oxygenates by supercritical

hydrothermal liquefaction of lignocellulosic model systems. Biomass and Bioenergy 2015.

83, 206-215.

43. Carrier, M., et al., Conversion of fern (Pteris vittata L.) biomass from a phytoremediation

trial in sub-and supercritical water conditions. biomass and bioenergy 2011. 35(2), 872-

883.

44. Carrier, M., Loppinet-Serani, A., Absalon, C., Aymonier, C., and Mench, M., Degradation

pathways of holocellulose, lignin and Ξ±-cellulose from Pteris vittata fronds in sub-and

super critical conditions. Biomass and bioenergy 2012. 43, 65-71.

45. Toor, S.S., Rosendahl, L., and Rudolf, A., Hydrothermal liquefaction of biomass: a review

of subcritical water technologies. Energy 2011. 36(5), 2328-2342.

46. Oasmaa, A. and Czernik, S., Fuel oil quality of biomass pyrolysis oils state of the art for

the end users. Energy & Fuels 1999. 13(4), 914-921.

47. Soltes, J. and Lin, S.-C.K., Hydroprocessing of biomass tars for liquid engine fuels, in

Progress in biomass conversion. 1984, Elsevier. p. 1-68.

48. SipilΓ€, K., Kuoppala, E., FagernΓ€s, L., and Oasmaa, A., Characterization of biomass-based

flash pyrolysis oils. Biomass and Bioenergy 1998. 14(2), 103-113.

49. Boelhouwer, J., Nederbragt, G., and Verberg, G., Viscosity data of organic liquids. Applied

Scientific Research 1951. 2(1), 249.

50. Schaschke, C., A dictionary of chemical engineering. OUP Oxford: 2014.

51. Patel, M. and Kumar, A., Production of renewable diesel through the hydroprocessing of

lignocellulosic biomass-derived bio-oil: A review. Renewable and Sustainable Energy

Reviews 2016. 58, 1293-1307.

52. Donnis, B., Egeberg, R.G., Blom, P., and Knudsen, K.G., Hydroprocessing of bio-oils and

oxygenates to hydrocarbons. Understanding the reaction routes. Topics in Catalysis 2009.

52(3), 229-240.

53. Elliott, D.C. and Hart, T.R., Catalytic hydroprocessing of chemical models for bio-oil.

Energy & Fuels 2008. 23(2), 631-637.

54. Xiu, S. and Shahbazi, A., Bio-oil production and upgrading research: A review. Renewable

and Sustainable Energy Reviews 2012. 16(7), 4406-4414.

55. Chiaramonti, D., et al., Development of emulsions from biomass pyrolysis liquid and diesel

and their use in enginesβ€”Part 1: emulsion production. Biomass and bioenergy 2003. 25(1),

85-99.

56. Chiaramonti, D., et al., Development of emulsions from biomass pyrolysis liquid and diesel

and their use in enginesβ€”Part 2: tests in diesel engines. Biomass and Bioenergy 2003.

25(1), 101-111.

57. Ikura, M., Stanciulescu, M., and Hogan, E., Emulsification of pyrolysis derived bio-oil in

diesel fuel. Biomass and bioenergy 2003. 24(3), 221-232.

84

58. Garcia-Perez, M., Adams, T.T., Goodrum, J.W., Geller, D.P., and Das, K., Production and

fuel properties of pine chip bio-oil/biodiesel blends. Energy & Fuels 2007. 21(4), 2363-

2372.

59. Baker, E. and Elliott, D., Catalytic upgrading of biomass pyrolysis oils, in Research in

thermochemical biomass conversion. 1988, Springer. p. 883-895.

60. Baldauf, W., Balfanz, U., and Rupp, M., Upgrading of flash pyrolysis oil and utilization in

refineries. Biomass and Bioenergy 1994. 7(1-6), 237-244.

61. Venderbosch, R., Ardiyanti, A., Wildschut, J., Oasmaa, A., and Heeres, H., Stabilization

of biomass‐derived pyrolysis oils. Journal of Chemical Technology & Biotechnology 2010.

85(5), 674-686.

62. Kwon, K.C., Mayfield, H., Marolla, T., Nichols, B., and Mashburn, M., Catalytic

deoxygenation of liquid biomass for hydrocarbon fuels. Renewable Energy 2011. 36(3),

907-915.

63. Tang, Z., Lu, Q., Zhang, Y., Zhu, X., and Guo, Q., One step bio-oil upgrading through

hydrotreatment, esterification, and cracking. Industrial & Engineering Chemistry Research

2009. 48(15), 6923-6929.

64. Wildschut, J., Mahfud, F.H., Venderbosch, R.H., and Heeres, H.J., Hydrotreatment of fast

pyrolysis oil using heterogeneous noble-metal catalysts. Industrial & Engineering

Chemistry Research 2009. 48(23), 10324-10334.

65. Landberg, K. Experimental and kinetic modelling study of hydrodeoxygenation of tall oil

to renewable fuel. Master Thesis, Chalmers University of Technology, Gothenburg,

Sweden, 2017

66. Grange, P., Laurent, E., Maggi, R., Centeno, A., and Delmon, B., Hydrotreatment of

pyrolysis oils from biomass: reactivity of the various categories of oxygenated compounds

and preliminary techno-economical study. Catalysis today 1996. 29(1-4), 297-301.

67. Weisser, O. and Landa, S., Sulphide catalysts, their properties and applications. Elsevier:

2013.

68. Elliott, D.C., Hart, T.R., Neuenschwander, G.G., Rotness, L.J., and Zacher, A.H., Catalytic

hydroprocessing of biomass fast pyrolysis bio‐oil to produce hydrocarbon products.

Environmental Progress & Sustainable Energy: An Official Publication of the American

Institute of Chemical Engineers 2009. 28(3), 441-449.

69. Cabrales-Navarro, F.A. Reactivity Study and Kinetic Modeling of Deasphalted Oil

Upgrading via Thermal and Catalytic Cracking. PhD, University of Calgary, 2016

70. Vitale, G., Frauwallner, M., Scott, C., and Pereira-Almao, P., Preparation and

characterization of low-temperature nano-crystalline cubic molybdenum carbides and

insights on their structures. Applied Catalysis A: General 2011. 408(1-2), 178-186.

71. Speight, J.G., High acid crudes. Gulf Professional Publishing: 2014.

72. ASTM Committee D-2 on Petroleum Products, L.F. and Lubricants, Standard test method

for acid number of petroleum products by potentiometric titration. ASTM International:

2011.

73. Carbognani, L., et al., Reliable determination of water contents of bitumen and vacuum

residua via coulometric Karl Fischer titration using tetrahydrofuran. Petroleum Science

and Technology 2014. 32(5), 602-609.

74. D, A., Standard Test Method for Water in Crude Oils by Coulometric Karl Fischer

Titration. Annual Book of Standards 2004.

85

75. Gray, M.R., Upgrading oilsands bitumen and heavy oil. University of Alberta: 2015.

76. D7169-05, A., Standard Test Method for Boiling Range Distribution of Samples with

Residues Such as Crude Oils and Atmospheric and Vacuum Residues by High Temperature

Gas Chromatography,. Annual Book of ASTM Standards. 5.

77. Carbognani, L., Lubkowitz, J., Gonzalez, M.F., and Pereira-Almao, P., High temperature

simulated distillation of Athabasca vacuum residue fractions. Bimodal distributions and

evidence for secondary β€œon-column” cracking of heavy hydrocarbons. Energy & Fuels

2007. 21(5), 2831-2839.

78. Rodriguez-DeVecchis, V.M., Carbognani Ortega, L., Scott, C.E., and Pereira-Almao, P.,

Use of Nanoparticle Tracking Analysis for Particle Size Determination of Dispersed

Catalyst in Bitumen and Heavy Oil Fractions. Industrial & Engineering Chemistry

Research 2015. 54(40), 9877-9886.

79. D5291, A., Standard Test Methods for Intrumental Determination of Carbon, Hydrogen

and Nitrogen in Petroleum Products and Lubricants. Annual Book of ASTM Standards

2002. 05.

80. Hassan, A., Carbognani, L., and Pereira-Almao, P., Development of an alternative setup

for the estimation of microcarbon residue for heavy oil and fractions: Effects derived from

air presence. Fuel 2008. 87(17-18), 3631-3639.

81. Venyaminov, S.Y. and Prendergast, F.G., Water (H2O and D2O) Molar Absorptivity in

the 1000–4000 cmβˆ’ 1Range and Quantitative Infrared Spectroscopy of Aqueous Solutions.

Analytical biochemistry 1997. 248(2), 234-245.

82. Silverstein, R.M., Webster, F.X., Kiemle, D.J., and Bryce, D.L., Spectrometric

identification of organic compounds. John wiley & sons: 2014.

83. Barcenas, M., Orea, P., Buenrostro-GonzΓ‘lez, E., Zamudio-Rivera, L.S., and Duda, Y.,

Study of medium effect on asphaltene agglomeration inhibitor efficiency. Energy & Fuels

2008. 22(3), 1917-1922.

84. Guo, X., Zheng, Y., Zhang, B., and Chen, J., Analysis of coke precursor on catalyst and

study on regeneration of catalyst in upgrading of bio-oil. Biomass and bioenergy 2009.

33(10), 1469-1473.

85. Carbognani Ortega, L.C., Josune; Pereira Almao, Pedro, Correlation of Thermogravimetry

and High Temperature Simulated Distillation for Oil Analysis: Thermal Cracking

Influence over both Methodologies, in The Boduszynski Continuum: Contributions to the

Understanding of the Molecular Composition of Petroleum. 2018, ACS Symposium

Series: Washington, DC.

86

Appendix I: Modifications of RTU-1

Figure 1 shows the RTU-1 diagram as designed and constructed by Cabrales Navarro before

modifications.69 The modifications made in the present research were done due to some limitations

found for some tanks at the high operating pressure used for hydrotreating bio-oil. In Figure 1, it

can be noticed that MB Tank 1 was used to separate the light products and gases from the heavy

products. First, the lights were passed through a condenser and the gases were separated (after

passing through Liquid Trap 1) and sent to the GC or exhaust. The light products were retained in

MB Tank 3 until the mass balance time was reached. Then, the pneumatic valves from the

automated sampling system opened, and the lights could be collected.

Figure 1. RTU-1 diagram before modifications69

87

Taking a look at Figure 3.1, it is noticed that the main modifications in RTU-1 were made in

the light products separation and collecting zone. The reason behind the alterations was that MB

Tank 1 presented some limitations of temperature when submitted to a pressure of 1400 psig.

Hence, the separation of light from heavy products could not be done in MB Tank 1. Additionally,

a N2 mass controller was used for bubbling in MB Tank 2.

The main modification done to the system was to change the separation system from MB Tank

1 which is submitted to high pressures (1400 psig) to MB Tank 2, which after the automated

sampling system is submitted to atmospheric pressure. For this purpose, a long line was connected

between MB Tank 2 and MB Tank 3 to serve as a condenser. Moreover, N2 bubbling was added

to MB Tank 2 to enhance the separation between the light and heavy products. Nitrogen flushing

lines were connected to MB Tank 2 and 3 for better collection of the samples. Finally, some lines

were added to send the remaining gases to the exhaust system.

88

Appendix II: Operational Data and Experimental Results

Table 1. Pilot plant data and operational conditions for reaction #1

RUN DATA Reaction #1

C1 C2 C2 C3 C4

Mass Balance MB1 MB1 MB3 MB1 MB2

Pressure [psig] 1400 1200 1200 1100 900

Reaction Temperature [Β°C] 310 310 310 310 310

WHSV [h-1] 0.2 0.2 0.2 0.2 0.2

Residence Time [min] 300 300 300 300 300

Mass Balance Time [min] 1385 1440 1440 1440 1440

Time on Stream [h] 53.1 202.3 250.3 466.3 322.3

Hydrocarbon Feed Flow [cc/min] 0.0395 0.0395 0.0395 0.0395 0.0395

Hydrogen to Oil ratio [v/v] 900 900 900 900 900

Hydrogen Feed Flow [sccm] 35.58 35.58 35.58 35.58 35.58

Total Liquid Hydrocarbon Feed [g] 54.76 59.21 59.21 59.21 59.21

Catalyst Matrix CAT-M2 CAT-M2 CAT-M2 CAT-M2 CAT-M2

LIQUID PRODUCTS Light Product - Light Collector [g] 2.65 4.36 4.21 4.20 4.03

Heavy Product [g] 50.04 52.06 53.52 52.45 53.87

Light Hydrocarbon Product [g] 0.56 1.07 1.23 0.40 1.06

Heavy Hydrocarbon Product [g] 49.53 51.82 53.26 51.34 53.65

Total Liquid Hydrocarbon Product [g] 50.09 52.89 54.48 51.74 54.71

Water in Light Collector [g] 2.08 3.29 2.99 3.80 2.97

Water in Total Liquid HC Product [g] 0.52 0.24 0.27 1.10 0.22

Total Water Collected [g] 2.60 3.53 3.25 4.90 3.20

% Light HC Product 1.12 2.02 2.25 0.78 1.93

% Heavy HC Product 98.88 97.98 97.75 99.22 98.07

GAS PRODUCTS Measured Gas Flow [sccm] 38.98 38.97 38.97 38.60 37.83

Hydrogen Content in Gas [% v/v] 97.70 97.90 97.90 98.10 98.10

Gas HC Product Flow [sccm] 0.90 0.82 0.82 0.73 0.72

Gas HC Product [g] 2.33 2.28 2.28 2.01 1.91

SOLID PRODUCTS CS2 Insolubles [g] 2.53 2.66 2.74 - -

TOTAL PRODUCTS Total Hydrocarbon Product [g] 52.42 55.17 56.76 53.75 56.62

USEFUL CALCULATIONS Overall Mass Balance [%] 100.5 99.1 101.4 99.0 101.0

Hydrocarbon Yield [%] 95.7 93.2 95.9 90.8 95.6

Water Yield [%] 4.7 6.0 5.4 8.3 5.3

Gas Yield [%] 4.2 3.9 3.8 3.4 3.2

Liquid Yield [%] 91.2 91.6 91.6 96.6 96.8

Liquid HC Yield [%] 86.5 85.6 86.2 88.2 91.5

CS2 Insolubles Yield [%] 4.6 4.5 4.6 - -

89

Table 2. Pilot plant data and operational conditions for reaction #2

RUN DATA Reaction #2

C1 C2 C3 C4 C5 C6

Mass Balance MB4 MB4 MB1 MB2 MB1 MB3

Pressure [psig] 1400 1400 1400 1400 1400 1400

Reaction Temperature [Β°C] 310 320 315 325 310 300

WHSV [h-1] 0.2 0.2 0.2 0.2 0.2 0.2

Residence Time [min] 300 300 300 300 300 300

Mass Balance Time [min] 720 720 720 720 720 720

Time on Stream [h] 78.0 294.0 546.0 714.0 798.0 952.0

Hydrocarbon Feed Flow [cc/min] 0.0878 0.0878 0.0878 0.0878 0.0878 0.0878

Hydrogen to Oil ratio [v/v] 1150 1150 1150 1150 1150 1150

Hydrogen Feed Flow [sccm] 101.00 100.97 100.97 100.97 100.97 101.00

Total Liquid Hydrocarbon Feed [g] 65.77 65.74 65.74 65.74 65.74 65.77

Catalyst Matrix CAT-M3 CAT-M3 CAT-M3 CAT-M3 CAT-M3 CAT-M3

LIQUID PRODUCTS Light Product - Light Collector [g] 4.02 4.02 6.34 4.94 4.53 3.97

Heavy Product [g] 58.42 58.42 56.31 56.31 57.95 59.96

Light Hydrocarbon Product [g] 0.80 0.80 1.65 1.41 1.17 0.98

Heavy Hydrocarbon Product [g] 57.77 58.16 56.13 56.07 57.74 59.56

Total Liquid Hydrocarbon Product [g] 58.57 58.96 57.78 57.48 58.90 60.53

Water in Light Collector [g] 3.22 3.22 4.69 3.52 3.36 3.00

Water in Total Liquid HC Product [g] 0.66 0.26 0.18 0.25 0.22 0.41

Total Water Collected [g] 3.88 3.48 4.87 3.77 3.58 3.40

% Light HC Product 1.36 1.35 2.86 2.46 1.98 1.61

% Heavy HC Product 98.64 98.65 97.14 97.54 98.02 98.39

GAS PRODUCTS Measured Gas Flow [sccm] 88.82 88.11 86.52 83.23 88.82 89.36

Hydrogen Content in Gas [% v/v] 96.90 96.60 97.30 97.00 96.90 98.00

Gas HC Product Flow [sccm] 2.75 3.00 2.34 2.50 2.75 1.79

Gas HC Product [g] 3.73 3.98 3.27 3.45 3.73 2.37

SOLID PRODUCTS CS2 Insolubles [g] 2.61 2.90 2.96 2.60 2.60 4.08

TOTAL PRODUCTS Total Hydrocarbon Product [g] 62.29 62.95 61.05 60.93 62.63 62.91

USEFUL CALCULATIONS Overall Mass Balance [%] 100.6 101.0 100.3 98.4 100.7 100.8

Hydrocarbon Yield [%] 94.7 95.7 92.9 92.7 95.3 95.7

Water Yield [%] 5.9 5.2 7.4 5.8 5.4 5.1

Gas Yield [%] 5.6 6.0 5.0 5.3 5.6 3.6

Liquid Yield [%] 90.4 89.6 90.5 90.7 90.4 90.3

Liquid HC Yield [%] 84.6 84.4 83.2 84.8 85.0 85.1

CS2 Insolubles Yield [%] 3.9 4.4 4.5 4.0 3.9 6.2

90

Table 3. Pilot plant data and operational conditions for reaction #2

RUN DATA Reaction #2

C7 C8 C9 C10 C11 C12

Mass Balance MB1 MB1 MB2 MB1 MB3 MB1

Pressure [psig] 1400 1400 1400 1400 1400 1400

Reaction Temperature [Β°C] 295 315 315 315 320 310

WHSV [h-1] 0.2 0.3 0.4 0.5 0.3 0.2

Residence Time [min] 300 200 150 120 200 300

Mass Balance Time [min] 720 960 360 300 480 720

Time on Stream [h] 1061.0 1127.0 1183.0 1200.0 1247.0 1318.0

Hydrocarbon Feed Flow [cc/min] 0.0878 0.1317 0.1757 0.2196 0.1317 0.0878

Hydrogen to Oil ratio [v/v] 1150 1150 1150 1150 1150 1150

Hydrogen Feed Flow [sccm] 101.00 151.46 202.06 252.54 151.46 100.97

Total Liquid Hydrocarbon Feed [g] 65.77 131.49 65.78 68.52 65.74 65.74

Catalyst Matrix CAT-M3 CAT-M3 CAT-M3 CAT-M3 CAT-M3 CAT-M3

LIQUID PRODUCTS Light Product - Light Collector [g] 4.83 11.13 4.09 3.52 4.01 3.83

Heavy Product [g] 59.43 114.53 59.78 61.29 59.96 57.19

Light Hydrocarbon Product [g] 3.84 6.87 1.33 2.11 1.29 1.90

Heavy Hydrocarbon Product [g] 59.23 113.51 59.17 59.62 59.61 55.99

Total Liquid Hydrocarbon Product [g] 63.06 120.38 60.50 61.73 60.90 57.90

Water in Light Collector [g] 1.00 4.26 2.76 1.41 2.72 1.92

Water in Total Liquid HC Product [g] 0.21 1.07 0.62 1.68 0.35 1.21

Total Water Collected [g] 1.21 5.33 3.38 3.09 3.07 3.13

% Light HC Product 6.08 5.71 2.20 3.41 2.12 3.29

% Heavy HC Product 93.92 94.29 97.80 96.59 97.88 96.71

GAS PRODUCTS Measured Gas Flow [sccm] 92.42 132.32 180.61 227.57 131.98 87.49

Hydrogen Content in Gas [% v/v] 98.00 97.70 97.70 97.70 97.70 97.70

Gas HC Product Flow [sccm] 1.85 3.04 4.15 5.23 3.04 2.01

Gas HC Product [g] 2.48 5.11 2.66 2.62 2.37 2.72

SOLID PRODUCTS CS2 Insolubles [g] 3.04 12.87 5.01 5.07 4.35 2.60

TOTAL PRODUCTS Total Hydrocarbon Product [g] 65.54 125.49 63.16 64.35 63.27 60.62

USEFUL CALCULATIONS Overall Mass Balance [%] 101.5 99.5 101.1 98.4 100.9 97.0

Hydrocarbon Yield [%] 99.7 95.4 96.0 93.9 96.2 92.2

Water Yield [%] 1.8 4.1 5.1 4.6 4.6 4.9

Gas Yield [%] 3.7 3.9 4.0 3.9 3.6 4.3

Liquid Yield [%] 91.7 86.3 88.5 88.6 89.9 91.6

Liquid HC Yield [%] 89.9 82.2 83.4 84.0 85.2 86.7

CS2 Insolubles Yield [%] 4.6 9.8 7.5 7.5 6.6 4.1

91

Table 4.SimDist raw data and calculated results for reaction #2

RUN DATA FEED Reaction #2

C1 C2 C3 C4 C6 C7

Pressure [psig] 1400 1400 1400 1400 1400 1400

Reaction Temperature [Β°C] 310 320 315 325 300 295

WHSV [h-1] 0.2 0.2 0.2 0.2 0.2 0.2

Catalyst Matrix CAT-M3 CAT-M3 CAT-M3 CAT-M3 CAT-M3 CAT-M3

Water Yield [%] 5.6 5.7 6.6 6.0 5.2 2.3

Gas Yield [%] 5.5 5.8 5.0 5.3 3.6 3.8

Naphtha Yield (28 - 190 Β°C) [%] 2.1 3.2 4.3 3.5 2.7 3.2 3.2

Jet Fuel Yield (190 - 260 Β°C) [%] 5.2 8.1 8.5 7.8 7.8 7.1 7.4

Diesel Yield (260 - 343 Β°C) [%] 10.5 16.6 18.1 16.6 20.0 14.1 13.8

VGO Yield (343 - 550 Β°C) [%] 26.5 27.1 26.8 27.1 24.7 25.4 27.6

Residue Yield (550Β°C+) [%] 55.8 33.6 30.8 33.4 33.6 41.3 42.0

CS2 Insolubles Yield [%] 3.9 4.4 4.5 4.0 5.6 4.5

Conversion (343 Β°C+) [%] 0.0 25.5 29.6 26.0 28.7 18.8 15.3

SIMULATED DISTILLATION

0 162.5 125.25 115.70 119.63 166.50 165.95 165.70

5 225.7 203.80 188.35 198.97 210.00 206.35 207.50

10 267.9 238.70 228.45 236.37 243.55 242.30 245.20

15 301.5 271.95 258.80 270.43 277.10 280.90 283.90

20 333.3 301.40 289.80 300.13 301.25 306.55 309.75

25 356.2 315.90 311.45 315.53 311.00 328.00 334.50

30 368.1 332.05 319.15 331.07 324.80 347.80 351.30

35 387.9 351.05 338.45 348.27 339.50 369.20 378.25

40 419.6 372.00 354.40 367.90 356.55 405.50 415.60

45 458.1 404.20 378.45 401.17 388.05 440.85 451.60

50 503.3 438.55 414.55 435.63 426.20 484.80 492.95

55 561.3 471.95 444.85 467.60 462.90 519.70 524.10

60 633.4 510.20 479.75 504.50 504.45 564.10 567.30

65 0.0 552.70 518.25 544.43 548.90 615.35 614.20

70 0.0 600.05 564.50 590.93 598.80 674.65 667.90

75 0.0 653.05 615.40 641.37 656.50 704.70 0.00

80 0.0 708.90 675.80 0.00 0.00 0.00 0.00

85 0.0 0.00 0.00 0.00 0.00 0.00 0.00

90 0.0 0.00 0.00 0.00 0.00 0.00 0.00

95 0.0 0.00 0.00 0.00 0.00 0.00 0.00

100 0.0 0.00 0.00 0.00 0.00 0.00 0.00

92

Table 5. Pilot plant data and operational conditions for reaction #3

RUN DATA Reaction #3

C2 C3 C4 C5 C6 C7

Mass Balance MB1 MB2 MB3 MB3 MB1 MB2

Pressure [psig] 1400 1400 1400 1400 1400 1400

Reaction Temperature [Β°C] 325 310 330 335 340 345

WHSV [h-1] 0.2 0.2 0.2 0.2 0.2 0.2

Residence Time [min] 300 300 300 300 300 300

Mass Balance Time [min] 720 720 720 720 720 720

Time on Stream [h] 141.5 213.5 297.5 417.5 453.5 525.5

Hydrocarbon Feed Flow [cc/min] 0.0976 0.0976 0.0976 0.0976 0.0976 0.0976

Hydrogen to Oil ratio [v/v] 900 1150 1150 1150 1150 1150

Hydrogen Feed Flow [sccm] 87.84 112.24 112.24 112.24 112.24 87.84

Total Liquid Hydrocarbon Feed [g] 70.27 73.08 73.08 70.27 70.27 70.27

Catalyst Matrix CAT-M3 CAT-M3 CAT-M3 CAT-M3 CAT-M3 CAT-M3

LIQUID PRODUCTS Light Product - Light Collector [g] 5.79 4.00 6.76 7.18 7.23 8.22

Heavy Product [g] 61.78 66.21 63.19 59.85 60.92 58.41

Light Hydrocarbon Product [g] 1.22 1.02 1.70 2.66 2.65 2.54

Heavy Hydrocarbon Product [g] 61.36 65.73 62.79 59.52 59.96 57.35

Total Liquid Hydrocarbon Product [g] 62.58 66.75 64.49 62.18 62.61 59.89

Water in Light Collector [g] 4.57 2.98 5.05 4.52 4.58 5.67

Water in Total Liquid HC Product [g] 0.43 0.49 0.41 0.35 1.00 1.08

Total Water Collected [g] 5.00 3.47 5.46 4.87 5.57 6.76

% Light HC Product 1.95 1.53 2.64 4.28 4.23 4.24

% Heavy HC Product 98.05 98.47 97.36 95.72 95.77 95.76

GAS PRODUCTS Measured Gas Flow [sccm] 81.92 91.59 91.15 94.85 92.01 92.06

Hydrogen Content in Gas [% v/v] 97.70 97.70 97.70 97.70 97.70 97.70

Gas HC Product Flow [sccm] 1.88 2.11 2.10 2.18 2.12 2.12

Gas HC Product [g] 2.35 2.76 2.60 2.62 2.48 2.38

SOLID PRODUCTS CS2 Insolubles [g] 3.21 6.13 5.03 2.37 2.22 2.72

TOTAL PRODUCTS Total Hydrocarbon Product [g] 64.93 69.51 67.09 64.80 65.09 62.27

USEFUL CALCULATIONS Overall Mass Balance [%] 99.5 99.9 99.3 99.1 100.6 98.2

Hydrocarbon Yield [%] 92.4 95.1 91.8 92.2 92.6 88.6

Water Yield [%] 7.2 4.7 7.5 7.0 7.9 9.8

Gas Yield [%] 3.4 3.8 3.6 3.8 3.5 3.4

Liquid Yield [%] 92.0 87.8 89.5 92.8 93.3 92.6

Liquid HC Yield [%] 84.9 83.1 82.0 85.8 85.5 82.8

CS2 Insolubles Yield [%] 4.6 8.4 6.9 3.4 3.1 3.9

93

Table 6.SimDist raw data and calculated results for reaction #3

RUN DATA FEED Reaction #3

C2 C4 C5 C6 C7

Pressure [psig] 1400 1400 1400 1400 1400

Reaction Temperature [Β°C] 325 330 335 340 345

WHSV [h-1] 0.2 0.2 0.2 0.2 0.2

Catalyst Matrix CAT-M3 CAT-M3 CAT-M3 CAT-M3 CAT-M3

Water Yield [%] 5.8 6.8 7.5 8.8 8.6

Gas Yield [%] 3.4 3.6 3.8 3.5 3.4

Naphta Yield (28 - 190 Β°C) [%] 2.1 3.2 3.6 4.2 4.0 5.1

Jet FuelYield (190 - 260 Β°C) [%] 5.2 8.1 7.9 9.2 9.3 9.4

Diesel Yield (260 - 343 Β°C) [%] 10.5 20.5 20.5 22.7 23.8 23.7

VGO Yield (343 - 550 Β°C) [%] 26.5 25.3 23.9 27.1 26.8 27.7

Residue Yield (550Β°C+) [%] 55.8 33.7 33.8 25.5 24.1 22.1

CS2 Insolubles Yield [%] 4.6 7.0 3.4 3.1 3.9

Conversion (343 Β°C+) [%] 0.0 28.3 29.9 36.0 38.1 39.5

SIMULATED DISTILLATION 0 162.5 166.40 165.45 165.30 166.77 165.05

5 225.7 204.95 197.20 191.60 192.80 182.50

10 267.9 239.45 234.45 227.60 227.30 221.85

15 301.5 271.50 266.70 256.55 256.50 247.35

20 333.3 296.35 292.65 283.45 281.90 274.85

25 356.2 309.45 307.50 302.75 301.20 296.10

30 368.1 320.05 315.85 310.75 308.60 308.90

35 387.9 336.60 332.30 322.10 319.00 314.35

40 419.6 351.65 348.05 336.85 333.17 328.90

45 458.1 378.95 368.30 351.65 348.00 341.45

50 503.3 414.25 403.60 374.30 365.97 356.10

55 561.3 444.25 431.85 408.15 398.70 382.40

60 633.4 491.35 478.35 435.90 427.57 415.45

65 0.0 537.00 525.40 471.70 461.83 441.10

70 0.0 591.10 576.55 513.45 502.87 477.55

75 0.0 652.00 637.40 560.05 547.70 519.25

80 0.0 701.35 695.05 615.00 598.93 565.60

85 0.0 0.00 0.00 0.00 0.00 0.00

90 0.0 0.00 0.00 0.00 0.00 0.00

95 0.0 0.00 0.00 0.00 0.00 0.00

100 0.0 0.00 0.00 0.00 0.00 0.00

94

Table 7. Pilot plant data and operational conditions for reaction #5 and #6

RUN DATA Reaction #4 Reaction #5

C1 C2 C3 C1 C2 C3

Mass Balance MB3 MB2 MB2 MB2 MB2 MB1

Pressure [psig] 1400 1400 1400 1400 1400 1400

Reaction Temperature [Β°C] 330 340 335 330 335 340

WHSV CAT-M3 [h-1] 0.2 0.2 0.2 0.2 0.2 0.2

WHSV CAT-+ [h-1] 0.5 0.5 0.5 - - -

Residence Time [min] 300 300 300 300 300 300

Mass Balance Time [min] 720 720 720 720 720 720

Time on Stream [h] 66.0 126.0 186.0 49.0 85.0 121.0

Hydrocarbon Feed Flow [cc/min] 0.0691 0.0691 0.0691 0.0927 0.0927 0.0927

Hydrogen to Oil ratio [v/v] 1150 1150 1150 1150 1150 1150

Hydrogen Feed Flow [sccm] 79.47 79.47 79.47 106.61 106.61 106.61

Total Liquid Hydrocarbon Feed [g] 49.75 49.75 49.75 66.74 66.74 66.74

Catalyst Matrix CAT-M3+ CAT-M3+ CAT-M3+ CAT-M4 CAT-M4 CAT-M4

LIQUID PRODUCTS

Light Product - Light Collector [g] 4.95 6.04 5.39 4.06 4.56 5.66

Heavy Product [g] 43.51 41.93 40.45 58.32 61.17 57.69

Light Hydrocarbon Product [g] 1.76 2.06 2.21 0.99 1.26 1.25

Heavy Hydrocarbon Product [g] 43.42 41.82 40.33 57.94 60.98 57.54

Total Liquid Hydrocarbon Product [g] 45.18 43.88 42.54 58.93 62.24 58.78

Water in Light Collector [g] 3.18 3.98 3.17 3.08 3.30 4.41

Water in Total Liquid HC Product [g] 0.09 0.12 0.13 0.38 0.19 0.16

Total Water Collected [g] 3.27 4.10 3.30 3.46 3.49 4.57

% Light HC Product 3.90 4.69 5.20 1.67 2.02 2.12

% Heavy HC Product 96.10 95.31 94.80 98.33 97.98 97.88

GAS PRODUCTS

Measured Gas Flow [sccm] 54.30 65.72 60.64 104.41 103.24 104.41

Hydrogen Content in Gas [% v/v] 97.70 97.70 97.70 97.70 97.70 97.70

Gas HC Product Flow [sccm] 1.25 1.51 1.39 2.40 2.37 2.40

Gas HC Product [g] 1.55 1.74 1.61 2.77 2.74 2.77

SOLID PRODUCTS

CS2 Insolubles [g] 3.72 2.56 3.00 5.98 6.62 5.41

TOTAL PRODUCTS

Total Hydrocarbon Product [g] 46.73 45.61 44.15 61.70 64.98 61.55

USEFUL CALCULATIONS

Overall Mass Balance [%] 100.5 99.9 95.4 97.6 102.6 99.1

Hydrocarbon Yield [%] 93.9 91.7 88.7 92.4 97.4 92.2

Water Yield [%] 6.5 8.2 7.0 5.3 5.1 6.9

Gas Yield [%] 3.1 3.5 3.4 4.2 4.0 4.2

Liquid Yield [%] 89.5 91.3 90.3 86.6 86.3 87.6

Liquid HC Yield [%] 82.9 83.1 83.3 81.3 81.2 80.7

CS2 Insolubles Yield [%] 7.4 5.2 6.3 9.2 9.7 8.2

95

Table 8.SimDist raw data and calculated results for reaction #5 and #6

RUN DATA FEED Reaction #4 Reaction #5

C1 C2 C3 C1 C3

Pressure [psig] 1400 1400 1400 1400 1400

Reaction Temperature [Β°C] 330 340 335 330 340

WHSV CAT-M3 [h-1] 0.2 0.2 0.2 0.2 0.2

WHSV CAT-+ [h-1] 0.5 0.5 0.5 - -

Catalyst Matrix CAT-M3+ CAT-M3+ CAT-M3+ CAT-M4 CAT-M4

Water Yield [%] 7.3 8.2 7.9 5.7 6.1

Gas Yield [%] 3.2 3.5 3.4 4.1 4.3

Naphta Yield (28 - 190 Β°C) [%] 2.1 3.7 4.6 3.8 4.3 3.4

Jet Fuel Yield (190 - 260 Β°C) [%] 5.2 7.1 7.7 8.6 8.0 7.9

Diesel Yield (260 - 343 Β°C) [%] 10.5 15.8 17.4 21.7 17.5 19.0

VGO Yield (343 - 550 Β°C) [%] 26.5 22.2 23.1 24.7 24.3 24.9

Residue Yield (550Β°C+) [%] 55.8 40.7 35.5 30.8 36.0 34.5

CS2 Insolubles Yield [%] 7.3 5.1 6.2 9.1 8.3

Conversion (343 Β°C+) [%] 0.0 23.5 28.9 32.6 26.7 28.3

SIMULATED DISTILLATION

0 162.5 164.95 164.40 167.00 163.75 165.75

5 225.7 195.60 184.60 193.50 187.65 200.25

10 267.9 237.55 227.90 229.30 227.40 235.10

15 301.5 272.75 260.95 260.20 258.70 267.20

20 333.3 302.40 291.65 286.80 288.65 294.65

25 356.2 316.75 312.90 304.10 308.80 312.15

30 368.1 334.00 322.50 310.10 317.90 319.65

35 387.9 353.35 340.70 325.30 337.10 336.85

40 419.6 377.40 357.55 339.60 354.80 352.85

45 458.1 417.70 386.50 354.70 380.00 374.50

50 503.3 454.30 425.95 382.00 416.25 410.35

55 561.3 501.00 461.95 418.00 446.75 441.35

60 633.4 557.70 509.55 449.30 487.70 478.50

65 0.0 628.40 569.10 494.20 532.55 523.05

70 0.0 701.40 648.15 543.00 583.90 574.25

75 0.0 0.00 0.00 0.00 0.00 0.00

80 0.0 0.00 0.00 0.00 0.00 0.00

85 0.0 0.00 0.00 0.00 0.00 0.00

90 0.0 0.00 0.00 0.00 0.00 0.00

95 0.0 0.00 0.00 0.00 0.00 0.00

100 0.0 0.00 0.00 0.00 0.00 0.00

96

Appendix III. Hydrogen consumption calculation

To calculate the Hydrogen consumption it is important to know how much hydrogen is

entering to the system (through the feed and the hydrogen flow) and how much hydrogen is exiting

the process. In this case, hydrogen is coming out in the form of the gases produced in hydrotreating,

in the HDT-bio-oil and in the water produced. The assumption made for this calculation is that the

density of the feed was assumed to be 1 g/mL.

To calculate the hydrogen utilized to hydrogenate the HDT-bio-oil, a balance in carbon will

be made and the H/C ratio will be used. Then, the first step is to calculate the grams of C and H in

the feed entering the HDT process per hour. The elemental analysis of the feed are used to calculate

the flow of H and C per hour for the feed.

�̇�𝐻 (π‘œπ‘–π‘™) [𝑔 𝐻

β„Ž] = (οΏ½Μ‡οΏ½π‘œπ‘–π‘™ [

𝑔 π‘œπ‘–π‘™

β„Ž] βˆ— π‘₯𝐻 𝑓𝑒𝑒𝑑 [

𝑔𝐻

𝑔 π‘œπ‘–π‘™])

�̇�𝐢 (π‘œπ‘–π‘™) [𝑔 𝐢

β„Ž] = (οΏ½Μ‡οΏ½π‘œπ‘–π‘™ [

𝑔 π‘œπ‘–π‘™

β„Ž] βˆ— π‘₯𝐢 𝑓𝑒𝑒𝑑 [

𝑔𝐢

𝑔 π‘œπ‘–π‘™])

Next, the grams of H and C per hour in the gases exiting the system are calculated.

�̇�𝐻 (π‘”π‘Žπ‘ )𝑖 [𝑔 𝐻

β„Ž] = (οΏ½Μ‡οΏ½π‘”π‘Žπ‘  [

π‘šπΏ π‘”π‘Žπ‘ 

β„Ž] βˆ— π‘₯𝑖 [

π‘šπΏ 𝑖

π‘šπΏ π‘”π‘Žπ‘ ]) βˆ— (πœŒπ‘– [

𝑔 𝑖

π‘šπΏ 𝑖]) βˆ— (

1

𝑀𝑖[π‘šπ‘œπ‘™ 𝑖

𝑔 𝑖]) βˆ— (#π‘šπ‘œπ‘™ [

π‘šπ‘œπ‘™ 𝐻

π‘šπ‘œπ‘™ 𝑖]) βˆ— 𝑀𝐻 [

𝑔 𝐻

1 π‘šπ‘œπ‘™ 𝐻]

�̇�𝐢 (π‘”π‘Žπ‘ )𝑖 [𝑔 𝐢

β„Ž] = (οΏ½Μ‡οΏ½π‘”π‘Žπ‘  [

π‘šπΏ π‘”π‘Žπ‘ 

β„Ž] βˆ— π‘₯𝑖 [

π‘šπΏ 𝑖

π‘šπΏ π‘”π‘Žπ‘ ]) βˆ— (πœŒπ‘– [

𝑔 𝑖

π‘šπΏ 𝑖]) βˆ— (

1

𝑀𝑖[π‘šπ‘œπ‘™ 𝑖

𝑔 𝑖]) βˆ— (#π‘šπ‘œπ‘™ [

π‘šπ‘œπ‘™ 𝐢

π‘šπ‘œπ‘™ 𝑖]) βˆ— 𝑀𝐢 [

𝑔 𝐢

1 π‘šπ‘œπ‘™ 𝐢]

97

�̇�𝐻 (π‘”π‘Žπ‘ )π‘‘π‘œπ‘‘π‘Žπ‘™ [𝑔 𝐻

β„Ž] = βˆ‘ �̇�𝐻 (π‘”π‘Žπ‘ )𝑖

𝑛

𝑖=1

�̇�𝐢 (π‘”π‘Žπ‘ )π‘‘π‘œπ‘‘π‘Žπ‘™ [𝑔 𝐢

β„Ž] = βˆ‘ �̇�𝐢 (π‘”π‘Žπ‘ )𝑖

𝑛

𝑖=1

Once, the carbon and hydrogen from the gases are calculated, the grams of H and C from the

HDT-bio-oil per hour can be obtained. Additionally, the grams of H per hour from the water are

calculated.

�̇�𝐢 (π»π·π‘‡π‘œπ‘–π‘™) [𝑔 𝐢

β„Ž] = (�̇�𝐢 (π‘œπ‘–π‘™) βˆ’ �̇�𝐢 (π‘”π‘Žπ‘ )π‘‘π‘œπ‘‘π‘Žπ‘™ )

(𝐻

𝐢)

π»π·π‘‡π‘œπ‘–π‘™[𝑔 𝐻

𝑔 𝐢] =

π‘₯𝐻2π»π·π‘‡π‘œπ‘–π‘™ [𝑔𝐻

𝑔 π‘œπ‘–π‘™]

π‘₯𝐢 π»π·π‘‡π‘œπ‘–π‘™ [𝑔𝐢

𝑔 π‘œπ‘–π‘™]

�̇�𝐻 (π»π·π‘‡π‘œπ‘–π‘™) [𝑔 𝐻

β„Ž] = (�̇�𝐢 (π»π·π‘‡π‘œπ‘–π‘™) [

𝑔 𝐢

β„Ž] βˆ— (

𝐻

𝐢)

π»π·π‘‡π‘œπ‘–π‘™[𝑔 𝐻

𝑔 𝐢])

οΏ½Μ‡οΏ½ 𝐻(𝐻2𝑂) [𝑔 𝐻

β„Ž] = (�̇�𝐻2𝑂 [

𝑔 𝐻2𝑂

β„Ž] βˆ—

1

𝑀𝐻2𝑂[π‘šπ‘œπ‘™ 𝐻2𝑂

𝑔 𝐻2𝑂] βˆ— #π‘šπ‘œπ‘™ [

π‘šπ‘œπ‘™ 𝐻

π‘šπ‘œπ‘™ 𝐻2𝑂] βˆ— 𝑀𝐻 [

𝑔 𝐻

π‘šπ‘œπ‘™ 𝐻])

Knowing the mass of hydrogen per hour that entered and exited the process, it can now be

calculated the hydrogen consumption.

�̇�𝐻 (𝐸π‘₯𝑖𝑑) [𝑔 𝐻

β„Ž] = �̇�𝐻 (π‘”π‘Žπ‘ )π‘‘π‘œπ‘‘π‘Žπ‘™ + �̇�𝐻 (π»π·π‘‡π‘œπ‘–π‘™) + οΏ½Μ‡οΏ½ 𝐻(𝐻2𝑂)

𝐻 π‘π‘œπ‘›π‘ π‘’π‘šπ‘’π‘‘ [𝑔 𝐻

β„Ž] = �̇�𝐻 (𝐸π‘₯𝑖𝑑) βˆ’ �̇�𝐻 (π‘œπ‘–π‘™)

yield𝐻2 π‘π‘œπ‘›π‘ π‘’π‘šπ‘’π‘‘ [𝑔 𝐻2

𝑔 π‘œπ‘–π‘™] =

𝐻2 π‘π‘œπ‘›π‘ π‘’π‘šπ‘’π‘‘ [𝑔 𝐻

β„Ž] βˆ—

1𝑀𝐻

[π‘šπ‘œπ‘™ 𝐻

𝑔 𝐻] βˆ— #π‘šπ‘œπ‘™ [

π‘šπ‘œπ‘™ 𝐻2π‘šπ‘œπ‘™ 𝐻

] βˆ— 𝑀𝐻2[

𝑔 𝐻2

π‘šπ‘œπ‘™ 𝐻2]

οΏ½Μ‡οΏ½π‘œπ‘–π‘™ [𝑔 π‘œπ‘–π‘™

β„Ž]