Ucalgary 2016 Cabralesnavarro Fredy

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 246

University of Calgary

PRISM: University of Calgary's Digital Repository

Graduate Studies The Vault: Electronic Theses and Dissertations

2016

Reactivity Study and Kinetic Modeling of Deasphalted


Oil Upgrading via Thermal and Catalytic Steam
Cracking

Cabrales Navarro, Fredy

Cabrales Navarro, F. (2016). Reactivity Study and Kinetic Modeling of Deasphalted Oil Upgrading
via Thermal and Catalytic Steam Cracking (Unpublished doctoral thesis). University of Calgary,
Calgary, AB. doi:10.11575/PRISM/27538
http://hdl.handle.net/11023/3487
doctoral 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

Reactivity Study and Kinetic Modeling of Deasphalted Oil Upgrading via Thermal and

Catalytic Steam Cracking

by

Fredy Antonio Cabrales Navarro

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF DOCTOR OF PHILOSOPHY

GRADUATE PROGRAM IN CHEMICAL ENGINEERING

CALGARY, ALBERTA

DECEMBER, 2016

© Fredy Antonio Cabrales Navarro 2016

i
Abstract

Solvent deasphalting (SDA) is a well-known process where asphaltenes are removed from
the oil using a paraffinic solvent to produce a lighter and better quality deasphalted oil (DAO).
SDA has uses in refineries and bitumen upgraders. In the latter use, the DAO still needs further
upgrading to make it a transportable oil that meets pipeline specifications. This doctoral thesis
covers the use of thermal cracking as a DAO upgrading technology, as well as catalytic steam
cracking using innovative Ni/K and Ni/Ce ultradispersed (UD) and Ni/Ce fixed-bed (FB) catalyst
formulations as an alternative path to further improve upon the thermal cracking performance. A
detailed reactivity study and comprehensive kinetic modeling for these DAO processing methods
is conducted and discussed throughout the work.

The experimental data for the reactivity experiments is obtained in a research-scale pilot
plant equipped with an up-flow open tubular reactor, which was procured, designed and
constructed as part of this doctoral project. A complete set of characterization techniques is used
for detailed evaluation of the performance of the processes and the analysis of relevant properties
related to the quality of the liquid and gas products.

The effect of operating conditions on DAO thermal cracking including Liquid Hourly Space
Velocity (LHSV), reaction temperature, total pressure and steam partial pressure was assessed
and was found to be consistent with results in the literature regarding thermal cracking of
residual hydrocarbons. Considerable generation of asphaltenes during thermal cracking was
found to be the main drawback of this process and their production was analyzed from a kinetic
point of view. Catalytic steam cracking was found to have the same relevant kinetic and rate
controlling steps as that of thermal cracking in terms of conversion and the effect of operating
variables, however having a catalytic effect to promote water dissociation to induce
hydrogenation and enhancement of the overall product quality. Operating conditions, as well as
the state of oxidation of the catalyst active phase were found to play a critical role for the
appropriate performance of the catalytic process.

Keywords: Deasphalted Oil, Asphaltenes, Catalytic Steam Cracking, Thermal Cracking,


Upgrading.

ii
Acknowledgements

First of all, I would like to thank Dr. Pedro Pereira-Almao for the opportunity of being
part of the Catalysis for Bitumen Upgrading (CBU) in the Schulich School of Engineering at the
University of Calgary. His supervision and advice during the development of this doctoral work
are greatly appreciated. I am certain I would not have had a better professional formation as a
doctor than what I received under his guidance.

I would like to thank all staff members of the CBU Research Group, especially Dr.
Carlos Scott, Lante Carbognani, Dr. Josefina Scott, Alejandro Coy, Dr. Gerardo Vitale and Dr.
Azfar Hassan for all the insightful technical discussions that guided this research work. In
addition, thanks to all members for their support, friendship and all those moments of joy that
made this such a pleasant experience.

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


the Schulich School of Engineering for offering me such an outstanding academic formation.
Thanks to all professors, administrative and technical staff members for their assistance every
time it was needed. The economic support received from the NSERC/NEXEN/AIEES Industrial
Research Chair in Catalysis for Bitumen Upgrading and the following scholarships is also
greatly appreciated: Eyes High International Doctoral Scholarship (2013), Alberta Innovates –
Technology Futures graduate scholarship (2014-2016) and the Graduate Excellence Scholarship
Award for PhD students (2014, 2015, 2016).

Thanks to my family for always believing in me and for always being there to support me
in this very long path of academic and professional projects towards the pursue of my PhD
degree in Chemical Engineering. Thanks also to my girlfriend Erika Scheele for being the best
company I could ever have had during the last year of my PhD formation and for being always
with me in all the good and the difficult times of this achievement.

Finally, thanks to Canada for making me feel like home and for giving me countless
opportunities.

iii
Dedication

To my parents Fredy and Elsa, my sister Paula, my brother Rafael and


my nephew Juan Esteban for all their unconditional and strengthening
support during the journey of this doctoral work

To Queen E, with all my love!

iv
Table of Contents

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

Acknowledgements ........................................................................................................................ iii

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

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

List of Tables ................................................................................................................................. xi

List of Figures and Illustrations ................................................................................................... xiii

List of Symbols, Abbreviations, Nomenclatures ........................................................................ xvii

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

1.1. Background and Motivation ................................................................................................1

1.2. Scope of the Research ..........................................................................................................2

1.3. Organization of the Thesis ...................................................................................................3

1.4. Gantt Chart ...........................................................................................................................4

1.5. References ............................................................................................................................6

Chapter 2: Detailed Description of Experimental Resources ..............................................7

2.1. Design and Construction of Research-Scale Pilot Plant ......................................................7

2.1.1. Feed Section .......................................................................................................................10

2.1.2. Reaction Section ................................................................................................................12

2.1.3. Hot Separation and Collection ...........................................................................................13

2.1.3.1. Semi-continuous Collection ...........................................................................................14

2.1.3.2. Continuous Collection ...................................................................................................15

v
2.1.4. Cold Separation and Collection .........................................................................................17

2.1.5. Gas Release and Analysis ..................................................................................................17

2.1.6. Control and Data Acquisition ............................................................................................18

2.2. Characterization Techniques ..............................................................................................21

2.2.1. Ultradispersed Catalyst Characterization ...........................................................................21

2.2.1.1. Metal Content.................................................................................................................21

2.2.1.2. Particle Size ...................................................................................................................22

2.2.2. Characterization of Feedstock and Liquid Products ..........................................................23

2.2.2.1. Product Distribution .......................................................................................................23

2.2.2.2. Viscosity ........................................................................................................................24

2.2.2.3. API Gravity ....................................................................................................................25

2.2.2.4. Sulfur and Nitrogen........................................................................................................26

2.2.2.5. Asphaltenes Stability Index (P-value) ...........................................................................26

2.2.2.6. Microcarbon Residue .....................................................................................................30

2.2.2.7. Saturates, Aromatics, Resins and Asphaltenes (SARA) ................................................30

2.2.3. Characterization of Gas Products.......................................................................................32

2.2.3.1. Gas Chromatography .....................................................................................................32

2.2.3.2. Mass Spectrometry.........................................................................................................32

2.3. References ..........................................................................................................................33

Chapter 3: Pilot Plant Commissioning...............................................................................35

3.1. Experimental Plan ..............................................................................................................36

vi
3.1.1. Semi-continuous Product Collection .................................................................................36

3.1.2. Continuous Product Collection ..........................................................................................36

3.1.3. Products Characterization ..................................................................................................37

3.2. Results ................................................................................................................................37

3.2.1. Semi-continuous Product Collection .................................................................................37

3.2.2. Continuous Product Collection ..........................................................................................47

3.3. References ..........................................................................................................................50

Chapter 4: Reactivity and Comprehensive Kinetic Modeling of Deasphalted Vacuum


Residue Thermal Cracking ............................................................................................................51

4.1. Abstract ..............................................................................................................................51

4.2. Introduction ........................................................................................................................52

4.3. Experimental Methods .......................................................................................................54

4.3.1. Experimental Setup ............................................................................................................54

4.3.2. Experimental Plan ..............................................................................................................56

4.3.3. Feedstock and Product Characterization ............................................................................58

4.3.4. Kinetic Modeling ...............................................................................................................59

4.4. Results and Discussion ......................................................................................................65

4.4.1. Effect of Partial and Total Steam Pressure ........................................................................65

4.4.2. Reactivity Analysis ............................................................................................................66

4.4.3. Kinetic Study .....................................................................................................................72

4.5. Conclusions ........................................................................................................................86

vii
4.6. Acknowledgement .............................................................................................................87

4.7. References ..........................................................................................................................87

Chapter 5: Catalytic Steam Cracking of Deasphalted Vacuum Residue using a Ni/K


Ultradispersed Catalyst ..................................................................................................................90

5.1. Abstract ..............................................................................................................................90

5.2. Introduction ........................................................................................................................91

5.3. Experimental Methods .......................................................................................................94

5.3.1. Catalytic Feed Preparation .................................................................................................94

5.3.2. Reactivity Test Unit ...........................................................................................................95

5.3.3. Characterization Techniques ..............................................................................................96

5.3.3.1. Feedstock and Liquid Products ......................................................................................96

5.3.3.2. Gas Samples ...................................................................................................................97

5.3.4. Experimental Plan ..............................................................................................................98

5.3.4.1. Evidence of Water Splitting with Isotopic Water ..........................................................98

5.3.4.2. Reactivity Experiments ..................................................................................................98

5.3.5. Kinetic Modeling ...............................................................................................................99

5.4. Results and Discussion ....................................................................................................100

5.4.1. Catalytic Feedstock Preparation ......................................................................................100

5.4.2. Evaluation of Water Splitting ..........................................................................................101

5.4.3. Reactivity Analysis ..........................................................................................................104

5.4.4. Kinetic Study ...................................................................................................................110

viii
5.5. Conclusions ......................................................................................................................113

5.6. Acknowledgement ...........................................................................................................114

5.7. References ........................................................................................................................114

Chapter 6: Comparative Upgrading of Deasphalted Vacuum Residue via Catalytic Steam


Cracking using Ni/Ce-based Ultradispersed and Fixed-bed Catalyst Formulations ...................116

6.1. Abstract ............................................................................................................................116

6.2. Introduction ......................................................................................................................116

6.3. Experimental Methods .....................................................................................................123

6.3.1. Catalytic Feedstock Preparation ......................................................................................123

6.3.2. Supported Catalyst Preparation........................................................................................125

6.3.3. Reactivity Tests ................................................................................................................125

6.3.4. Characterization Techniques ............................................................................................127

6.3.4.1. UDC- Characterization ................................................................................................127

6.3.4.2. Supported Catalyst Characterization............................................................................127

6.3.4.3. Liquid Products ............................................................................................................128

6.3.4.4. Gas Products ................................................................................................................129

6.3.5. Experimental Procedures for Catalyst Evaluations..........................................................129

6.3.5.1. Ni/Ce Ultradispersed Catalysts ....................................................................................129

6.3.5.2. Ni/Ce Fixed-bed Catalysts ...........................................................................................130

6.4. Results ..............................................................................................................................131

6.4.1. Ultradispersed Catalyst Preparation .................................................................................131

ix
6.4.2. Supported Catalyst Characterization................................................................................131

6.4.3. Catalytic Steam Cracking with Ultradispersed Catalyst ..................................................137

6.4.4. Catalytic Steam Cracking in a Fixed-bed Reactor ...........................................................141

6.4.5. Comparison of Thermal vs. UDC vs. FB Processing ......................................................145

6.5. Conclusions ......................................................................................................................147

6.6. Acknowledgement ...........................................................................................................148

6.7. References ........................................................................................................................148

Chapter 7: Conclusions and Recommendations ..............................................................151

7.1. Conclusions ......................................................................................................................151

7.2. Recommendations ............................................................................................................154

Appendixes ..................................................................................................................................156

Appendix A – Standard Operating Procedure (SOP) – RTU-1 ...................................................156

Appendix B – Chapter 3 Operational Data and Experimental Results ........................................190

Appendix C – Chapter 4 Operational Data and Experimental Results ........................................196

Appendix D – Chapter 5 Operational Data and Experimental Results ........................................212

Appendix E – Chapter 6 Operational Data and Experimental Results ........................................219

Appendix F – Copyrights Authorization Letter ...........................................................................225

x
List of Tables

Table 2.1. Comparison of characteristics of previous design vs. new design of the RTU-1 .......... 9

Table 2.2. Control parameters ....................................................................................................... 19

Table 2.3. Microscopic images for P-value determination ........................................................... 28

Table 2.4. Microscope images for P-value analysis of samples with high presence of solids ..... 29

Table 3.1. South American DAO feedstock characterization ....................................................... 35

Table 3.2. Statistical analysis of temperature variation for thermal and catalytic experiments ... 42

Table 3.3. Statistical analysis of pressure variation for thermal and catalytic experiments ......... 42

Table 3.4. Summary of experimental information for semi-continuous commissioning runs.


Section A: thermal cracking; Section B: catalytic steam cracking ............................................... 44

Table 4.1. Feedstock characterization ........................................................................................... 56

Table 4.2. Effect of pressure on converted DAO thermal cracking. Results reported for the whole
liquid product recovered after reaction ......................................................................................... 66

Table 4.3. Summary of experimental data for thermal cracking runs .......................................... 68

Table 4.4. Statistical analysis of experimental variation for each kinetic model lump ................ 75

Table 4.5. Estimated kinetic parameters for case 1, case 2 and case 3 for converted DAO ......... 82

Table 4.6. Estimated kinetic parameters for case 1, case 2 and case 3 for virgin DAO ............... 83

Table 4.7. Comparison of estimated kinetic parameters with and without using re-
parameterization of the Arrhenius equation for case 1 – converted DAO .................................... 84

Table 4.8. Kinetic parameters for case 3 - converted DAO using a 6-lumps kinetic model ........ 84

Table 5.1. Catalyst concentration in DAO+CAT ....................................................................... 100

Table 5.2. DAO feedstock properties before and after catalyst incorporation ........................... 101

Table 5.3. Summary of experimental conditions and whole liquid product characterization for
catalytic steam cracking of DAO ................................................................................................ 107

Table 5.4. Estimated kinetic parameters for upgrading of DAO via catalytic steam cracking ... 113

xi
Table 6.1. Catalyst concentration and particle diameter in DAO+CAT ..................................... 131

Table 6.2. Textural properties of supported catalyst................................................................... 132

Table 6.3. Metals content for supported (fresh) catalyst ............................................................ 132

Table 6.4. Summary of experimental conditions and whole liquid product characterization for
thermal and catalytic steam cracking of DAO using Ni/Ce UDC .............................................. 141

Table 6.5. Summary of experimental conditions and whole liquid product characterization for
thermal and catalytic steam cracking of DAO using Ni/Ce-based fixed-bed catalyst. Section A:
reactions at 360 °C. Section B: reactions at 380 °C.................................................................... 145

Table 6.6. Whole liquid product characterization comparison for thermal and catalytic runs at
isoconversion .............................................................................................................................. 146

xii
List of Figures and Illustrations

Figure 1.1 Gantt chart. .................................................................................................................... 5

Figure 2.1. RTU-1 diagram. .......................................................................................................... 10

Figure 2.2. RTU-1 pictures. Left: front vertical plane of the unit; right: reactor.......................... 10

Figure 2.3. RTU-1 feed section..................................................................................................... 12

Figure 2.4. RTU-1 reactor & thermocouple profile probe schematic. .......................................... 13

Figure 2.5. RTU-1 hot separation section. .................................................................................... 14

Figure 2.6. RTU-1 continuous collection configuration in hot separator. .................................... 16

Figure 2.7. RTU-1 cold separation section. .................................................................................. 17

Figure 2.8. RTU-1 gas release and analysis section. .................................................................... 18

Figure 2.9. Hardware for data acquisition and control. ................................................................ 21

Figure 2.10. Asphaltene agglomerates size in precipitated sample in P-value analysis. .............. 29

Figure 3.1. Reactor pressure and temperature profiles for thermal cracking experiments
(condition 1 to 4)........................................................................................................................... 38

Figure 3.2. Reactor pressure and temperature profiles for catalytic steam cracking experiments
(condition 5 to 9)........................................................................................................................... 39

Figure 3.3. Detailed reactor temperature profile for catalytic steam cracking experiment
(condition 9). ................................................................................................................................. 40

Figure 3.4. Reactor temperature profile for catalytic steam cracking experiments (condition 10
and 11). ......................................................................................................................................... 41

Figure 3.5. Product properties as function of conversion for commissioning runs using the semi-
continuous product collection system. .......................................................................................... 45

Figure 3.6. Product distribution and microcarbon content as function of conversion for
commissioning runs using the semi-continuous product collection system. ................................ 45

Figure 3.7. Microscope image of sample for condition 9 (435 °C-2 h-1) (0.15 mL nC16 added)
during the P-value analysis. .......................................................................................................... 46

xiii
Figure 3.8. Microscope image of sample for condition 11 (440 °C-2 h-1) (0.15 mL nC16 added)
during the P-value analysis. .......................................................................................................... 47

Figure 3.9. Temperature and pressure profile for continuous product collecting system at 423 °C
and 2 h-1......................................................................................................................................... 48

Figure 3.10. Temperature and pressure profile for continuous product collecting system at 380
°C and 0.25 h-1. ............................................................................................................................. 49

Figure 3.11. Simulated distillation comparison for product fractions gathered at 423 °C and 2 h -1
for continuous product collecting system. .................................................................................... 49

Figure 3.12. Simulated distillation comparison for product fractions gathered at 380 °C and 0.25
h-1 for continuous product collecting system. ............................................................................... 50

Figure 4.1. OrCrude upgrading scheme. ................................................................................... 53

Figure 4.2. Schematic of the experimental setup for reactivity tests. ........................................... 55

Figure 4.3. Proposed kinetic model for DAO thermal cracking. .................................................. 60

Figure 4.4. 6-lumps kinetic model for DAO thermal cracking. .................................................... 65

Figure 4.5. Product distributions and P-value vs. residue (560 °C+) conversion for converted and
virgin DAO thermal cracking. ...................................................................................................... 70

Figure 4.6. Viscosity and API gravity vs residue (560 °C+) conversion. ..................................... 71

Figure 4.7. MCR residue and P-value profiles vs. residue (560 °C+) conversion........................ 72

Figure 4.8. Predicted model composition vs. experimental composition for case 1(unconstrained)
converted DAO kinetic model. ..................................................................................................... 73

Figure 4.9. Predicted model composition vs. experimental composition for case 1(unconstrained)
virgin DAO kinetic model. ........................................................................................................... 74

Figure 4.10. Experimental vs. model compositions as a function of space time at 423 °C for case
1 (unconstrained) converted DAO kinetic model. ........................................................................ 74

Figure 4.11. Comparison between the modeling error by lump for converted and virgin DAO
kinetic modeling............................................................................................................................ 75

xiv
Figure 4.12. Average lump composition with error bars .............................................................. 76

Figure 4.13. Relative initial velocities of formation of each lump at 423 °C for DAO (560 °C+) as
percentage of the global initial velocity. ....................................................................................... 81

Figure 4.14. Predicted model compositions vs. experimental compositions for case 3 - converted
DAO for a 6-lumps kinetic model. ............................................................................................... 85

Figure 5.1. Catalyst incorporation unit – skid............................................................................... 95

Figure 5.2. Schematic of the experimental setup for reactivity tests. ........................................... 96

Figure 5.3. Reaction pathways for kinetic modeling of DAO catalytic steam cracking. ............. 99

Figure 5.4. Mechanism of H2O splitting by K+ .......................................................................... 102

Figure 5.5. Mass spectrometry analysis for catalytic steam cracking of DAO using O18 isotopic
water. ........................................................................................................................................... 103

Figure 5.6. H2 and CO2 comparison for thermal tracking and catalytic steam cracking. Ci:
catalytic conditions. .................................................................................................................... 104

Figure 5.7. Product distribution for thermal cracking and catalytic steam cracking. ................. 105

Figure 5.8. Microscope images of P-value for products gathered at 440 ºC via catalytic steam
cracking. ...................................................................................................................................... 107

Figure 5.9. Comparison of solid contents on the whole liquid products from catalytic steam
cracking and thermal cracking. ................................................................................................... 109

Figure 5.10. MCR and viscosity profile vs. conversion of HC (560 °C+).................................. 110

Figure 5.11. Predicted model composition vs. experimental composition for kinetic modeling of
catalytic steam cracking of DAO. ............................................................................................... 111

Figure 5.12. Average absolute error by lump for kinetic modeling of catalytic steam cracking of
DAO. ........................................................................................................................................... 111

Figure 6.1. Transformation of molecular groups for thermal cracking and catalytic steam
cracking of vacuum residue and vacuum gas oil.20.................................................................... 121

xv
Figure 6.2. Effects of operating variables and condensation reactions when choosing a catalytic
steam cracking catalyst. .............................................................................................................. 123

Figure 6.3. Ni/Ce ultradispersed catalyst preparation unit – skid. .............................................. 125

Figure 6.4. Schematic of packed-bed and empty tube reactors. Reactor drawing from Swagelok
database. ...................................................................................................................................... 126

Figure 6.5. Textural properties for the supported catalyst: a) adsorption-desorption isotherms and
b) pore width distribution............................................................................................................ 134

Figure 6.6. XRD diffractograms for the supported catalyst (fresh, spent and regenerated).
Vertical red lines indicate the CeO2 phase (PDF#01-073-6318); vertical blue lines indicate the
NiO phase (PDF#01-075-0269); Vertical green lines indicate the TiO2 phase (PDF#01-089-
4921). .......................................................................................................................................... 135

Figure 6.7. XRD diffractograms comparison for fresh, spent and regenerated catalyst. ............ 136

Figure 6.8. Comparison of H2 and CO2 GC results for thermal and catalytic steam cracking using
Ni/Ce UDC. AR: Atomic Ratio (Molar). .................................................................................... 138

Figure 6.9 Comparison of H2 and CO2 QMS results for thermal and catalytic steam cracking
using Ni/Ce UDC. ....................................................................................................................... 138

Figure 6.10. Conversion and liquid product distribution for thermal and catalytic steam cracking
of DAO using Ni/Ce UDC. ......................................................................................................... 139

Figure 6.11. Comparison of H2 and CO2 QMS results for thermal and catalytic steam cracking
using Ni/Ce-based fixed-bed catalyst. ........................................................................................ 142

Figure 6.12. GC analysis of H2 and CO2 and H2/CO2 ratios for fixed-bed reactions. ................ 143

Figure 6.13. Liquid product distributions at isoconversion. ....................................................... 147

xvi
List of Symbols, Abbreviations, Nomenclatures

AAE Average Absolute Error

AE Absolute Error

AGO Atmospheric Gas Oil

AIEES Alberta Innovates Energy and Environment Solutions

API American Petroleum Institute

AR Atomic Ratio

Asp-C5 Asphaltenes (as pentane insoluble)

ASTM American Society for Testing and Materials

BPV Back Pressure Valve

CAT Catalyst

CBU Catalyst for Bitumen Upgrading

CSC Catalytic Steam Cracking

DAO Deasphalted Oil

Dilbit Diluted Bitumen

DLS Dynamic Light Scattering

Ea Activation Energy

FID Flame Ionization Detector

FPD Flame Photometric Detector

GAAE Global Average Absolute Error

GC Gas Chromatography

xvii
HC Hydrocarbon

HDT Hydrotalcite

HVGO Heavy Vacuum Gas Oil

ICP Inductively Coupled Plasma

ki Kinetic Constant for reaction i

LHSV Liquid Hourly Space Velocity

LVGO Light Vacuum Gas Oil

MB-0 Mass Balance-Number

MCR Microcarbon Residue

MS Mass Spectrometry

mx Mass of x

nC5 Pentane

np Number of Experiments at each Temperature

NSERC Natural Sciences and Engineering Research Council

nt Number of Evaluated Temperatures

NTA Nanoparticle Tracking Analysis

O.D. Outside Diameter

ODE Ordinary Differential Equations

PI-000 Pressure Indicator-Number

PID Proportional, Integral, Derivative

PLC Programmable Logic Controller

xviii
PTV Programmable Temperature Vaporizing

Pv P-value

Pw0 Partial Pressure Number

QMS Quadrupole Mass Spectrometer

RE Relative Error

rk Coefficient of Determination for Arrhenius Fitting of Reaction k

RTU-1 Reactivity Test Unit 1

SARA Saturates, Aromatics, Resins and Asphaltenes

SC Steam Cracking

SCO Synthetic Crude Oil

SDA Solvent Deasphalting

SimDist Simulated Distillation

S.G Specific Gravity

SOP Standard Operation Procedure

SSE Sum of Squared Errors

T Temperature

TAN Total Acid Number

TC Thermal Cracking

TC-000 Temperature Controller-Number

TCD Thermal Conductivity Detector

TI Internal Fluid Temperature Indicator

xix
TLC Thin Layer Chromatography

TLC Thin Liquid Chromatography

UDC Ultradispersed Catalyst

V-000 Valve-Number

VGO Vacuum Gasoil

VR Vacuum Residue

WGS Water Gas Shift

wi Weight Fraction of Lump i

WHSV Weight Hourly Space Velocity

W.T. Wall Thickness

WTM Wet Gas Meter

yi Yield of lump i

x Density of x

 Space Time

i Reaction Rate for Reaction i

xx
Chapter 1: Introduction

1.1. Background and Motivation

In the last decade, a sharp worldwide depletion of conventional hydrocarbon sources has
forced the oil and energy sector to move toward unconventional alternatives to fulfill the
increasing demand for petroleum-based fuels currently used as primary energy source.
Therefore, the exploitation of heavy, extra-heavy hydrocarbons and bitumen has become very
common in most oil producer countries and appropriate technologies to handle these heavy oils
are being developed for their production as well as their upgrading and refining.1,2 Due to their
heavy nature and high viscosity, most of these resources are not transportable and need
upgrading process or dilution in order to meet pipeline specifications: The initial approach to
give added-value to Canadian bitumen was via full upgrading, where a synthetic crude oil (SCO)
with very good properties (i.e. API 30-40) was produced. However, the high capital cost required
to build these multi-billion dollar upgrading facilities combined with reduced commodities prices
hindered the economic viability of these projects. A different approach to make the heavy
bitumen transportable was to produce a diluted bitumen (Dilbit) using a light hydrocarbon as
diluent. Nonetheless, the high cost of diluent and the considerable amount required to reach
transportability conditions does not make it the most economically attractive option. In addition,
quality of the oil is not improved when using diluent.3,4 Partial upgrading, where bitumen is
upgraded to reach Canadian transportability conditions (API: 20; Viscosity 350 cSt at carrier
reference temperature; Olefins < 1% wt 1-decene equivalent), has the potential to maximize the
added-value to the bitumen with a minimized capital cost that makes it economically attractive
using simplified upgrading schemes at a smaller scale as compared to full-upgrading and, in
addition, no diluent is needed.3,4 In fact, the federal government recently launched the National
Partial Upgrading Program to support initiatives to develop technologies in this area, supporting
field pilot level tests, aiming to increase the competitiveness of the Canadian industry in the
global markets and to fill technical gaps present during the development stage of these
technologies. In this way, this area of investigation is a matter of national interest and the
development of new more cost efficient upgrading methods has become the center of attention
for many research institutions and universities related to the energy sector, especially in countries
like Canada where there are enormous reserves of heavy and extra-heavy oils.5

1
In this order of ideas, this doctoral work, developed under the NSERC/NEXEN/AIEES
Industrial Research Chair in Catalysis for Bitumen Upgrading directed by Dr. Pedro Pereira-
Almao in the Catalysis for Bitumen Upgrading (CBU) research group at the University of
Calgary, focuses on the study of an alternative upgrading process applicable to particular heavy
hydrocarbon fractions using innovative nano-catalytic and steam activation processes as well as
the use of conventional thermal cracking technologies as a benchmark.

In the catalytic steam cracking process, an in-house formulated ultra-dispersed catalyst is


incorporated in the feedstock through a decomposition reaction performed in a catalyst
manufacturing skid designed for this purpose.6 Then, the catalyst-loaded hydrocarbon is fed to an
up-flow reactor mixed with a steam stream where it is heated to catalytic cracking reaction
conditions. In a different approach, a supported catalyst with the similar functionality is used in a
fixed-bed reactor to perform the catalytic steam cracking reaction. In both cases, dissociation of
the steam promoted by the catalytically active sites induces hydrogenation, preventing olefin
formation and initial condensation of remaining asphaltene molecules, allowing the process to
reach higher temperatures than conventional thermal cracking processes, without compromising
the stability of the products and, at the same time, inhibiting the production of considerable
amounts of coke as byproduct.7,8 As a result, higher amounts of light hydrocarbon fractions are
expected to be produced, increasing the economic viability of the process.

This research proposal may bring relevant contributions in the bitumen upgrading area,
ranging from the enhancement of the current process up to the development of completely new
alternatives that can have significant economic, environmental, and also technical benefits to the
oil and gas industry.

1.2. Scope of the Research

The main objective of this work is to gain understanding on the upgrading mechanism of
non-asphaltene containing heavy hydrocarbons when processed via thermal cracking and
catalytic steam cracking. The specific objectives to be fulfilled in this research are:

 Conduct a systematic study of variables effect, such as temperature, space velocity, total
pressure and steam partial pressure, on the reactivity of partially cracked and virgin DAO
via thermal cracking.

2
 Investigate the use of catalytic steam cracking of DAO with a conventional Ni/K
ultradispersed catalyst (UDC) formulations including systematic variables study of the
effect of temperature and space velocity.
 Validate the mechanism of water splitting taking place during catalytic steam cracking
reactions.
 Develop comprehensive kinetic models involving the formation of asphaltenes for both
processes.
 Investigate a new Ni/Ce catalyst formulation aiming to outperform the conventional in-
house formulations (Ni/K-UDC).
 Compare the performance of ultradispersed vs. fixed-bed for Ni/Ce-based catalyst
formulations.
 Analyze the effect of the oxidation state of the catalyst active phase when conducting
catalytic steam cracking reactions.

1.3. Organization of the Thesis

This thesis was prepared using the manuscript-based model of the Faculty of Graduate
Studies at the University of Calgary, where research is presented as a collection of scientific
article-type manuscripts. In this work, 5 different manuscripts are included, three of them in the
process of publication in relevant scientific journals of the corresponding research area. The
following works were also presented in widely renowned international conferences:

 Fredy A. Cabrales-Navarro and Pedro Pereira-Almao, “Kinetic Modeling of Deasphalted


Oil Upgrading via Thermal Cracking and Catalytic Steam Cracking using K-Ni
Ultradispersed Catalyst”. Presented at: World Heavy Oil Congress 2015, Edmonton, AB,
Canada.
 Fredy A. Cabrales-Navarro and Pedro Pereira-Almao, “Reactivity Study of Deasphalted
Oil Upgrading via Catalytic Steam Cracking using Ultradispersed Catalysts,”. Presented
at: 2014 AIChE Annual Meeting, Atlanta, GA, USA.

The introduction and motivation of this doctoral work is presented in Chapter 1.

Chapter 2 aims to give a detailed description of all experimental resources used during the
development of this project. This includes the procurement, design and construction of the

3
research-scale pilot plant where the core of the experimental evaluations was conducted. Also,
detailed information and description of the analytical characterization methods is provided.

Chapter 3 includes the commissioning of the constructed pilot plant with a preliminary
baseline study that allowed to evaluate the reliability, operability, safety factors, among other
factors of the experimental setup before the development of the core experimental work. These
two first chapters represent the first stage of development of this project, where a considerable
amount of time commitment, which accounts for approximately 30% of the total time of the
project, was allocated.

Chapter 4 presents a thorough study of vacuum residue DAO thermal cracking including
reactivity of two different feedstocks and new comprehensive kinetic modeling including
asphaltenes generation. This chapter is to be submitted for publication as a scientific article.

Chapter 5 includes a reactivity study and kinetic modeling of catalytic steam cracking of
DAO using a conventional Ni/K formulation that had not been previously used for upgrading of
this particular type of feedstock. Results in Chapter 5 are compared with the baseline thermal
cracking results presented in Chapter 4 as baseline. This chapter is to be submitted for
publication as a scientific article.

Results in Chapter 5 directed this research towards the exploration of catalyst formulations
with lower activation temperatures. In this way, Ni/Ce-based UDC formulations as well as a
fixed-bed formulation were evaluated. These results are presented in Chapter 6, which is also to
be submitted for publication as a scientific article.

Chapter 7 presents a summary of the conclusions of the work, as well as recommendations


for possible continuation of research in this area.

Finally, appendices A to E present the following additional information: standard operation


procedure of the pilot plant (SOP), raw pilot plant data for each experiment conducted, detailed
characterization information and data post-processing with mass balance, conversion, product
distribution, reaction yields, etc.

1.4. Gantt Chart

The Gantt chart presented in Figure 1.1 provides a reference for the timelines followed
during the planning and execution of this investigation.

4
Figure 1.1 Gantt chart.

5
1.5. References

(1) Pereira-Almao, P.; Flores, C.; Zbinden, H.; Guitian, J.; Solari, R. B.; Feintuch, H.; Gillis, D.
Aquaconversion technology offers added value to E. Venezuela synthetic crude oil production.
Oil Gas J. 2001, 99 (20), 79-85.
(2) Pereira-Almao, P.; Marzin, R.; Zacarias, L.; Cordova, J.; Carrazza, J.; Marino, M. Steam
conversion process and catalyst. U.S. Patent No. US5885441, 1999.
(3) De Klerk, A.; Zerpa-Reques, N. G.; Xia, Y.; Omer, A. A. Integrated central processing
facility (cpf) in oil field upgrading (ofu). U.S. Patent No. US20140138287A1, 2014.
(4) Zerpa-Reques, N. G.; De Clerk, A.; Xia, Y.; Omer, A. A. Upgrading of hydrocarbon material.
U.S. Patent No. US20130043033, 2016.
(5) Alberta's Energy Reserves 2014 and Supply/Demand Outlook 2015-2024. Alberta Energy
Regulator: Calgary, 2015.
(6) Vasquez, A. Synthesis, Characterization and Model Reactivity of Ultra Dispersed Catalysts
for Hydroprocessing. MSc Thesis, University of Calgary, Calgary, 2007.
(7) Trujillo-Ferrer, G. Thermal and Catalytic Steam Reactivity Evaluation of Athabasca Vacuum
Gasoil. MSc Thesis, University of Calgary, Calgary, 2008.
(8) Fathi, M. M. Comparative Upgrading of Arab Light Vacuum Residuum via Aquaprocessing
and Thermal Cracking. PhD Thesis, University of Calgary, Calgary, 2011.

6
Chapter 2: Detailed Description of Experimental Resources

2.1. Design and Construction of Research-Scale Pilot Plant

The reactivity tests for the different feedstocks were carried out in a Reactivity Test Unit
(RTU-1) pilot plant equipped with an up-flow open tubular reactor that emulates the
performance of an industrial thermal cracking or visbreaking unit. As indicated by Trujillo-
Ferrer1, there are different definitions for a pilot plant. These definitions are based on equipment
size, purpose, operating group, etc.
Based on size, which is a very common definition, pilot plants can be classified as follows1:
 Laboratory-scale, bench-top test plants or micro-units: these are small equipment that fit
on a bench-top or inside a small laboratory hood. They range from 0.5 to 1.0 m2 and use
1/16” to 1/4” tubing, and are usually totally manual and continuously attended.
 Integrated pilot plants or research-scale pilot plants: these pilot plants are the most
common for Research & Development (R&D) of chemical processes in industrial
organizations. These equipment range in size from 2 to 14 m2 and use 1/4” to 1” tubing
for construction. They are usually automated and commonly designed for unattended
operations.
 Demonstration units, semi-works units or prototype units: these units, built with
commercial-size pipes ranging from 1” to 8”, resemble the commercial scale plants in
terms of automation and operation. In terms of size, they are in the order of 900 m2.

For this research, and based on the definitions used in the Catalysis for Bitumen Upgrading
group, a pilot plant is defined as a specialized equipment that allows representation and
evaluation of all process variables relevant to the industrial equipment and of paramount
importance for process scale-up (e.g. heat transfer, fluid dynamics, reactivity) regardless of the
size of the unit. Based on its size, the RTU-1 is classified as a research-scale pilot plant.
A first version of the unit, which was operated for several years in the CBU research group
at the University of Calgary, was built by Trujillo-Ferrer1 in 2008. Then, the unit was redesigned,
rebuilt, and validated as part of this doctoral project during 2013. The former experimental setup
was completely dismantled and the pieces and equipment in good condition were reused. The
main focus of the new design was to build a pilot plant with a vertical type of layout, capable of
handling high pressures and a wider range of operability conditions with a simpler operation of

7
the unit. The new pilot plant can be operated by one person since most of the manipulation is in
one vertical plane, in contrast to the previous top-bench type layout. Due to optimized space
requirements of the new unit, two pilot plants can be accommodated in the same enclosure. Even
though the previous setup was flexible in temperature and Liquid Hourly Space Velocity
(LHSV) operating conditions, the operating pressure was limited to 450 psig. For the new
design, the unit was revamped up to a rating for safe operation of 1800 psig. Even though this
high pressure is not necessary for either thermal cracking or catalytic steam cracking reactions
investigated in this work, this versatility to operate at higher pressures was included in the design
so that the unit could be used to carry out hydrocracking or bitumen in-situ upgrading
evaluations in other research projects conducted at the CBU group. It is important to highlight
that due to the pressure-temperature rating of some pieces, replacement of some parts might be
needed depending on the operating temperature of each section of the unit. A cold separation
section and a depressurization trap were included in order to improve the collection of light
products and water, which was very problematic with the previous design. A major improvement
was also made in the liquid products collection system with the incorporation of a continuous
automated collection device consisting of a configuration of computer-controlled pneumatic
valves with an intermediate collector vessel. Table 2.1 presents a summary of the major changes
in design and operability of the RTU-1 for the new design compared to the previous unit.
The new RTU-1 is composed of five sections: the feed section, reaction section, hot
separation section, cold separation section, and finally the gas release and analysis section. A
general diagram of the new RTU-1 is shown in Figure 2.1, and pictures of the experimental setup
built are shown on Figure 2.2. Operation details are given in the Standard Operating Procedure
(SOP) of the pilot plant created as part of this doctoral work, found in Appendix A. Details about
each specific section of the unit are given in the upcoming subsections.

8
Table 2.1. Comparison of characteristics of previous design vs. new design of the RTU-1

Previous RTU-1 New RTU-1

Top-bench Vertical Configuration

Layout

Feed Temperature [°C] 20-140 20-140


Feed Flow [cc/min] 0.01-15 0.01-15
Operating Pressure [psig] 0-450 0-1800
Reaction Temp. [°C] Up to 500 Up to 500
LHSV [h-1] 0.1-4.0 0.1-6.0
Hot Separator Yes Yes
Cold Separator No Yes
Depressurization trap No Yes
Products collection Semi-continuous Semi-continuous/Continuous
LabVIEW control software Yes Yes

9
Figure 2.1. RTU-1 diagram.

Figure 2.2. RTU-1 pictures. Left: front vertical plane of the unit; right: reactor.
2.1.1. Feed Section

The feed section, whose schematic is shown in Figure 2.3, is equipped with a Teledyne
ISCO series 500D dual-pump continuous flow system with dual pneumatic valves controlled by

10
a Series D Controller, which refills one pump while the other one is delivering fluid. Two feed
tanks are available to supply feedstock to the pumps. The main feed tank is a custom made 10 L
stainless steel vessel of 6.7” O.D. and 15” height manufactured at the Engineering Machine Shop
at the University of Calgary, where enough feedstock can be stored for several days of
continuous work. The main feed tank is equipped with a spring-type relief valve that opens at a
pressure of 100 psig in case an over pressurization of the vessel due to malfunctioning of the
heating device occurs. The second feed tank is an auxiliary 1 L Swagelok vessel operated at
room temperature where lighter hydrocarbons, such as gasoline, toluene or vacuum gasoil
(VGO) are stored for cooling down and cleaning the unit after each reaction. Under normal
operation, a blanket of 70 psig of N2 is used for pressurization of both tanks with the objective of
creating an oxygen-free environment to preserve the feedstock from oxidation and also to
provide head pressure to uniformly refill the ISCO pump cylinders without any bubbles. A
second spring-type relief valve that opens at a pressure of 1500 psig directs the feed flow to an
auxiliary 500 ml Swagelok vessel placed on the pump outlet line in case there is reactor or lines
plugging downstream. This setting can be easily changed depending on the operating conditions.
Additionally, the feed section is equipped with a Shimadzu LC-20AD high performance liquid
chromatograph pump for water injection. A heater is placed in the water outlet line (TC-201) to
generate superheated steam before the mixing point with the oil. A Brooks Instrument 5850EM
hydrogen mass flow controller is installed for hydrogen injection if needed. The feedstock
pumping section comprised of the ISCO pump cylinders and valves, main feed tank, relief valve,
and interconnecting lines made of ¼” O.D. 316 stainless steel tubing from Swagelok, are
equipped with heating tapes for operation at temperatures up to 140 °C (TC-101 to TC-108).
Every heated piece is insulated with Superwool insulation (Ref. 6# SW 607, supplied by
Inproheat-Edmonton) to diminish heat losses. The limiting factor for a further increase in
temperature of the feed section determined by the pneumatic valves of the ISCO pumps, whose
rating goes up to 140 °C. Operation above this temperature would damage the seals of these
valves. The feed section temperature is set depending on the feedstock properties in such a way
that its viscosity is below 400 cP in order to ensure appropriate performance of the
pumping/refilling cycles of the ISCO pumps without any flow disruption. However, most heavy
oils and fractions, such as: vacuum residue (VR), DAO, VGO, can be pumped at 140 °C.

11
Figure 2.3. RTU-1 feed section.

2.1.2. Reaction Section

The RTU-1 is equipped with a 103.46 mL open tubular reactor operated in up-flow mode.
An Omega custom thermocouple provided with 7 sensing points is assembled inside the reactor
for temperature monitoring as shown in Figure 2.4. The thermocouple probe was designed and
custom-made for this specific reactor in such a way that the 5 sensing points are distributed
inside the reaction zone whereas the outermost two points indicate the temperatures before the
inlet and after the outlet of the reactor. The reactor was fabricated with 50 cm length 316
stainless steel Swagelok tubing 3/4” O.D. and 0.049” wall thickness. The aspect ratio of the
reactor is higher than 10 to diminish radial gradients and back-mixing in such a way that a plug-
flow regime can be assumed. Reactor dimensions were evaluated by Trujillo-Ferrer1 and this
configuration was found to be the most appropriate for this application. A fixed-bed can be
placed in the reactor in case a supported catalyst has to be tested, or it can be operated as an
empty tube. There are three individually controlled heating sections in the reactor (TC-203, TC-
204 & TC-205) in such a way that there is versatility to adjust any of the sections output
independently to reach the desired temperature profile. A 3/8” O.D. and 10 cm length 316

12
stainless steel tube was used as pre-heater before the reactor entrance in order to reach a
temperature close to reaction conditions ensuring no reaction is occurring at this place, in this
way reducing the heat load at the reactor entrance. The products coming out from the reaction
zone go to a hot separator.

Figure 2.4. RTU-1 reactor & thermocouple profile probe schematic.

2.1.3. Hot Separation and Collection

The products coming out of the reactor, which consist of a mixture of reaction gases, steam
and liquid hydrocarbons go to a hot separator operated at a temperature which ensures that water
condensation does not occur. If the temperature is too low, such that the oil/steam mixture is
cooled below the water boiling point at the operating pressure, a water-in-oil emulsion can form
in the collected liquid product, making product characterization and closure of the mass balances
difficult. For the 300 psig operating pressure normally used for catalytic steam cracking and
thermal cracking, the separator was kept at a constant temperature of 250 °C at its wall (internal
temperature ≈ 240 °C). This temperature guarantees water removal from the liquid products. The

13
gaseous separated stream that consists of gaseous hydrocarbons, steam and light hydrocarbons
that are carried away, is processed in the cold separation section. The hot separation section for
the RTU-1 shown in Figure 2.5 consists of the following double ended 304L stainless steel tanks
supplied by Swagelok: one 1 gallon stability tank, two 1 L mass balance vessels and one 40 mL
intermediate vessel. Depending on the products collection mode (semi-continuous or
continuous), different tanks can be used to collect the mass balance as explained in more detail
below.

Figure 2.5. RTU-1 hot separation section.

2.1.3.1. Semi-continuous Collection

In this mode of operation, both stability tank (ST) 1 and mass balance (MB) tank 1 act as hot
separators. Under normal operation of the unit, the stream coming out of the reactor is aligned to
either one of the vessels while the other one is being depressurized and drained to collect the
liquid product. Then, it is re-pressurized to the operating pressure. A
pressurization/depressurization panel to facilitate this task was built in the unit. Normally, the
stability tank receives the product from the start-up and stabilization period due to its large size
that makes it capable of collecting product for several hours of running. Once the unit is stable
for starting a mass balance, the stream is directed to the mass balance tank 1 for a given time.
This redirection of the stream coming out of the reactor is made with a pair of manually actuated

14
3-way valves. After the mass balance time is finished, the stream is aligned back to the stability
tank and sample is collected from the mass balance tank after depressurization. This procedure
can be repeated as much as needed to collect the required amount of liquid product. The
drawbacks of this semi-continuous approach is that the operator has to be present any time an
alignment has to be done, and most importantly, the concentration of the gas phase is altered
every time the tanks are switched since they are re-pressurized with industrial nitrogen that
dilutes the reaction gases in the unit. In this way, continuous gas concentration monitoring (i.e.
mass spectrometry) is not suitable since nitrogen entering the system modifies the measurements
in the equipment. An accurate and reliable online gas analysis is critical for identifying catalytic
activity as explained in Chapters 5 and 6. Therefore, incorporation of a continuous collecting
system that would allow the user to perform the required analysis and also to facilitate the
operation of the pilot plant was considered in a second stage of design of the unit.

2.1.3.2. Continuous Collection

The main issue for continuous collection of product is that the separator is operated at a
temperature of 250 °C and the reaction pressure is normally 300 psig, therefore opening of the
draining valve would not be safe since the hot liquid would splash as soon as the valve is opened,
depressurizing the unit. To address this issue, the liquid product has to be collected in a vessel
operated preferably at atmospheric pressure and a lower temperature than the one needed for
water separation, but higher than the minimum temperature that guarantees flow of the liquid
product (140 °C set point in this project). The main challenge comes with the device that allows
transition of product flow from the high pressure tank to the atmospheric collector. Controlling
the pressure in the outlet stream of the reactor is very challenging as there is production of small
solid particles (coke) that can easily clog any controlling device. A mass flow controller that
operates reliably at the required temperature with very heavy hydrocarbons is not common for
research-scale prototypes. In addition, since the density of the product under the operating
conditions of the hot separator and the amount of light products that are carried to the cold
separation section are not known before collecting the products, the only way to close a mass
balance would be by controlling a liquid level in the hot separator and keeping it at a constant
value. Similarly, implementation of liquid level controlling devices for heavy oil applications for
the size of this equipment is not suitable. The alternative adopted in this design was to

15
incorporate a set of two computer-controlled pneumatic valves with a small 40 mL Swagelok
vessel in between. The set of valves is placed at the bottom of the hot separator as observed in
Figure 2.6. The pneumatic valves are timed in such a way that when the upper valve (V-301)
opens, approximately 90% of the small vessel fills up with liquid product and the remaining
space contains reaction gases and nitrogen. A detailed calculation procedure for the valves
timing is presented in the SOP of the unit in Appendix A. The 10 % remaining volume acts as
leverage for changes in the density of the liquid product and thermal expansion at the operating
temperature. This causes a small pressure drop in the unit, less than 3% of the operating pressure
which does not induce any type of instability in the reactor profile and does not compromise the
accuracy of the results as corroborated in the pilot plant commissioning presented in Chapter 3.
The upper valve is left open for 60 s and then it closes. Then, after 10 s, the lower valve (V-302)
is opened for 60 s and the product is released to the atmospheric collector. After this point, the
cycle starts again. In order to recover the pressure set point, a constant amount of nitrogen is
injected at the bottom of the hot separator. This stream also helps to strip the water and carry it to
the cold separator. In this way, the dilution of the reaction gases with nitrogen within the unit is
kept constant and the product collection is completely continuous. The operator has to be present
to drain the collector tank to start and close a mass balance, but no valve switching or
depressurization/pressurization is required.

Figure 2.6. RTU-1 continuous collection configuration in hot separator.

16
2.1.4. Cold Separation and Collection

In this section, represented by the schematic on Figure 2.7, gases are separated from the
liquid product and water after passing through a condenser. It is very similar in design to the hot
separation section. The main difference is that it operates at room temperature and the capacity
of the vessels is smaller since the sample amount collected is usually less than 10% wt. of the
material collected in the hot collectors. In this case, a 500 mL stability tank 2, a 150 mL mass
balance tank 3, and a 75 mL mass balance tank 4, all 304L stainless steel vessels supplied by
Swagelok, were installed. The product collection is carried out in the same way as explained in
Section 2.1.3, either in a semi-continuous or continuous mode. In continuous mode, a 10 mL
double-ended vessel is placed in between the pneumatic valves to drain the product from the cold
separator. Gases coming out from the cold separation zone pass through a back pressure valve
(BPV), which maintains the operating pressure at the given set point, and the gas stream is sent
to the gas release and depressurization zone.

Figure 2.7. RTU-1 cold separation section.

2.1.5. Gas Release and Analysis

As illustrated in Figure 2.8, this section is equipped with a condenser and a liquids trap to
collect any condensable liquid that may damage gas analysis equipment, two KOH traps for gas
sweetening, a Gas Chromatograph (GC) for gas composition analysis, a Shinagawa W-NK-0.5

17
Wet Gas Meter (WTM) for gas flow measurements, and also an extra trap to collect light
hydrocarbons and water flashed out when the collector tanks are depressurized during a mass
balance. A Quadrupole Mass Spectrometer (QMS) can be coupled to the outlet line of the BPV,
which operated at atmospheric pressure.

Figure 2.8. RTU-1 gas release and analysis section.

2.1.6. Control and Data Acquisition

A computer software programmed in LabVIEW code was created by technicians from


Electronics & Instrumentation at the Schulich School of Engineering at the University of Calgary
for control of all the instrumentation, including: heaters, thermocouples, mass flow controllers,
pumps, solenoid pneumatic valves, as well as for monitoring of pressure and temperature in the
unit. Continuous assistance was provided to the technicians for developing the software,
including the visual interfaces, controllers tuning, as well as to tailor the program to provide the
user with critical experimental information for data post-processing. A total of 25 120W heating
tapes provided by HTS Amptek were installed for temperature control and 3 were left as spare.
Each heating tape is connected to a Magnecraft Solid State Relay 25 (6225AXXSZS-DC3) and is
coupled with a wall thermocouple to measure the temperature of the heater. A
Proportional/Integral/Derivative PID control loop was programmed within the LabVIEW code to
control the output of each heater; control parameters (proportional term, integral term, and
derivative term) were individually tuned for each one of them and are reported in Table 2.2.
These tuning parameters are adjustable in the control program at any moment if required. A total
of 11 standard type K thermocouples supplied by Omega were placed along the process lines,
feed tank and vessels for measurement of the internal fluid temperature (TI), and an Omega
custom built thermowell with 7 sensing points was installed for capturing the temperature profile

18
along the reaction zone, for a total of 19 internal points for temperature monitoring and control.
A total of four 0-3000 psig high temperature pressure transmitters (PI) model MT104P3MXS-
3/8”D from MPI Melt Pressure, were placed along the reaction and hot separation zone, while
two 0-2000 psig low temperature transmitters model DPG1000L-2KG from Omega were placed
at the cold separation zone for measurement of the pressure at different points of the unit. The
two built-in pressure transmitters of the Teledyne ISCO feed pumps were used for monitoring
pressure in the feed section.
Table 2.2. Control parameters
Proportional Integral Derivative
Temperature Controller
Kc Ti (min) Td (min)
TC-101 Feed Tank 2.710 2.544 0
TC-102 Line to Pumps 0.899 1.602 0
TC-103 Pump B Cylinder 2.445 4.096 0
TC-104 Pump A Cylinder 2.309 3.175 0
TC-105 Pump B Valves 1.016 1.735 0
TC-106 Pump A Valves 1.030 2.466 0
TC-107 Pumps Discharge 0.942 2.526 0
TC-108 Pressure Relief Valve 3.131 5.005 0
TC-109 Line to Reactor 1 0.990 1.876 0
TC-201 Steam Generator 1.084 2.310 0
TC-202 Line to Furnace 0.905 1.442 0
TC-203 Furnace 1.030 1.691 0
TC-204 Reactor Bottom Section 0.988 1.917 0
TC-205 Reactor Middle Section 1.230 2.210 0
TC-206 Reactor Top Section 1.127 1.656 0
TC-301 Products Outlet 0.907 1.232 0
TC-302 Line to Stability Tank 0.858 1.614 0
TC-303 Line to Mass Balance Tank 0.901 2.048 0
TC-304 Stability Tank 3.604 2.726 0
TC-305 Mass Balance Tank 1 1.127 2.287 0
TC-306 Line from ST1 to Light Products Outlet 0.998 1.289 0
TC-307 Line from MB2 to Light Products Outlet 0.898 1.462 0
TC-308 Light Products Outlet 0.931 1.890 0
TC-309 Intermediate Vessel 1.000 0.500 0
TC-310 Mass Balance Tank 2 1.000 0.500 0

The data acquisition system coupled with the LabVIEW program consists of state of the art
hardware provided by National Instruments. The following list provides a brief description of all
the hardware installed in the pilot plant and a picture of the compact acquisition setup is
presented in Figure 2.9.

19
 NI PS-15 Power Supply, 24 VDC, 5 A, 100-120/200-240 VAC Input: power supply to
provide energy to the acquisition modules.
 Voltage/Current Analog Input, 500 S/s, 16 Ch module: current input module for
measuring signals from pressure sensors and mass flow controllers.
 NI 9203 8-Channel +/-20 mA, 200 kS/s, 16-Bit Analog Input Module: current input
module for measuring signals from mass flow controller – this unit had poor immunity to
electrical noise. It was adequate for reading the mass flow controller signal but could not
meet requirements for pressure measurement.
 NI 9401 8-Channel, 100 ns, TTL Digital Input/Output Module: module to turn Solid
State Relays on/off as required.
 NI 9265 4-Channel 20 mA, 100 kS/s per Channel, 16-Bit, Current Output Module:
module to send set point to mass flow controllers.
 NI 9213 16-ch TC, 24-bit C Series module: thermocouple input module to measure
temperatures.
 cRIO-9074 CompactRIO Controller and Chassis Integrated System, 400 MHz Power PC
controller, 2M Gate FPGA, 8-slots: Real Time Controller – software is loaded into this
unit to measure and control the plant. This system receives set points from host computer
and send acquired data to the host computer. The unit is similar to a Programmable Logic
Controller (PLC).
 NI 9144 8-slot Deterministic Ethernet Chassis for C Series Modules: expansion chassis
for above system – allows the data acquisition system to have more than the 8 modules
that can be installed in the cRIO-9074.
 NI 9924 25 pin Dsub terminal block for screw terminal connectivity to 25 pin Dsub C
Series modules: terminal blocks for some of the measurement/control modules.
The LabVIEW program that controls all the instrumentation of the unit is subdivided in
7 interactive sections: feed section, reaction section, heavy products, light products, heaters
and pump control, alarm settings, and heater tuning. The program is set to measure all the
operation data every 5 seconds and records it in an excel spreadsheet every minute for post-
processing.

20
Figure 2.9. Hardware for data acquisition and control.

2.2. Characterization Techniques

In this section, the most relevant characterization techniques used for analysis of feedstock,
UDC feedstocks, fixed-bed catalysts, as well as the liquid and gas products from the reactions
carried out throughout the development of this doctoral investigation are explained in detail.
Most of these techniques are modifications of ASTM methods commonly used for heavy oil and
catalyst characterization that have been adapted for a feasible implementation with the resources
available at the CBU group at the University of Calgary, while keeping comparable levels of
accuracy and reproducibility as those of ASTM methods. It is important to highlight that due to
the paper-based structure of this thesis, all these techniques are briefly described again in each
manuscript/article-type chapter where they were used.

2.2.1. Ultradispersed Catalyst Characterization

2.2.1.1. Metal Content

Determining the amount of catalyst particles incorporated in the DAO matrix is of


paramount importance to corroborate its adequate preparation. For this purpose, the Inductively
Coupled Plasma (ICP) technique as indicated on the ASTM D7691-05 norm was used. A
Thermoelectron spectrometer model IRIS Intrepid II was used for the analysis. Preparation of the
sample via microwave assisted digestion (ASTM D7455) before the ICP analysis was needed.
For this purpose, an amount of sample between 0.25 to 0.30 g was added to a teflon digestion

21
cell along with 10.5 mL of nitric acid (70% Aldrich) and 1 mL of an aqueous solution of
phosphoric acid 85% wt. for samples containing tungsten. When determining cerium
concentration as required for the work presented in Chapter 6, the digestion was done using only
nitric acid, since phosphoric acid reacts with cerium to form insoluble salts that are not properly
digested in the acid media, which hinders the accurate determination of the concentration of this
metal. The closed teflon cells were placed in a CEM microwave digester model MARS6 and
were heated to 210 °C during 20 min using a heating ramp of 10 °C/min, always ensuring that no
vapor leaks were observed from any of the cells. After digestion, the content of the cell was
carefully transferred to a 25 mL volumetric flask using a disposable plastic pipette and it was
diluted with distilled water until completing the 25 mL. A sample of the solution was then
injected to the ICP equipment where a plasma flame decomposes all the materials to its basic
elemental atomic forms. As indicated by Scott2, atoms pass to excited electronic states due to the
energy supplied by the plasma flame and, upon returning to their fundamental state, they emit
photons at characteristic wavelengths for each element. Emitted light is dispersed at different
wavelengths and is detected in the optical system of the equipment to generate spectra. Each line
in the spectra and its position corresponds to a certain wavelength and a metal, and its intensity is
proportional to its concentration in the injected solution. Standard solutions of known
concentration of metals of interest are injected to the equipment and the intensity of the signals
are correlated to its concentration, which allows the user to estimate the actual metal
concentration in the sample solution and consequently in the analyzed DAO sample.

2.2.1.2. Particle Size

Measurement of the particle size is critical for assessment of catalytic activity of UDC since
this parameter has to be as close as possible to nanometric range (<100 nm) in order for the
catalyst to have a comparable catalytic activity as that of a fixed-bed catalyst.3 Before 2015, no
reliable technique was available for determining the actual particle size diameter of catalyst
particles suspended in heavy oil matrices. Researchers4 were measuring particle sizes using
Dynamic Light Scattering (DLS) and suspending the catalyst particles on colorless base oils, and
not on the actual feedstock. In 2015, Rodriguez-DeVecchis et al.5 developed a method called
Nanoparticle Tracking Analysis (NTA), using a NanoSight model NS300 equipment. In this
technique, the particles suspended in diluted heavy oil samples are detected and identified at the

22
given visual range of the equipment through light dispersion as a consequence of irradiating
them with a laser. After this stage, the equipment follows the movement of each particle during a
certain time and calculates the hydrodynamic radius following Stokes Law (Eq. 2.1) for a
subsequent estimation of the bulk particle size based on the analysis of all visualized particles.

̅̅̅̅̅̅̅̅̅ 2𝑘𝑇𝑡
(𝑥, 𝑦)2 = Eq. 2.1
3𝑟𝜂

Where 𝑥 and 𝑦 are the position coordinates, 𝑘 is the Boltzmann constant, T is the
temperature, 𝑡 is the total time a singular particle is tracked, 𝜂 is the viscosity of the medium, and
𝑟 is the hydrodynamic radius.

In this method, 0.5 g of sample diluted in 9.5 g of toluene (99.9% Aldrich) in glass vials are
submitted to mechanical agitation during 10 min in a Burrel model 75 shaker. Next, samples are
sonicated during 10 min in a Branson 3510 sonicator. Afterwards, the sample chamber of the
equipment is purged by injecting a portion of the diluted and homogenized sample corresponding
to a volume of approximately 1 mL and the focus of the equipment is adjusted for a proper
visualization of the particles. After that, 5 videos of the moving particles were taken for analysis,
always ensuring replacement of the sample in the sample chamber, and then the results were
analyzed for the prepared dilution. This procedure was repeated for 4 more sequential dilutions
prepared by diluting 1 g of the former solution and adding 9 g of toluene each time. The particle
size was estimated as the average of the best dilutions analyzed.

2.2.2. Characterization of Feedstock and Liquid Products

2.2.2.1. Product Distribution

The economic value of a certain crude oil is directly related to its product distribution. Crude
oils with large amounts of heavy compounds are difficult to process and require specific deep
conversion processing or upgrading in order to provide added-value. Commonly, due to the large
amount of compounds that are present in a given crude oil, products are classified as pseudo
components or lumps defined by their boiling point range. Each petroleum producer or refiner
has their own classification. For the purpose of this work, the following liquid products were
defined: naphtha is the fraction that boils between 28-190 °C, kerosene 190-260 °C, diesel 260-
343 °C, light vacuum gas oil (LVGO) 343-453 °C, heavy vacuum gas oil (HVGO) 453-560°C,

23
and vacuum residue (VR) boils at temperatures above 560°C. VR can be subdivided into
deasphalted oil (DAO) 560 °C+, and pentane (nC5) insoluble asphaltenes. The amount of each
lump is obtained from a distillation curve that relates boiling point with the amount of crude oil
distilled.

In this work, the Simulated Distillation (SimDist) characterization technique was used for
the quantification of the liquid products distribution following an in-house modified version of
the ASTM norm D-7169-05 by Carbognani et al.6 Instead of injecting 0.2 µL of the original
sample, 1 µL of a filtered solution prepared by diluting 0.15 g of sample with 20 mL of carbon
disulfide (CS2) was injected for analysis in an Agilent 6890N chromatograph equipped with
automatic injection, a PTV injection port and a 5 m x 0.53 mm capillary column with a 0.1 µm
film methyl silicone stationary phase (Ref. P/N SS 112-102-01 from Separation Systems Inc).
This modification reduces the error caused by volumetric injection of the sample from 25% to
5% as well as reduces the relative error induced by the presence of nanoparticles in the sample.
For example, the presence of a 100 nm particle results in an error of 20% when injecting only 0.2
µL of sample. Chem-Station software (Rev. A. (10.02) 1757) was used to control the
chromatographic events. In addition, SimDist Expert v.8 software, also by Separation Systems
Inc., was used for data post-processing.

2.2.2.2. Viscosity

Viscosity is one of the most important parameters when characterizing a crude oil sample
and is a measure of the resistance to flow of a fluid, thus it is related to its transportability. As
explained in the introduction of this work, virgin bitumen is not transportable. For this reason, it
has to be either blended with diluents or upgraded into a synthetic crude oil in order to reach
Canadian transportability specifications, among which viscosity is one of the main factors (API:
19; viscosity: 350 cSt at carrier reference temperature; olefins: < 1% wt 1-decene equivalent). In
this work, viscosity was measured in a Brookfield viscometer model DV-II+ Pro that is equipped
with a water recirculation system model TC-502 for measurements at temperatures between 0 to
100 °C. The measurement starts by setting up the temperature controlling system at the required
temperature. In this work, all measurements were carried out at 25 °C, except for the feedstocks,
which were measured at 60 and 100 °C. Next, the measuring device (spindle) was screwed to the
bottom of the rotatory shaft and the viscometer was closed with the sample cell, adjusting the

24
distance between the spindle and the bottom of the cell to 0.1 mm. Then, an amount of sample
enough to cover the whole surface of the spindle was placed in the sample cell. After closing, the
rotation engine was started and the velocity was adjusted in such a way that the shear generated
by the cohesive forces between the fluid and the metal plates measured by the equipment’s
dynamometer was within the 50-70% range. Rotation was continued until completing at least 5
full rotations of the spindle to guarantee proper formation of the fluid film, always ensuring that
the temperature set-point was at the required value since small variations significantly affect the
measurement. Finally, the viscosity value (in cP) was reported in the equipment data book.

2.2.2.3. API Gravity

API gravity is a representation of the specific gravity (S.G.) of the fluid in a more sensitive
scale and is defined by Eq. 2.2 from the ASTM D287-92 norm.

141.5
𝐴𝑃𝐼 = − 131.5 Eq. 2.2
𝑆. 𝐺.

It is a very important parameter when assessing the transportability of a crude oil as shown
in the criteria presented in the previous Subsection 2.2.2.2. Moreover, as for viscosity, it is an
indication of the nature of the chemical compounds present in the crude oil since heavier
compounds such as asphaltenes have considerably lower API, which decreases the economic
value of the oil.7

Specific gravity was determined using a digital densitometer, Rudolph Research Analytical
model DDM2911, as described by Carbognani et al.8 Since it is not possible to inject heavy
viscous samples directly in the equipment cell, two diluted sample solutions at concentrations
close to 0.5 % v/v and 1% v/v were analyzed and the density of the pure sample was extrapolated
by assuming an ideal mixture for diluted solutions of components with similar polarity and
additive volumes as represented mathematically in Eq. 2.3.

𝑚𝑠𝑎𝑚𝑝𝑙𝑒
𝜌𝑠𝑎𝑚𝑝𝑙𝑒 = Eq. 2.3
(1⁄𝜌𝑡𝑜𝑙𝑢𝑒𝑛𝑒 )(𝑚𝑠𝑎𝑚𝑝𝑙𝑒 + 𝑚𝑡𝑜𝑙𝑢𝑒𝑛𝑒 ) − (1/𝜌𝑡𝑜𝑙𝑢𝑒𝑛𝑒 )𝑚𝑡𝑜𝑙𝑢𝑒𝑛𝑒

25
2.2.2.4. Sulfur and Nitrogen

Sulfur and nitrogen have significant presence in heavy feedstocks as sulfided or nitrogenated
compounds. This complicates processing of these feedstocks, mainly in catalytic processes since
these compounds are active sources of catalyst deactivation. Also, their potential production of
NOx and SOx, which are gases highly harmful to the environment, is a limiting factor when
processing crude oils with high content of these compounds. In this way, their presence in high
proportions has a strong influence on the commercial value of a crude oil.7,9

Sulfur and nitrogen contents were determined in an Antek 9000NS analyzer.10 An amount of
0.2 mg of sample and 2.8 g of toluene were weighed in a glass vial. The vial was placed in the
equipment that injects 0.2 µL of sample inside a furnace to conduct pyrolysis at a temperature of
1000 °C, having an oxygen atmosphere to guarantee complete conversion of nitrogen and sulfur
to their corresponding oxides NO2 and SO2. To determine the amount of sulphur, the gas effluent
is irradiated with light at a given wavelength and the SO2 absorbs a photon and passes to an
excited state. When relaxing to its fundamental state, the molecule emits a photon at a specific
energy different from the irradiation energy, whose intensity is related to the amount of sulphur.
On the other hand, to determine the amount of nitrogen, the gas stream is put in contact with
ozone, which reacts with NO2 to form NO2 in an excited state. This excited NO2 releases a
characteristic photon when decaying to its fundamental state and the intensity of the emitted light
is proportional to the amount of nitrogen present in the sample, which is calculated based on
previous calibration of the equipment.10

2.2.2.5. Asphaltenes Stability Index (P-value)

P-value is a measure of the state of peptization of the asphaltenes present in the sample. This
method, described in detail by Di Carlo and Janis11, relies on determining the maximum dilution
of the sample with hexadecane at which asphaltenes are not precipitated when observed in an
optical microscope with a magnification of 40X. P-value is calculated following Eq. 2.4.

𝐻𝑒𝑥𝑎𝑑𝑒𝑐𝑎𝑛𝑒 [𝑚𝑙]
P − value = 1 + Eq. 2.4
1 𝑔 𝑠𝑎𝑚𝑝𝑙𝑒

In this method, 1 g of sample was added to a 20 mL griffin beaker with a 12 mm magnetic


stirrer. The desired amount of hexadecane, calculated by assuming a P-value and determining the

26
volume of hexadecane from Eq. 2.4, was added using a 10 mL burette with subdivisions of at
least 0.05 mL. A micropipette could also be used if more accuracy is needed. A condenser tube
was coupled with a rubber plug to the top of the beaker to minimize hexadecane loses upon
heating. Next, the griffin beakers were immersed in a glycerin bath maintained at 100 °C for 30
min to guarantee enough contact between the hexadecane and the oil sample; this is called the
digestion period. After this stage, a clip was used to stir the contents of the beaker and a tiny
drop was transferred to a microscope slide and covered with its corresponding cover slip. Finally,
the sample was inspected in an optical digital microscope model DC3-163 provided by National
Instruments, coupled with a computer-controlled camera with the software Motic Images Plus
2.0, with a magnification of 40X for visual confirmation of precipitated asphaltenes. It is
important that the microscope slide is hot in order to melt paraffins that may interfere with the
visual determination of flocculated asphaltenes. A sample was defined as precipitated at the
evaluated P-value following the criteria presented in Table 2.3. In addition, an example of the
size of the asphaltene agglomerates is presented in Figure 2.10.

To determine the actual P-value of one sample, several sample dilutions were determined
simultaneously. The limiting factor for the number of determinations that can be made at the
same time was the number of griffin beakers that can be properly placed in the hot oil bath. For
this work, it was possible to determine six samples at the same time.

27
Table 2.3. Microscopic images for P-value determination

Description/Observations Microscope Image

P-value 1.0 (Blank, no n-hexadecane added)

No significant presence of precipitated asphaltenes or


solids is evidenced

P-value 1.50

No significant presence of precipitated asphaltenes is


observed

P-value 1.55

Asphaltenes are starting to form agglomerates that are


not clearly seen in the microscope image

P-value 1.60 (0.60 mL of n-hexadecane added)

Presence of precipitated asphaltenes networks is clearly


observed. This is defined as P-value of the sample
(operator dependent).

P-value 1.65

Presence of precipitated asphaltenes is observed

28
Figure 2.10. Asphaltene agglomerates size in precipitated sample in P-value analysis.

Significant presence of solids in the sample can hinder the P-value determination since it
would not be possible to differentiate them from precipitated asphaltenes. To address this issue,
it is important to observe the sample without hexadecane dilution (blank) to identify solids
presence before doing the analysis. In case significant presence of solids is evidenced, the
organic part of the sample has to be diluted with a solvent such as chloroform (CHCl3) and then
filtered in a vacuum filtering unit and finally it has to be rotoevaporated to remove the solvent
added before doing any determination. Table 2.4 shows an example of a sample with high
presence of solids.

Table 2.4. Microscope images for P-value analysis of samples with high presence of solids

Blank Blank (Filtered) P-value = 1.15


(Filtered)
Significant presence of No precipitated asphaltenes
solids in sample in sample Precipitated

29
2.2.2.6. Microcarbon Residue

Microcarbon Residue (MCR) or Conradson Carbon is an indirect measurement of how


prone a crude oil is to forming coke. This property can be correlated with catalyst deactivation
by coke deposition and pore plugging, and the tendency of an oil to form solids and fouling when
submitted to thermal processing, among other factors. In this way, higher values in MCR have a
detrimental effect on the economic value of a crude oil.

MCR was determined following the muffle furnace method developed by Hassan et al.12,
which is an alternative to ASTM methods ASTM D-189, D-524 and D-4530 that considerably
reduced the analysis time per sample. In this method, approximately 20 mg of sample were
weighed in a 2 mL glass vial on a 5 digits Mettler Toledo XS 205 balance. Vials were placed on
the sample platform of the MCR assembly inside the muffle furnace. This platform can
accommodate up to 26 samples and is equipped with 26 nitrogen injection tubes (1 per sample)
to create an oxygen free environment. Then, vials were covered using a glass cover with a 1/8”
orifice in the middle, the muffle furnace was closed, and nitrogen flow was started at 50 mL/min
during 45 min to guarantee the absence of oxygen inside the furnace, which would result in an
incorrect analysis. The next step was to start heating using a heating ramp of 10 °C/min until
reaching a temperature of 520 °C that was maintained for 20 min. After heating, the furnace was
let to cool down and the samples were placed in a numbered sample holder for weighting. MCR
weight percentage was calculated following Eq. 2.5. Each sample was analyzed in duplicate.

𝑀𝑎𝑠𝑠 𝑜𝑓 𝑠𝑎𝑚𝑝𝑙𝑒 𝑎𝑓𝑡𝑒𝑟 ℎ𝑒𝑎𝑡𝑖𝑛𝑔 [𝑔]


𝑀𝐶𝑅 [%] = ∗ 100 Eq. 2.5
𝑀𝑎𝑠𝑠 𝑜𝑓 𝑠𝑎𝑚𝑝𝑙𝑒 [𝑔]

2.2.2.7. Saturates, Aromatics, Resins and Asphaltenes (SARA)

Hydrocarbon SARA group type distribution separation analysis was carried out following
the procedure described by Carbognani et al.13 This method is divided into two sections. The first
is the microdeasphalting of the sample to determine the content of pentane (nC5) insoluble
asphaltenes. The second part is the analysis of the Saturates-Aromatic-Resins group types for the
maltene phase (deasphalted oil) using thin layer chromatography with flame ionization detection
(TLC-FID). For microdeasphalting, 0.4 g of sample weighted to the nearest 0.1 mg were placed
in a 100 mL beaker along with 20 mL of n-pentane (ratio of 50/1 vol. nC5/wt. sample). The

30
beaker was then covered with a Petri dish in order to minimize evaporation of the paraffinic
solvent and is subsequently placed over a heating plate maintained at a temperature of 50 °C
during 30 min for digestion of the sample. During this period, the content was gently mixed
several times with a metallic spatula to promote formation of the solid asphaltene particles.
Pentane was added as needed to maintain a constant solvent content. Afterwards, the mixture
was cooled down to ambient temperature and the precipitate was filtered out by vacuum filtration
in a 47 mm Millipore unit provided with 0.45 µm nylon membranes from Pall Corporation. The
membrane with the precipitated solids was placed in a Petri dish inside an oven operated at 80
°C, left for 5 min to dry out any remaining amount of solvent. Then, the membrane was cooled
down to ambient temperature and weighted to quantify the amount of nC5 asphaltenes in the
sample. In the SAR analysis section, the maltene plus solvent fraction is rotoevaporated to
remove the nC5 using a round bottom 100 mL flask in a rotatory evaporator provided with
nitrogen stripping to create an inert environment, avoiding oxidation of resins. Then, the
collected maltene was diluted with toluene (99.9 % Aldrich) in order to get a concentration of 10
mg of maltenes per mL of toluene. SAR distribution for the maltene fraction was subsequently
conducted in an Iatroscan model MK-6 chromatograph controlled with a Peak Simple
chromatographic data system. S-III silica chromarods activated by burning at least 5 times were
used for the analysis. One microliter of the diluted maltene solution was spotted at the bottom of
at least three rods per sample (triplicate analysis). Next, the chromarods were placed in a TLC
chamber filled with 100 mL of heptane (Aldrich) and the Saturates fraction was eluted up to 10
cm in the chromarods during approximately 30 min. Subsequently, chromarods were placed in a
different TLC chamber filled with 100 mL of toluene (Aldrich) for elution up to 6 cm to separate
the Aromatics fraction for approximately 15 min. The Resins fraction remains at the bottom of
the rods. The next step was to scan the chromarods in the Iatroscan chromatograph FID operated
with a mixture of air and hydrogen at flowrates of 2 mL/min and 160 mL/min respectively, with
a scanning speed of 30 s/rod. The contents of SAR HC group types were estimated in the Peak
Simple software applying a pre-established response factor to the Resins fraction and
renormalizing.

31
2.2.3. Characterization of Gas Products

2.2.3.1. Gas Chromatography

In order to close the mass balance, it is required to calculate the total amount of hydrocarbon
and permanent gases generated during the thermal and catalytic reactions. For this, the molar
composition of this stream has to be determined. This analysis was conducted via Gas
Chromatography (GC) in a SRI Instruments chromatograph model 8610C equipped with 4
packed columns, 2 switching valves, 2 thermal conductivity detectors (TCD), 1 flame
photometric detector (FPD), and one flame ionization detector (FID). The first TCD is operated
using helium as carrier gas with a flow of 20 mL/min to detect light hydrocarbons within the C1-
C5 range, hydrogen at high concentrations, and the remaining permanent gasses (CO2 and CO)
with a set of a 183 cm molecular sieve column (model MS13X) and a silica gel column of the
same length. Due to the proximity in conductivity between H2 and He, it is difficult to detect low
concentrations of H2 using He as carrier. For this reason, the second TCD used argon carrier
instead, with a flow of 10 mL/min to detect hydrogen at low concentrations using a 3 feet
molecular sieve column (model MSX13X) for separation. The following oven temperature
profiles were used for gas separation in the columns: initial temperature of 35 °C maintained for
5 min, subsequent increase of 15 °C/min up to 215 °C, and constant at 215 °C for 20 min. Sulfur
compounds (H2S and SO2) are analyzed with the FPD detector operated with a mixture of
hydrogen and air at flowrates of 42 mL/min and 150 mL/min respectively using a 60 m capillary
column (model MXT-1) for separation. The FID was used to quantify the amount of light
hydrocarbons within the C1-C5 range to crosscheck the results obtained from the first TCD. The
FID was operated with a combustion mixture of hydrogen and air at flowrates of 25 mL/min and
250 mL/min respectively, using the same column as in the FPD.

2.2.3.2. Mass Spectrometry

Quadrupole Mass Spectrometry (QMS) is a characterization technique used to determine


molecular weights of chemical species present in a gas stream. Furthermore, unlike GC where
each analysis can take around 1 h, QMS provides continuous measurements. In this way, it is
very practical to identify changes in the concentration profiles as soon as any change in the
reactions takes place. In principle, this method determines the mass to charge ratio (m/e) of the
gas stream constituents after ionization.14 In the works described in Chapter 5 and 6, a Pfeiffer

32
OmniStar QMS described in detail by Fathi and Pereira-Almao15 was connected inline for certain
experimental conditions under catalytic steam cracking and thermal cracking processing. This
benchtop equipment consists of a heated inlet system to prevent condensation of water and light
hydrocarbons in the capillary gas inlet tube, a PrismaPlus mass spectrometer, a dry-compressing
diaphragm vacuum pump, and a HiPace turbopump. The equipment samples the gas stream at
atmospheric pressure after the back pressure valve, and it goes to the vacuum chamber, where it
is bombarded with electrons for components ionization. Next, the ions are separated by mass to
charge ratio, for further detection in a Faraday detector or a secondary electron multiplier. This
detector measures a specific current for each ion, which is correlated to its partial pressure and
concentration in the gas phase.16 The equipment is computer-controlled with a software that
allows scanning of atomic masses ranging from 1 to 300 amu as well as post-processing of data.

2.3. References

(1) Trujillo-Ferrer, G. Thermal and Catalytic Steam Reactivity Evaluation of Athabasca Vacuum
Gasoil. MSc Thesis, University of Calgary, Calgary, 2008.
(2) Scott, C. E., Chemical Reactors Design, Catalyst Characterization Lecture. University of
Calgary: Calgary, AB, 2012.
(3) Scott, C. E.; Pereira-Almao, P. Catalysis for heavy oils and bitumen upgrading. Current
Topics in Catalysis 2014, 11, 1-24.
(4) Vasquez, A. Synthesis, Characterization and Model Reactivity of Ultra Dispersed Catalysts
for Hydroprocessing. MSc Thesis, University of Calgary, Calgary, 2007.
(5) Rodriguez-DeVecchis, V. M.; Carbognani Ortega, L.; Scott, C. E.; Pereira-Almao, P. Use of
Nanoparticle Tracking Analysis for Particle Size Determination of Dispersed Catalyst in
Bitumen and Heavy Oil Fractions. Ind. Eng. Chem. Res. 2015, 54 (40), 9877-9886.
(6) Carbognani, L.; Lubkowitz, J.; Gonzalez, M. F.; Pereira-Almao, P. High Temperature
Simulated Distillation of Athabasca Vacuum Residue Fractions. Bimodal Distributions and
Evidence for Secondary “On-Column” Cracking of Heavy Hydrocarbons. Energy Fuels 2007, 21
(5), 2831-2839.
(7) Gray., M. R., Upgrading oilsands bitumen and heavy oil. The University of Alberta Press:
Edmonton, 2015.
(8) Carbognani Ortega, L.; Rogel, E.; Vien, J.; Ovalles, C.; Guzman, H.; Lopez-Linares, F.;
Pereira-Almao, P. Effect of Precipitating Conditions on Asphaltene Properties and Aggregation.
Energy Fuels 2015, 29 (6), 3664-3674.
(9) James, G. S., Feedstock Evaluation and Composition. In Hydroprocessing of Heavy Oils and
Residua, CRC Press: 2007; pp 15-33.
(10) Antek - MODEL 9000 Nitrogen/Sulfur Analyzer (PAC). Evisa, Ed. 2010.
(11) Di Carlo, S.; Janis, B. Composition and visbreakability of petroleum residues. Chem. Eng.
Sci. 1992, 47 (9), 2695-2700.

33
(12) Hassan, A.; Carbognani, L.; 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.
(13) Carbognani, L.; Gonzalez, M. F.; Pereira-Almao, P. Characterization of Athabasca Vacuum
Residue and Its Visbroken Products. Stability and Fast Hydrocarbon Group-Type Distributions.
Energy Fuels 2007, 21 (3), 1631-1639.
(14) de Hoffmann, E.; Stroobant, V., Mass Spectrometry: Principles and Applications. Wiley:
2007.
(15) Fathi, M. M.; Pereira-Almao, P. Catalytic Aquaprocessing of Arab Light Vacuum Residue
via Short Space Times. Energy Fuels 2011, 25 (11), 4867-4877.
(16) Garcia Hubner, E. Catalysts Evaluation for Catalytic Steam Cracking of De-asphalted Oil in
a Fixed Bed Reactor. MSc Thesis, University of Calgary, Calgary, 2015.

34
Chapter 3: Pilot Plant Commissioning

For the purpose of this work, the RTU-1 was designed for thermal and catalytic steam
cracking of DAO using ultradispersed catalysts. In order to verify the adequate performance in
terms of stability, safety, control, and other relevant operational parameters when using the
experimental setup with the process of interest, a preliminary experimental evaluation was
conducted using an in-house prepared DAO obtained by bench-scale pentane solvent
deasphalting of a 250 °C+ fraction of South American heavy oil with the characteristics shown in
Table 3.1. Even though it is not the feedstock used for the core experimental work of this
investigation, which was not available at the moment of the commissioning of the unit, it is very
similar and can be used for testing purposes. The semi-continuous mode of product collection
was used in this first stage of commissioning. When the continuous collection system explained
in Section 2.1.3.2 was integrated to the unit in 2015, the experimental setup was recommissioned
using an industrial DAO feedstock provided by Nexen Energy ULC that was used throughout the
development of this doctoral work. A detailed description of the characteristics of this industrial
feedstock is provided in Section 4.3.2.

Table 3.1. South American DAO feedstock characterization


PRODUCTS DISTRIBUTION [% wt.]
Naphtha (28 - 190 °C) 0.0
Kerosene (190 - 260 °C) 2.3
Diesel (260 - 343 °C) 11.7
LVGO (343 - 453 °C) 27.2
HVGO (453 - 560 °C) 23.3
Vacuum Residue (560°C+) 35.6
PROPERTIES
API Gravity 11.2
MCR [%wt] 9.34
Viscosity @25°C [cP] 19297
Hydrogen (H) [%wt] 11.8
Carbon (C) [%wt] 85.3
Saturates [%wt] 20.8
Aromatics [%wt] 55.1
Resins [%wt] 17.4
Asphaltenes-C5 [%wt] 6.7

35
3.1. Experimental Plan

3.1.1. Semi-continuous Product Collection

As explained in section 2.1.3.1, the first stage of design of the pilot plant was based on a
semi-continuous product collection. Under this mode of operation, thermal cracking of a South
American DAO using 5 %wt. of steam was studied (5% wt. of the hydrocarbon volumetric
flowrate multiplied by its density at standard conditions). Four conditions were tested at a LHSV
of 2h-1: condition 1 at 415 °C, condition 2 at 420 °C, condition 3 at 425 °C, and condition 4 at
423 °C, collecting two mass balances, each at 60 min intervals for each evaluated condition, and
at a reaction pressure of 300 psig.
For the commissioning under DAO catalytic steam cracking conditions, the Ni/K catalyst
matrix was incorporated into the feed using a skid preparation unit as described elsewhere1,
targeting a concentration of 300 ppm of Ni and 400 ppm of K. As in the thermal cracking case, 5
% wt. of water was used, and a total of 5 conditions were tested, using as initial temperature the
last temperature used in thermal cracking (423 °C). The other conditions were set along the run
based on the conversion obtained. In this way, 4 of the 5 catalytic reaction conditions were
established; these are: condition 5 at 423 °C and LHSV 2 h-1, condition 6 at 427 °C and LHSV 2
h-1 , condition 7 at 425 °C LHSV 2 h-1, condition 8 at 430 °C LHSV 2.5 h-1. Also, it was decided
to carry out another reaction at 435 °C and LHSV 2 h-1 in a run carried out separately. For each
condition, two mass balances of 60 min were collected. The pressure of the test (300 psig) was
the same for all of the evaluated conditions. The final stage of the catalytic evaluation, it was
decided to test the reactivity of the DAO at the extreme condition of 440 °C and LHSV 2 h-1
using the same feedstock. First, the previous run at 435 °C was validated, and then the reaction at
440 °C was performed. Similarly to the previous evaluations, for each condition two mass
balances of 60 min were collected.

3.1.2. Continuous Product Collection

The continuous collecting system was incorporated to the RTU-1 after one year of operation
(2015). In order to corroborate the stability of the unit when operated in this mode and to
evaluate the possible effect that the pressure drop caused by the drainage of small amounts of
liquid product from the separators would have on the reactor temperature and pressure profiles
which would reflect on the quality of the produced data, a second stage of commissioning was

36
conducted. In order to do this, two scenarios were evaluated under thermal cracking processing
conditions: a high and a low LHSV. Since the reaction temperature cannot be the same for the
two conditions because the severity of the reaction would change drastically and coking might
occur if the low LHSV is run at a high temperature as may be needed for the high LHSV case,
both experiments were run at different temperatures. The first one at a high temperature of 423
°C, and the second one at low temperature for thermal cracking of 380 °C. A stabilization period
was allowed for each run before collecting any mass balance. A total of 3 mass balances were
collected for each run. The duration of each mass balance was equal to 2 times the inverse of the
LHSV for each reaction in order to collect a similar amount of sample for mass balances. The
pressure and temperature were monitored carefully during the experimental runs and SimDist
analysis of the products was carried out to evaluate the reproducibility and consistency of all the
mass balances in this mode of operation. The timing of the pneumatic valves cycle for the
continuous collection, calculated using the guidelines established in the SOP of the unit included
in Appendix A, were as follows: 7.5 and 60 min for LHSV of 2 h-1 and 0.25 h-1 respectively.

3.1.3. Products Characterization

Analytical characterization of liquid products was carried out for the first mass balance of all
the conditions. In order to check the reproducibility of the experimental runs, the second mass
balances of condition 4 and 9, which were the most severe conditions providing stable products,
were used. The following characterization techniques described in detail in Section 2.2 were
used for the reactivity evaluation and commissioning using the semi-continuous collection
system: SimDist, MCR, SARA, P-value, API gravity, viscosity and GC analysis. For the
continuous collection system, since the purpose was to corroborate the stable operation of the
unit, only SimDist was used.

3.2. Results

3.2.1. Semi-continuous Product Collection

As can be seen in Figure 3.1, where the reaction temperature and pressure as a function of
time are shown for the whole thermal cracking commissioning run, the operation was very
smooth, without any significant variation on neither pressure nor reaction temperature during the
experiments. This indicates a satisfactory behavior of the unit in terms of operation and stability.

37
For each condition, a stabilization period equal to 3.5 times the inverse of the LHSV was used
before collecting the mass balances.

Figure 3.1. Reactor pressure and temperature profiles for thermal cracking experiments
(condition 1 to 4).

Figure 3.2 shows the reaction temperature and pressure as a function of time for the catalytic
runs (C5-C9). In the same way as in the steam cracking runs, there was no significant change in
both monitored variables, and the operation of the unit was very smooth.

Figure 3.3 shows the temperature profile for each one of the 5 sensing points along the
tubular reactor for the catalytic run C9. As illustrated, there are no significant instabilities during
the experimental run for any of the reactor sections (bottom, middle and top).

38
Figure 3.2. Reactor pressure and temperature profiles for catalytic steam cracking experiments
(condition 5 to 9).

39
Figure 3.3. Detailed reactor temperature profile for catalytic steam cracking experiment
(condition 9).

In the temperature profiles for the stability period and mass balances of the catalytic
reactions C10 and C11 presented in Figure 3.4, a stable operation at 435 °C was observed as in
the previous case (C9), with a target reaction temperature and pressure at the established set
point for the respective condition. Conversely, soon after the temperature reaction was re-
adjusted to reach 440 °C for the last evaluated condition (C11), a sharp drop of the temperatures
at the bottom of the reactor was seen (refer to Figure 3.4), believed to be a result of the
deposition of coke on the reactor or thermowell walls (coke fouling), followed by a continuous
smooth decrease on the overall reactor temperatures, which is also believed to have been caused
by the activation of the endothermic coke producing (condensation) reactions. Due to these
instabilities evidenced at this level of severity, the product from condition 11 was not
characterized since the sample was not representative of a stable process from an operational
point of view.

40
Figure 3.4. Reactor temperature profile for catalytic steam cracking experiments (condition 10
and 11).

Table 3.2 and Table 3.3 present a brief statistical analysis of the temperature and pressure
variations respectively for the evaluated thermal and catalytic conditions during the stability and
mass balance periods. An average relative variance of 0.12 % and 0.28 % was obtained for the
temperature and pressure respectively. As an example, this means than for a base case of a
reaction temperature of 425 °C, the temperature fluctuated within a range of ±0.5 °C around the
set-point. In the same way, considering a base case of a reaction pressure of 300 psig, the
pressure fluctuated within a range of ±0.8 psig around the set-point. This is a very good
indication of the tight control of the operating conditions for the designed equipment.

41
Table 3.2. Statistical analysis of temperature variation for thermal and catalytic experiments

Average Average
Standard Relative
Target Measured Relative
Experiment Deviation Variance
Temperature[°C] Temperature Variance
[°C] [%]
[°C] [%]
Condition 1 415 415.2 0.54 0.13
Thermal Condition 2 420 420.1 0.67 0.16
Runs Condition 3 425 425.0 0.32 0.07
Condition 4 423 423.0 0.27 0.06
Condition 5 423 423.0 0.62 0.15 0.12
Condition 6 427 427.3 0.52 0.12
Catalytic
Runs Condition 7 425 424.7 0.55 0.13
Condition 8 430 430.4 0.38 0.09
Condition 9 435 434.8 0.71 0.16

Table 3.3. Statistical analysis of pressure variation for thermal and catalytic experiments

Average Average
Target Standard Relative
Measured Relative
Experiment Pressure Deviation Variance
Pressure Variance
[psig] [psig] [%]
[psig] [%]
Condition 1 299.8 2.09 0.70
Thermal Condition 2 299.2 0.82 0.28
Runs Condition 3 298.5 0.47 0.16
Condition 4 297.7 0.76 0.26
Condition 5 300 300.0 0.85 0.28 0.28
Condition 6 299.7 0.53 0.18
Catalytic
Runs Condition 7 298.2 0.57 0.19
Condition 8 299.4 0.59 0.20
Condition 9 301.0 0.98 0.33

In Table 3.4, a summary of the results for both thermal cracking (Section A) and catalytic
steam cracking (Section B) is presented, including reaction conditions and the analytical tests
used to monitor the performance of the process and product quality. As illustrated in Table 3.4, it
was possible to go from 16.4 % of conversion of the 350 °C+ fraction at the lowest temperature,
up to 22.6 % at the highest for the thermal cracking reactions. Product properties were plotted as
a function of conversion as presented on Figure 3.5 to facilitate the visualization of the results.
Reasonable trends were observed for the MCR content, which increased as the temperature
increased, reaching a maximum value of 12.62 % at the highest temperature. As expected,

42
viscosity reduced with the increase of reaction severity. Also, the P-value, which is an indication
of the stability of the asphaltene molecules in the product, showed a decreasing trend with
severity for thermal cracking, with values lower than the accepted limit of 1.15 for the most
severe condition (C3), indicating the limit of the non-catalytic process at these conditions. For
the catalytic steam cracking reactions, it was possible to go from 23.9 % of conversion of the 350
°C+ fraction at the lowest temperature (423 °C), up to 34.7 % at the highest (435 °C). Comparing
condition 4 (SC) with condition 5 (CSC), which are both at the same conditions, an increase of 2
points in conversion was observed, which suggest that the catalyst was not activated at that
temperature since the difference is insignificant. However, at 435 °C a significant increase of
conversion was observed, probably caused by the activation of the catalyst at that level of
severity. Reasonable trends were observed for the MCR content, which increased with
temperature, reaching a maximum value of 12.76 % at the highest temperature (C9-MB1).

In Figure 3.6, the product distribution by cuts as a function of the conversion for the DAO
SC and CSC runs is shown, which is directly related to the severity of the reaction. As can be
observed for the thermal cracking reactions, the yield of the heavier cuts (LVGO, HVGO, and
residue) tend to decrease as the conversion increases, and the yields of the lighter fractions
(naphtha, kerosene, and diesel) slightly increase, which is a typical trend for a conventional
visbreaking process.2 The yield of gases did not change significantly with the conversion.
Similarly to thermal cracking, the yields of the heavier cuts (LVGO, HVGO, and residue) tend to
decrease as the conversion increases for the CSC processing, and the yields of the lighter
fractions (naphtha, kerosene, and diesel) tend to increase. However, in this case, the yields of
LVGO were higher than the yields of residue, which benefits the formation of a total lighter
product, with lower viscosity than thermal cracking products. For the catalytic runs, the gas yield
showed an overall increase with the severity of the reaction.

43
Table 3.4. Summary of experimental information for semi-continuous commissioning runs.
Section A: thermal cracking; Section B: catalytic steam cracking

Thermal Cracking
Section A Feed C4 C4
C1 C2 C3
MB1 MB2
Temperature [°C] N/A 415 420 425 423 423
LHSV [h-1] N/A 2 2 2 2 2
HC Mass Balance [%] N/A 97.1 94.8 95.9 96.0 95.4
Conversion (350 °C+)[%] N/A 16.4 18.8 22.6 21.7 20.4
P-value N/A >1.15 >1.15 <1.15 1.15 1.15
Asphaltenes-C5 [%wt] 6.7 8.3 8.4 11.2 10.6 10.1
API Gravity 11.2 10.7 11.0 12.0 11.7 11.2
MCR [%wt] 9.34 10.28 11.69 12.62 12.54 11.71
Viscosity @25°C [cP] 19297 3176 1864 936 1184 1060

Catalytic Steam Cracking


Section B C9 C9
C5 C6 C7 C8
MB1 MB2
Temperature [°C] 423 427 425 430 435 435
LHSV [h-1] 2 2 2 2.5 2 2
HC Mass Balance [%] 98.1 97.9 99.7 98.2 97.1 98.4
Conversion (350 °C+)[%] 23.9 27.9 25.3 26.4 34.7 33.9
P-value Presence of solids observed
Asphaltenes-C5 [%wt] 10.1 14.8 11.5 11.7 14.5 13.7
API Gravity 10.5 11.3 10.6 11.5 12.7 12.4
MCR [%wt] 12.38 12.35 12.22 12.52 12.76 13.51
Viscosity @25°C [cP] 1181 720 1665 720 238 311

44
Figure 3.5. Product properties as function of conversion for commissioning runs using the semi-
continuous product collection system.

Figure 3.6. Product distribution and microcarbon content as function of conversion for
commissioning runs using the semi-continuous product collection system.

45
When determining stability of the liquid products from the catalytic runs, there were some
challenges due to the presence of solid particles that hindered the visualization of asphaltenes by
optical microscopy since it was not clear whether the observed particles were precipitated
asphaltenes or simply small particles of catalyst or coke. Consequently, a desalting process for
one particular condition was carried out in order to remove solids (including the catalyst) and
water-soluble salts. It was found that after the desalting there were neither solids nor precipitated
asphaltenes at the P-value stability limit of 1.15 as observed in Figure 3.7. The desalting was
carried out by mixing the oil with water and centrifuging it to separate the solids from the liquid
hydrocarbon phase and the water. Salts present in the product were dissolved in the water phase.
This mimics removal of UDC particles in an industrial operation.3 In this way, the increase in
stability when using the Ni/K UDC was confirmed, allowing higher levels of conversion, thus
leading to increased amounts of lighter products compared to thermal processing. For the
catalytic condition C11 (440 °C-2 h-1), the desalting process did not show the same result as for
condition 9, instead, asphaltene agglomerates were observed after the desalting process at the P-
value stability limit of 1.15 as observed in Figure 3.8, which is in agreement with the
observations made during the experimental run which indicate that the reaction was too severe,
leading to coke formation and triggering of asphaltenes precipitation.

Before Desalting After Desalting

Figure 3.7. Microscope image of sample for condition 9 (435 °C-2 h-1) (0.15 mL nC16 added)
during the P-value analysis.

46
Before Desalting After Desalting

Figure 3.8. Microscope image of sample for condition 11 (440 °C-2 h-1) (0.15 mL nC16 added)
during the P-value analysis.

Due to the proximity in density of the whole product and water, desalting turned out to be a
difficult task to be carried out at laboratory scale as done with the previous two desalted samples.
To avoid this, future samples were cleaned of solids following the CHCl3 filtration method
explained in more detail in Section 2.2.2.5.

3.2.2. Continuous Product Collection

As represented in Figure 3.9, where the average temperatures and pressure in the reactor are
plotted against time for the continuous collecting run carried out at high LSHV, pressure drops
for each actuation of the pneumatic drainage valves were very small (less than 3%) and the
operating pressure set point was recovered well in advance before the next actuation of the
valves occurred. The high frequency of the valves at this LHSV did not cause any kind of
instability in the reactor temperatures and the operation of the unit was very stable. In Figure
3.10, where a similar profile is presented for the low LHSV case, even though the pressure drop
in each actuation was slightly higher than in the previous case, it was still below 3%. In addition,
it is important to highlight that, unlike the previous case, in this test each actuation of the valves
that occurred every hour caused a minimum drop in the reactor temperature (less than 2 °C) that
was recovered immediately by the reactor PID control. Furthermore, in this run the back pressure
valve (BPV) was slightly opened during the first mass balance in order to lower the pressure
closer to the set point since it had stabilized at a slightly higher pressure before opening the mass

47
balance period; it is more difficult to control the pressure at low LHSV since the volumetric flow
of gas generated, which the BPV uses for pressure control, is very low compared to the first case.
The operability of the bench-scale pilot plant continuous collecting system at low LHSV was
also smooth and all the operating parameters were successfully controlled. Mass balance closure
for the 6 mass balances collected was 100.2±1.6 %, which is a positive indication of the stability
and reliability of the collecting system.

Figure 3.11 and Figure 3.12 present the SimDist results for the light and heavy product
fractions collected for each one of the three mass balances at both evaluated conditions, as well
as the SimDist of the whole product resulting of the blending of the light and heavy fraction at
the obtained proportions. As can be observed form the plots, all the distributions of hydrocarbons
by boiling point are practically the same for all of the collected samples. This corroborates the
stable operation of the unit from an analytical point of view. Based on the previous results, a
satisfactory performance of the unit can be guaranteed within the evaluated range of LHSV of
0.25-2.0 h-1 at which most industrial upgrading and refining processes take place. This serves as
a proof of the versatility of the unit under different operating conditions.

Figure 3.9. Temperature and pressure profile for continuous product collecting system at 423 °C
and 2 h-1.

48
Figure 3.10. Temperature and pressure profile for continuous product collecting system at 380
°C and 0.25 h-1.

Figure 3.11. Simulated distillation comparison for product fractions gathered at 423 °C and 2 h-1
for continuous product collecting system.

49
Figure 3.12. Simulated distillation comparison for product fractions gathered at 380 °C and 0.25
h-1 for continuous product collecting system.

3.3. References

(1) Fathi, M. M.; Pereira-Almao, P. Catalytic Aquaprocessing of Arab Light Vacuum Residue
via Short Space Times. Energy Fuels 2011, 25 (11), 4867-4877.
(2) Speight, J. G. Visbreaking: A technology of the past and the future. Scientia Iranica 2012, 19
(3), 569-573.
(3) Gray., M. R., Upgrading oilsands bitumen and heavy oil. The University of Alberta Press:
Edmonton, 2015.

50
Chapter 4: Reactivity and Comprehensive Kinetic Modeling of Deasphalted Vacuum
Residue Thermal Cracking

Fredy A. Cabrales-Navarro* and Pedro Pereira-Almao


Department of Chemical and Petroleum Engineering, Schulich School of Engineering,
University of Calgary, 2500 University Drive N.W., Calgary, AB T2N 1N4 , Canada
* Corresponding author (facabral@ucalgary.ca)

4.1. Abstract

Upgrading and Refining of Heavy Oil and Bitumen has become a costly practice in the last
decades, creating a need for new processes to appropriately convert these heavy feedstocks into
lighter and more valuable materials. Solvent Deasphalting (SDA) is a carbon rejection process
where asphaltenes are removed from the oil using a paraffinic solvent, producing a lighter
deasphalted oil (DAO) stream that can be further upgraded without the limitations of asphaltenes
instabilities. In this work, upgrading via thermal cracking of deasphalted vacuum residue was
assessed in a research-scale pilot plant equipped with an up-flow open tubular reactor. Two
different feedstocks were evaluated: converted and virgin DAO. Reactivity experiments were
carried out at temperatures within the interval 380-423 °C and Liquid Hourly Space Velocities
(LHSV) of 0.25-3 h-1. The effects of operating pressure and steam partial pressure on thermal
crackability of DAO were also evaluated not having significant effect within the 150-300 psig
range. Experimental results showed increased reactivity of virgin DAO compared to converted
DAO as well as better quality and higher selectivity to more valuable products. A lumped kinetic
modeling including asphaltenes generation was evaluated. Different scenarios were assessed to
incorporate chemical considerations within the complex lumped kinetic model developed, which
have a significant effect on the kinetic parameters as well as the possible interpretation of the
results. Overall, typical behavior was observed for thermal cracking with global activation
energies of 241 kJ/mol 226 kJ/mol for converted and virgin DAO (560 °C+) respectively. Good
fitting of the model with experimental data with relative errors around 7 % and clear kinetic
relevance of the asphaltenes generation reaction during thermal cracking was highlighted.

Keywords: Virgin DAO, Converted DAO, Asphaltenes, Lump, Thermal Cracking.

51
4.2. Introduction

Asphaltenes, a group of heavy hydrocarbons found in larger proportions in heavy and extra-
heavy oils, defined by their insolubility in n-paraffinic solvents such as pentane or hexane and
solubility in aromatics like toluene are directly related to the increase in viscosity and density of
the oil.1 For this reason, the use of solvent deasphalting to separate the asphaltenic compounds
has become practical in the heavy oil upgrading industry.
The presence of asphaltenes not only limits the application of catalytic processes for further
upgrading of the oil due to catalyst poisoning, but also the achievement of increased conversions
in conventional processes such as thermal cracking (TC) due to their instability when submitted
to moderate severity conditions.2,3 In this way, there is a new window of opportunities for new
technologies for deep conversion of deasphalted oil (DAO) streams into lighter and more
valuable compounds. These potential developments need appropriate kinetic models that would
allow the simulation of said processes on an industrial scale. This valuable information is
abundant for hydrocracking4-6 and conventional thermal cracking7-10 using hydrocarbon fractions
such as vacuum residue (VR), but little information is available in the open literature for DAO
thermal cracking. As an example, analyzing the upgrading scheme shown in Figure 4.1, which is
based on a combination of SDA/TC with recycling of the thermally cracked DAO commercially
known as the OrCrudeTM process11, it is crucial to model the thermal cracking operation and to
accurately estimate the amount of valuable products, such as atmospheric gas oil (AGO) and
vacuum gas oil (VGO). Moreover, the amount of pitch obtained, which is a combination of
pentane insoluble asphaltenes and resins characterized by having a lower economic value and
being difficult to process, would depend not only on the amount of asphaltenes originally present
in the bitumen, but on the amount of asphaltenes generated from the thermal cracking process.
For this reason, it is essential to predict the amount of these heavy compounds generated in order
to have an accurate estimation of the pitch yield in the upgrading scheme.

52
Figure 4.1. OrCrude upgrading scheme.

For the kinetic modelling of DAO upgrading via thermal cracking, a traditional lumping
technique in which the hydrocarbon is divided into lumps of fractions defined by their boiling
point is proposed. Several authors have used this approach for thermal cracking of vacuum
residue using lumping schemes that can range from a simple two lumps model such as the work
of Del Bianco et al.12 to more complex reaction pathways involving 5 or 6 different lumps such
as investigations by Singh et al.13 or Mohaddecy and Sadighi14 respectively. Kataria et al.10 and
Singh et al.9 presented a very detailed chronological summary about kinetic modeling
approaches for visbreaking of residual feedstocks. On the other hand, these types of lumping
kinetic models have also been successfully used for hydrocracking and catalytic steam cracking
(CSC) of vacuum residue. Sanchez, et al.5 followed this approach for moderate hydrocracking of
Maya heavy crude with a Ni/Mo catalyst. Also, it was successfully implemented by Puron et al.4
for hydrocracking of a vacuum residue fraction from Maya heavy oil using the catalytic matrixes
NiMo/Al2O3, and NiMo/Al2O3-Cr. Furthermore, this methodology has been used for kinetic
modeling bitumen hydroprocessing near reservoir conditions using ultradispersed catalysts.15-17 It
was also preliminarily used by Fathi and Pereira-Almao18 for thermal cracking and catalytic
steam cracking of Arab light vacuum residue at high space velocities using ultradispersed
catalysts under very similar conditions to the ones used in this work.
A common flaw of most mentioned models is that in the considered pathways, the formation
of coke precursors (asphaltenes) is not considered. Asphaltenes are known to promote the
formation of insoluble materials such as coke and also to increase the yield of gases. Thus, their

53
incorporation in the model is of relevance considering that they seriously affect the conversion
limit of the cracking process; most models assume working conditions under which those
reaction paths are negligible. For bituminous oils, it is not appropriate to neglect the influence of
asphaltenes on the reaction scheme, since there are considerable amounts of them within the
feedstock. Even for deasphalted oil fractions, where the majority of the asphaltenes has been
removed, their generation during the reaction is significant. As a result, they have a relevant
effect on the kinetic model.
This work targets the development of a multi-lump kinetic model that was progressively
analyzed to reduce to 8-lumps incorporating the formation of asphaltenes during thermal
cracking of converted and virgin DAO obtained from a bitumen upgrading facility located in
Northern Alberta, Canada. The main objective of the work is to gain understanding about the
reactivity of DAO thermal cracking and the kinetics of formation of pentane insoluble
asphaltenes when included as a separate lump in the kinetic model, an approach not reported in
the open literature thus far. In addition, this work aims to create a more robust practical model
differentiating the reactivity of vacuum gas oil by splitting it into heavy vacuum gas oil (HVGO)
and light vacuum gas oil (LVGO), effective streams in most industrial vacuum distillation units.
Using a similar reasoning, the distillates lump was split into diesel and kerosene instead of the
simplified lumping as distillates. Analyzing the results of our modeling it was found that a
reduced number of kinetic constants were really needed to express the kinetic process. For
instance the lighter lumps diesel, kerosene and naphtha do not react to lighter products at the
evaluated conditions. In addition, the effect of varying total operating pressure and steam partial
pressure on the DAO reactivity was assessed.

4.3. Experimental Methods

4.3.1. Experimental Setup

The reactivity tests were carried out in an experimental setup equipped with an up-flow open
tubular reactor that resembles a conventional thermal cracking unit. A general scheme of the
reactivity test unit is shown in Figure 4.2.

54
Figure 4.2. Schematic of the experimental setup for reactivity tests.

The reactivity test unit is composed of five sections: the feed section, reaction section, hot
separation section, cold separation section, and finally the gas release, depressurization and
analysis section. The feed section is equipped with a Teledyne ISCO series 500D dual-pump
continuous flow system with dual pneumatic air-valves controlled by a Series D Controller,
which refills one pump while the other one is delivering DAO. Also, there is a Shimadzu LC-
20AD high performance liquid chromatography pump for water injection. Feed streams were
mixed inline before entering the reaction zone, which comprises an up-flow open tubular reactor
with a volume of 103.46 mL, operated at constant temperature and pressure, and equipped with
an Omega custom 7 sensing points thermocouple profile probe. The products coming out from
the reaction zone go to a hot separator in which the water, gases, and a fraction of the lightest
hydrocarbons are separated from the product by keeping the separator at a constant temperature
(mostly 250 °C). The separated gaseous stream is then passed through a condenser in order to
liquefy the light hydrocarbons and water, and the whole stream is sent to the cold separation
section where gases are separated from the liquid product. Finally, gases coming out from the
cold separation zone passed through a back pressure valve, which maintains the operating

55
pressure at the given set point, and the gas stream is sent to gas release and analysis. In this zone,
the exiting gas stream can be directed to mass flow measurements, gas chromatography (GC)
analysis or directly to the gas exhaust. Gas stream is first sweetened by bubbling in an aqueous
potassium hydroxide (KOH) trap and then it is either vented or sent to gas flow measurement in
a Shinagawa W-NK-0.5 wet gas meter (WTM). The other mode of operation is to flow the
stream through the GC for gas composition analysis, followed by gas sweetening in a similar
KOH trap and finally the stream is vented.

4.3.2. Experimental Plan

Two industrial deasphalted vacuum residues obtained from a bitumen upgrading facility
located in Northern Alberta, Canada were characterized. The first material (converted DAO)
comes from solvent deasphalting of a mixture of unconverted vacuum residue from a thermally
cracked DAO plus virgin vacuum residue from bitumen as shown in the upgrading scheme in
Figure 4.1. The second material (virgin DAO) was obtained from solvent deasphalting of
completely virgin vacuum residue as in the scheme in Figure 4.1 without the thermal cracking
recycling loop. Characteristics of both feedstocks are presented in Table 4.1.

Table 4.1. Feedstock characterization

Converted Virgin
Analysis
DAO DAO
API Gravity 6.0 6.3
Microcarbon Residue [%wt.] 12.95 12.96
Viscosity @60°C [cP] 5339 N/A
Viscosity @100°C [cP] 300 1328
Saturates [%wt.] 4.0 2.4
Aromatics [%wt.] 75.0 76.0
SARA
Resins [%wt.] 20.2 20.8
Asphaltenes-C5 [%wt.] 0.7 0.8
Naphtha (28 - 190 °C) [%wt.] 0.18 0.00
Kerosene (190 - 260 °C) [%wt.] 2.70 0.00
Cut Diesel (260 - 343 °C) [%wt.] 5.25 0.00
Yields LVGO (343 - 453 °C) [%wt.] 7.00 0.56
HVGO (453 - 560 °C) [%wt.] 17.15 12.63
Vacuum Residue (560°C+) [%wt.] 67.72 86.81

56
Thermal cracking in the presence of 5 % wt. steam and operating pressure within the 260 to
400 psig range has been carried out in the Catalysis for Bitumen Upgrading Research Facility at
the University of Calgary for different feedstocks such as Arab light Vacuum Residue19 and
Athabasca vacuum gas oil20 with satisfactory results that allow to apply a similar methodology
for a residual deasphalted oil feedstock. A total of 12 conditions were evaluated for converted
DAO thermal cracking experiments required for the kinetic modeling, consisting of three
different LHSVs at four different temperatures: 380, 409, 416 and 423 ºC. The LHSV set at each
temperature were chosen in such a way that the most severe conditions were as close as possible
to the stability limit of the liquid products in terms of P-value stability index (1.10-1.20).21,22
LHSVs spanned a common range for thermal cracking of 1-3 h-1, except for the low temperature
evaluation carried out at 380 ºC where lower space velocities between 0.25-0.75 h-1 were used in
order to have conversions within similar levels. Due to the limited availability of virgin DAO,
only 6 conditions were evaluated, selecting two LHSVs at three different temperatures: 409, 416
and 423 ºC. All reactions were carried out at a pressure of 300 psig and 5 %wt. steam. Due to the
need to heat up the DAO feedstock to be able to mobilize it, LHSV was defined at pumping
conditions for all experiments as expressed in Eq. 4.1.

𝑉𝑜𝑙𝑢𝑚𝑒𝑡𝑟𝑖𝑐 𝑓𝑙𝑜𝑤 𝑜𝑓 𝐷𝐴𝑂 𝑎𝑡 140 °𝐶 𝑎𝑛𝑑 300 𝑝𝑠𝑖𝑔 [𝑚𝑙/ℎ]


LHSV [ℎ−1 ] = Eq. 4.1
𝑅𝑒𝑎𝑐𝑡𝑜𝑟 𝑉𝑜𝑙𝑢𝑚𝑒 [𝑚𝑙]

Additionally, looking forward to evaluating the influence of pressure on the converted DAO
thermal cracking, a set of 6 experiments was carried out varying the total operating pressure and
the steam partial pressure. The first 3 experiments were run at a constant pressure of 300 psig
and changing the steam content from 2.5, 5.0 and 7.5 %, and compensating with nitrogen in
order to keep similar gas levels in the reactor as those produced under the maximum steam
content of 7.5 % wt., assuming that hydrocarbon gases generated by changing steam/nitrogen
content are not significant. In this way, the liquid hold-up is the same for the three conditions and
any difference on product quality or reactivity can be attributed to the effect of the partial
pressure of steam in the reactor. The last three experiments were performed using a total pressure
of 150, 225, and 300 psig with constant 5% wt. steam.

57
In order to compare the reactivity at each condition, the conversion of the hydrocarbons that
boil at temperatures above 560 ºC (HC 560°C+) was used as defined in Eq. 4.2:

% 𝑤𝑡 𝐻𝐶 (560 °C +) 𝐹𝑒𝑒𝑑 − % 𝑤𝑡 𝐻𝐶 (560 °C +) 𝑃𝑟𝑜𝑑𝑢𝑐𝑡


Conversion = ∗ 100% Eq. 4.2
%𝑤𝑡 𝐻𝐶 (560 °C +) 𝐹𝑒𝑒𝑑

4.3.3. Feedstock and Product Characterization

In order to assess and compare the performance of the reactivity tests in terms of quality of
the products and to provide the information required for kinetic modeling, liquid and gas product
samples were collected and analyzed. Simulated distillation is performed to determine weight
percentage of each lump following the ASTM D7169-05 method in an Agilent 6890N
chromatograph equipped with automatic injection as described in detail by Carbognani et al.23 In
order to determine the state of peptization of the asphaltenes in the liquid product samples, the P-
value method described by Di Carlo and Janis22 was used. This method consists of the
determination of the maximum dilution of the sample with hexadecane at which asphaltenes are
not precipitated when observed in an optical microscope with a magnification of 40X. The P-
value is calculated following Eq. 4.3.

𝐻𝑒𝑥𝑎𝑑𝑒𝑐𝑎𝑛𝑒 [𝑚𝑙]
P − value = 1 + Eq. 4.3
1 𝑔 𝑠𝑎𝑚𝑝𝑙𝑒

Microcarbon Residue (MCR) was determined following the muffle furnace method
developed by Hassan et al.24 This analysis provides an indirect measurement of the capability of
the samples to produce coke and also gives useful information about the extent of advance of the
undesired condensation reactions taking place during thermal cracking of residual hydrocarbons.
The API gravity of the liquid products was calculated as defined in the ASTM D287-92 standard
method, for which it is necessary to determine the specific gravity of the sample. For this, a
digital densitometer Rudolph Research Analytical model DDM2911 was used as described with
more detail by Carbognani et al.25 Two dilutions in toluene at concentrations close to 0.5 % v/v
and 1 %v/v are prepared and their densities are measured in the equipment. Then, the density of
the pure sample is extrapolated by assuming an ideal mixture for diluted solutions of components
with similar polarity and assuming additive volumes following Eq. 4.4.

58
𝑚𝑠𝑎𝑚𝑝𝑙𝑒
𝜌𝑠𝑎𝑚𝑝𝑙𝑒 = Eq. 4.4
(1⁄𝜌𝑡𝑜𝑙𝑢𝑒𝑛𝑒 )(𝑚𝑠𝑎𝑚𝑝𝑙𝑒 + 𝑚𝑡𝑜𝑙𝑢𝑒𝑛𝑒 ) − (1/𝜌𝑡𝑜𝑙𝑢𝑒𝑛𝑒 )𝑚𝑡𝑜𝑙𝑢𝑒𝑛𝑒

Where 𝑚𝑠𝑎𝑚𝑝𝑙𝑒 is the mass of the sample, 𝜌𝑠𝑎𝑚𝑝𝑙𝑒 is the density of the sample, 𝜌𝑡𝑜𝑙𝑢𝑒𝑛𝑒 is the
density of toluene used for dilution and 𝑚𝑡𝑜𝑙𝑢𝑒𝑛𝑒 is the mass of toluene added. The viscosity was
measured in a Brookfield viscometer model DV-II+ coupled with a water recirculation system
model TC-502 that allows measurements between 0 and 100 °C. Hydrocarbon group-type
separation analysis of Saturates, Aromatics, Resins and Asphaltenes (SARA) was also carried
out for all liquid samples, not only to quantify the amount of asphaltenes generated by thermal
cracking required for kinetic modeling, but to acquire important information about the reactivity
of different hydrocarbon groups when submitted to thermal cracking. This analysis was
performed following the procedure developed by Carbognani et al.26 Gas samples were analyzed
inline using a SRI Instruments chromatograph model 8610C equipped with 4 packed columns,
two switching valves, 2 thermal conductivity detectors (TCD), one operating with argon as
carrier gas to detect hydrogen at low concentrations and the other with helium as carrier gas to
determine hydrogen at high concentrations and the remaining permanent gases and light
hydrocarbons. A Flame Photometric Detector (FPD) provided sulfur compounds analysis
(mainly H2S) and a Flame Ionization Detector (FID) was used for analysis of light hydrocarbons.

4.3.4. Kinetic Modeling

For modeling of the experimental data, a plug-flow reactor (PFR) model was assumed. In
this case, an isothermal reaction with negligible radial variations at steady state was considered.
The mathematical expression to represent the PFR model is given by:

𝑑𝑦𝑖 𝑑𝑦𝑖
= = 𝜎𝑖 Eq. 4.5
𝑑𝜏 𝑑(1/𝐿𝐻𝑆𝑉))

Where 𝑦𝑖 is the weight fraction of lump i, 𝜏 is the space time, and 𝜎𝑖 is the reaction rate for
lump i.

A first order rate law has shown good results not only for modeling hydrocracking reactions
using the lumping technique, but more importantly for catalytic steam cracking and thermal
cracking reactions under similar conditions as those proposed in this work.18 In this way, the first

59
order rate law expressed in Eq. 4.6, where 𝑘𝑖 is the kinetic constant for reaction i, was adopted
for each reaction.

𝜎𝑖 = 𝑘𝑖 ∗ 𝑦𝑖 Eq. 4.6

Figure 4.3. Proposed kinetic model for DAO thermal cracking.

Combining Eq. 4.5 and Eq. 4.6 and expressing the global reaction rate for each lump in
terms of the specific reaction rates following the kinetic scheme shown in Figure 4.3, 8 different
mass balance equations conforming a system of ordinary differential equations (ODE) is
obtained and is represented by Eq. 4.7 to Eq. 4.14. It was assumed that production of asphaltenes
(Asp-C5) comes directly from the hydrocarbons that boil above 560 °C that are not asphaltenes,
namely DAO (560 °C+). The recombination of these two lumps represents the vacuum residue or
VR (560 °C+).

𝑑𝑦𝐴𝑠𝑝−𝐶5
= 𝑘1 𝑦𝐷𝐴𝑂 (560 °C+) Eq. 4.7
𝑑(1/𝐿𝐻𝑆𝑉)

𝑑𝑦𝐷𝐴𝑂 (560 °𝐶+)


= − (𝑘1 + 𝑘2 + 𝑘3 + 𝑘4 + 𝑘5 + 𝑘6 + 𝑘7 )𝑦𝐷𝐴𝑂 (560 °𝐶+) Eq. 4.8
𝑑(1/𝐿𝐻𝑆𝑉)

60
𝑑𝑦𝐻𝑉𝐺𝑂
= 𝑘2 𝑦𝐷𝐴𝑂 (560 °C+) − (𝑘8 + 𝑘9 + 𝑘10 + 𝑘11 + 𝑘12 )𝑦𝐻𝑉𝐺𝑂 Eq. 4.9
𝑑(1/𝐿𝐻𝑆𝑉)

𝑑𝑦𝐿𝑉𝐺𝑂 Eq. 4.10


= 𝑘3 𝑦𝐷𝐴𝑂 (560 °C+) + 𝑘8 𝑦𝐻𝑉𝐺𝑂 − (𝑘13 + 𝑘14 + 𝑘15 + 𝑘16 )𝑦𝐿𝑉𝐺𝑂
𝑑(1/𝐿𝐻𝑆𝑉)

𝑑𝑦𝐷𝑖𝑒𝑠𝑒𝑙
= 𝑘4 𝑦𝐷𝐴𝑂 (560 °C+) + 𝑘9 𝑦𝐻𝑉𝐺𝑂 + 𝑘13 𝑦𝐿𝑉𝐺𝑂 Eq. 4.11
𝑑(1/𝐿𝐻𝑆𝑉)

𝑑𝑦𝐾𝑒𝑟𝑜𝑠𝑒𝑛𝑒
= 𝑘5 𝑦𝐷𝐴𝑂 (560 °C+) + 𝑘10 𝑦𝐻𝑉𝐺𝑂 + 𝑘14 𝑦𝐿𝑉𝐺𝑂 Eq. 4.12
𝑑(1/𝐿𝐻𝑆𝑉)

𝑑𝑦𝑁𝑎𝑝ℎ𝑡ℎ𝑎
= 𝑘6 𝑦𝐷𝐴𝑂 (560 °C+) + +𝑘11 𝑦𝐻𝑉𝐺𝑂 + 𝑘15 𝑦𝐿𝑉𝐺𝑂 Eq. 4.13
𝑑(1/𝐿𝐻𝑆𝑉)

𝑑𝑦𝐺𝑎𝑠𝑒𝑠
= 𝑘7 𝑦𝐷𝐴𝑂 (560 °C+) + 𝑘12 𝑦𝐻𝑉𝐺𝑂 + 𝑘16 𝑦𝐿𝑉𝐺𝑂 Eq. 4.14
𝑑(1/𝐿𝐻𝑆𝑉)

This system of ODEs was numerically integrated within the experimental range of space
times using the Runge-Kutta method implemented in a Matlab code. The modeling objective is
to find the set of unknown kinetic constants at which the modeled product distribution results as
close as possible to the experimental values. An objective function based on the sum of squared
errors (SSE) between experimental and modeled compositions defined by Eq. 4.15 was used to
quantify the fitting between experimental and modeled data, as follows:

𝑛𝑇 𝑛𝑝
2
𝑆𝑆𝐸 = ∑ ∑(𝑦𝑖𝑒𝑥𝑝 −𝑦𝑖𝑚𝑜𝑑 ) Eq. 4.15
𝑗=1 𝑖=1

𝑛𝑇 is the number of evaluated temperatures and 𝑛𝑝 represents the number of experimental


points evaluated at each temperature.

As suggested by Fathi and Pereira-Almao18, three criteria were considered to find the set of
kinetic constants that represent the experimental behavior of the process. The first one is the

61
minimization of the objective function, for which an optimization program was implemented in
the Matlab code using the nonlinear optimization algorithm fmincon. This allows for the
incorporation of linear and inequality constraints during the minimization. The second criterion
is that the global experimental reaction rate of transforming the DAO (560 ºC+) to the other 7
lumps has to be as close as possible to the sum of the specific reaction rates of DAO (560 ºC+) to
each lump (𝑘𝑔𝑙𝑜𝑏𝑎𝑙 𝐷𝐴𝑂 (560 °C+) = 𝑘1 + 𝑘2 + 𝑘3 + 𝑘4 + 𝑘5 + 𝑘6 + 𝑘7 ), with 𝑘𝑔𝑙𝑜𝑏𝑎𝑙 𝐷𝐴𝑂 (560 °C+)
calculated from the experimental results by estimating the slope of the plot (1/𝐿𝐻𝑆𝑉)𝑖
vs 𝑙𝑛(𝑦0 ⁄𝑦𝑡 )𝐷𝐴𝑂 (560 °C+) , where 𝑦0 is the concentration of DAO (560 °C+) in the feedstock and
𝑦𝑡 is the concentration of DAO (560 °C+) in the whole liquid products at each evaluated space
time. Finally, the last criterion considered was that the reaction rate constants should follow the
Arrhenius law expressed by Eq. 4.16:

𝐸𝑎𝑖
ln(𝑘𝑖 ) = ln(𝐴𝑖 ) − Eq. 4.16
𝑅𝑇

As indicated by Blasetti and de Lasa27, the kinetic parameters obtained from the Arrhenius
equation (𝐸𝑎𝑖 and 𝐴𝑖 ) can show a mutual adverse effect of one parameter estimate, which is
known as parameter correlation. Agarwal and Brisk28 used a re-parameterization equation that
has proved to reduce the correlation between 𝐸𝑎𝑖 and 𝐴𝑖 from the Arrhenius expression, which
consists of centering the temperature values around a reference temperature 𝑇 ∗ as shown in Eq.
4.17.

𝐸𝑎𝑖 1 1
𝑘𝑖 = 𝐴∗𝑖 𝑒 [− ( − )] Eq. 4.17
𝑅 𝑇 𝑇∗

Where,

𝐸𝑎𝑖
[− ]
𝐴∗𝑖 = 𝐴𝑖 𝑒 𝑅𝑇 ∗ Eq. 4.18

In order to determine whether there was any significant correlation effect


between 𝐸𝑎𝑖 and 𝐴𝑖 , kinetic parameters were estimated with and without parametrization of the
Arrhenius equation for a base case. A reference temperature of 407 °C (average temperature for
the evaluated range) was used for the re-parameterized equation (Eq. 4.17).

62
When estimating kinetic parameters, the values of the reaction rate constants can be
obtained at each temperature and then Eq. 4.16 is used to estimate the activation energy (𝐸𝑎𝑖 )
and the pre-exponential factor (𝐴𝑖 ) by using linear regression, or the minimization algorithm can
be expressed as a function of the activation energy and the pre-exponential factor for each
reaction and the whole set of experimental points is optimized in one step. The advantage of the
first case is that the number of parameters to be optimized at each temperature is lower, but there
is the disadvantage of having higher chances of fitting errors in the linear regression of the
Arrhenius equation that would represent higher global error for the kinetic model due to its
exponential nature. On the other hand, in the second scenario it is guaranteed that the kinetic
parameters follow the Arrhenius equation, but the optimization is much more difficult due to the
higher number of parameters and also the high non-linearities introduced in the system of ODEs
with the exponential form of the Arrhenius expression, causing the ODE solver not to converge
to a solution, which can be overcome by having an analytical solution for the system of ODEs
usually very complex for such a high number of reaction paths. To avoid this, a different
approach was considered in this work, with similarities to the one proposed by Da Silva.16 The
Arrhenius equation was not directly replaced in the system of ODEs, but its dependence was
included by means of modification in the objective function that accounts for the error in the
linear regression of Eq. 4.16 as represented in Eq. 4.19, where 𝑛𝑟 is the number of reactions in
the kinetic scheme, and 𝑟𝑘 is the coefficient of determination for the linear fitting of the
Arrhenius equation of each reaction. In this way, the whole set of data was optimized in a single
step. A similar approach was also implemented by Taghipour and Naderifar8 for kinetic
modeling of vacuum residue thermal cracking.

𝑛𝑇 𝑛𝑝 𝑛𝑟 2
2
𝑆𝑆𝐸 = ∑ ∑(𝑦𝑖𝑒𝑥𝑝 −𝑦𝑖𝑚𝑜𝑑 ) + (𝑛𝑟 − ∑ 𝑟𝑘 ) Eq. 4.19
𝑗=1 𝑖=1 𝑘=1

The Average Absolute Error (AAE) between the experimental and model results for each
lump was also calculated in the Matlab code following Eq. 4.20, where k represents each lump (1
to 8).

𝑒𝑥𝑝 𝑚𝑜𝑑
. 𝑛𝑇 𝑛𝑝 𝑦𝑘,𝑖 −𝑦𝑘,𝑖
∑𝑗=1 ∑𝑖=1 | 𝑒𝑥𝑝 | Eq. 4.20
𝑦𝑘,𝑖
𝐴𝐴𝐸𝑘 [%] = ∗ 100
𝑛𝑇 + 𝑛𝑝

63
These types of kinetic modeling have the characteristic of having multiple solutions.
Therefore, different patterns can be obtained for the kinetic constants, activation energies and
pre-exponential factors depending on the initial seed for the multivariable optimization, the
algorithm used, the number of reactions involved, etc. For this reason, it is of paramount
importance to incorporate chemical considerations into the mathematical modeling in such a way
that the obtained results are in agreement with what is expected for a thermal cracking reaction
of a residual hydrocarbon. To address these aspects, three scenarios were evaluated following the
analysis of lumped kinetic model for hydrocracking of heavy residue developed by Asaee et al.29

 Case 1 (No sequence): kinetic parameters were obtained without any chemical constraint
using the mathematical modeling described in this section.
 Case 2 (Sequence 1): activation energies of thermal cracking of heavier lumps are lower
than lighter ones since larger molecules require less energy to be cracked. Furthermore, a
sequence was imposed on the activation energies in such a way that the activation energy
of conversion of a lump to lighter products is lower than to heavier products. In addition,
constraints on the kinetic constant of gas production were imposed in such a way that gas
production proceeds more readily from heavier lumps (𝑘7 > 𝑘12 > 𝑘16 ).
 Case 3 (Sequence 2): activation energies of thermal cracking of heavier lumps are lower
than lighter ones. Furthermore, a sequence was imposed on the activation energies in
such a way that the activation energy of conversion of a lump to lighter products is higher
than to heavier products. Similar constraints as in case 2 were placed on the kinetic
constants for gas production.

Looking forward to evaluating if there was any difference in the calculated kinetic
parameters when comparing the proposed 8-lumps model presented in Figure 4.3, a simplified 6-
lumps model as shown in Figure 4.4 was used. The same mathematical approach described in
this section was followed. The number of kinetic constants to be determined at each temperature
using the simplified 6-lumps model is 8 instead of the 16 required to represent the reactions in
the 8-lumps model. The comparison was conducted only for converted DAO thermal cracking
without using any constraint in the optimization algorithm (case 1- no sequence).

64
Figure 4.4. 6-lumps kinetic model for DAO thermal cracking.

Finally, in order to quantify the experimental error of the measured weight fractions for each
one of the kinetic modeling lumps due variability for different mass balances in the pilot plant
run as well as the variability in the characterization techniques (SimDist and nC5-asphaltenes by
microdeasphalting), two different experiments from the pilot plant commissioning reported in
Chapter 3 were analyzed. In each one of those experimental runs conducted at different severity
levels, three mass balances were collected and analyzed. An average of the relative variance was
calculated for each lump.

4.4. Results and Discussion

4.4.1. Effect of Partial and Total Steam Pressure

As observed in the experimental data presented in Table 4.2, no significant effect on the
DAO reactivity was evidenced when changing the steam partial pressure keeping a constant
liquid hold up (Pw1, Pw2 and Pw3) in the reactor with values around 40%, which was somehow
expected since the effect of steam pressure would be predominant on the enhancement of the
product quality by reducing the effect of condensation reactions and coke formation, but not on
the reactivity. However, for the levels of steam evaluated in this work no significant
improvement on the product quality was observed when increasing the content of steam in the
reaction media, which suggests that the lower value of 2.5 %wt. already provides a reasonable
amount of water for the reaction. On the other hand, when keeping a constant flow of water

65
changing the total operating pressure (Pt1, Pt2 and Pt3), it was possible to observe a slight
increase of conversion at the maximum pressure of 300 psig (Pt3), possibly caused by the
increase on the liquid hold up and consequently on the effective liquid hydrocarbon residence
time in the reactor, giving it more time to react compared to the lower pressure cases (Pt1 and
Pt2). In this case (Pt3), a slight increase was also observed on the indicators of the product
quality, such as slight reduction on the microcarbon and asphaltenes content and increase on the
P-value. However, these enhancements are within the ranges of error of the given
characterization techniques, thus it is not possible to affirm that there is a significant
improvement. For this reason, it was not attempted to incorporate pressure as a variable for the
kinetic modeling of the process.

Table 4.2. Effect of pressure on converted DAO thermal cracking. Results reported for the whole
liquid product recovered after reaction

Analysis Condition
Pw 1 Pw 2 Pw 3 Pt1 Pt2 Pt3
Temperature [°C] 423
LHSV [h-1] 2
Pressure [psig] 300 300 300 150 225 300
Water Content [%wt.] 7.5 5.0 2.5 5.0 5.0 5.0
N2 Injected @rxn Conditions 0 12.8 25.6 0 0 0
HC Mass Balance [%]
[mL/min] 98.6 98.0 97.5 97.1 98.0 98.
Conversion HC (560 °C+) [%] 40.0 41.0 39.4 40.9 38.2 43.
9
P-value 1.15 1.15 1.15 1.10 1.10 1.1
3
Saturates [%wt] 13.3 11.6 10.0 12.7 12.4 11.
5
SARA Aromatics [%wt] 67.3 69.0 67.3 64.9 65.6 62.
9
Resins [%wt] 6.8 7.2 8.7 7.8 7.1 11.
6
Asphaltenes-C5 [%wt] 12.7 12.3 14.0 14.6 14.9 14.
4
API Gravity 8.9 8.6 7.8 8.6 8.4 8.1
1
MCR [%wt] 16.57 16.24 16.8 16.82 16.62 16.
Viscosity @25°C [cP] 2080 2863 2899 3011 2802 280
54
2
4.4.2. Reactivity Analysis

In Table 4.3, a brief summary of the experimental conditions, including mass balances and
conversion levels is presented. A range of around 26% of conversion points was explored for the
converted DAO runs spanning from a 17.1% of conversion of the 560 °C+ hydrocarbon fraction
up to 43.3 %. For the virgin material, a narrower range between 32.7 to 48.2 % was investigated

66
due to higher reactivity and feedstock available. Additionally, a very good closure of
hydrocarbon mass balances was obtained for the experimental set with an average value of 99.1
% and standard deviation of 1.1 %, which proves the stable operation and reliability of the
experimental setup used for this work and strengthen the robustness of the experimental data
used for mathematical modeling. Comparing experiment No. 4, No. 8 and No. 12 for the
converted DAO feedstock, which are all at the same space velocity of 2 h -1, an increase of 7.5
conversion points was observed by increasing 1.7% the reaction temperature from 409 °C to 416
°C, and an increase of 13.8 when the temperature increased by 3.4% from 409 °C to 423 °C.
Conversely, comparing for example experiment No. 4 and No. 6 both at the same temperature of
409 °C for the same feedstock, an increase of only 9.5 conversion points is observed when
decreasing the space velocity and thus the liquid residence time in the reactor by 25% from 2 h -1
to 1.5 h-1. In this way, it is possible to observe how sensible the DAO reactivity is to changes in
the reaction temperature compared to the other main variable available to control the reaction
severity, i.e., the space velocity. A similar behavior was observed for the virgin feedstock.
However, when comparing the reactivity of both feedstocks at similar operation conditions we
can see that the 560 °C+ fraction of the virgin material is much more reactive that that of the
converted one, leading to an increase of around 5 conversion points that increase the amount of
valuable light products obtained from this heavy fraction. The loss of reactivity for the converted
material is attributed to the consecutive cracking of the non-asphaltene containing hydrocarbon
fraction in the upgrading scheme illustrated in Figure 4.1.

67
Table 4.3. Summary of experimental data for thermal cracking runs

Converted DAO Virgin DAO


T LHSV
No. -1 HC Mass Conversion HC Mass Conversion
[°C] [h ]
Balance [%] [%] Balance [%] [%]
1 0.75 98.9 17.1
2 380 0.50 100.0 21.0 N/A
3 0.25 99.2 29.2
4 2.0 100.6 29.6 98.7 32.7
5 409 1.5 99.7 31.8 N/A
6 1.0 96.5 39.1 99.0 45.2
7 2.5 97.9 32.7 99.7 36.2
8 416 2.0 98.6 37.1 N/A
9 1.5 99.5 39.9 98.8 45.6
10 3.0 97.4 36.7 101.3 40.7
11 423 2.5 99.5 39.5 N/A
12 2.0 98.9 43.3 99.7 48.2

Reasonable trends were observed for the products distributions as a function of conversion,
which are directly related to the severity of the reaction, as can be seen in Figure 4.5. The
amount of heavy cut DAO (560 ºC+) tend to sharply decrease as the conversion increases, while
other less heavy cuts in the gasoil range (HVGO and LVGO) remain practically unchanged,
possibly because their rate of consumption is similar to their rate of generation under the thermal
reactions taking place. On the other hand, light cuts such as naphtha, kerosene and diesel
moderately increase. This is a typical behavior of a visbreaking process.2,30 Now considering the
undesirable products, gas yields between 3.8 to 6.2 % wt. and 2.8 to 4.7 % wt. were obtained for
the range of conversions evaluated, while the pentane insoluble asphaltenes in the thermal
cracking products were between 7.1 and 13.3 % wt and 4.6 to 10.0 % wt for converted and virgin
DAO respectively, for the minimum and the maximum conversions. Analyzing now the stability
of the liquid products represented by the P-value index shown in Figure 4.5 as a function of
conversion, it is possible to see a considerable deterioration of the stability of the asphaltenes
molecules produced during thermal processing, reaching the commonly accepted instability

68
value of 1.15-1.20 that suggests asphaltenes precipitation that may lead to equipment fouling in
industrial operations at the conversion level of 43 % for both feedstocks. Operation beyond this
conversion level, besides promoting instabilities in the liquid product, would trigger the
production of insoluble materials, mainly coke, and would also increase the yield of gas,
consequently reducing the effective liquid yield of more valuable products in the overall process,
which would not be economically convenient.

As illustrated in Figure 4.6, the ln(viscosity) measured at 25 ºC follows a linear trend of the
conversion with a correlation factor of 0.9922 for the converted DAO and 0.9955 for the virgin
one. As expected for a visbreaking process, the viscosity of the liquid product is reduced by
several orders of magnitude at the maximum conversions for both cases. It is important to
highlight the parallelism on the viscosity reduction trends for both feedstocks, which suggests
that the extent of viscosity reduction is mainly controlled by the degree of conversion achieved
and is not appreciably influenced by the difference in chemical nature of both materials in terms
of the higher amount of partially cracked molecules present in the converted DAO compared to
the virgin one. However, the viscosity values for virgin DAO and its products are higher than for
converted DAO, which can also be explained by the higher branching of the molecules in the
virgin and its products from thermal cracking, than for the converted DAO and its thermal
cracking products, which are less branched and more aromatic.

Conversely, API gravity slightly increased 2 points and 4 points at the maximum
conversions achieved for both converted and virgin DAO respectively, which suggests that there
is in fact a density increase attainable at similar operational conditions due to the recycling effect
in the upgrading loop. A similar behavior was observed for the amount of microcarbon residue
obtained for both cases, as discussed in the ensuing text.

69
Figure 4.5. Product distributions and P-value vs. residue (560 °C+) conversion for converted and
virgin DAO thermal cracking.

70
Figure 4.6. Viscosity and API gravity vs residue (560 °C+) conversion.

As observed in Figure 4.7, converted DAO has a higher tendency to form insoluble
materials (MCR) when compared to the virgin one at similar conversion due to the fact that even
though cracked molecules can be within the same boiling point range as virgin ones, their
chemical composition is different with the possible presence of high aromaticity compounds that
are more prone to condensation reactions leading to coke formation. Considering now the
stability of the asphaltenes present on the thermally cracked products, it was expected to obtain a
more stable product with a higher P-value at the same conversion levels for the virgin material,
but on the experimental results presented in Figure 4.7 it was observed that the P-value trend was
conversion driven, with both lines practically overlapping. One possible explanation for this
phenomenon is that the deterioration of the asphaltenes stability due to thermal cracking leading
to their precipitation at higher conversions is mainly controlled by the nature of the first
asphaltenes that are generated from thermal cracking of the virgin material and once they are
removed by solvent deasphalting and the DAO is submitted again to thermal cracking, the
asphaltenes generated in those subsequent steps systematically decrease their stability,
presumably because their bulk nature is similar as a function of conversion severity.

71
Figure 4.7. MCR residue and P-value profiles vs. residue (560 °C+) conversion.

4.4.3. Kinetic Study

A good fit between the experimental and model data was obtained for converted DAO and
virgin DAO thermal cracking kinetic model without any sequence or constraint on the activation
energies. The linear regression of the plot of compositions obtained by the model vs. the
experimental compositions has a coefficient of correlation close to 1 for both converted and
virgin feedstock as shown in Figure 4.8 and Figure 4.9 respectively. If the experimental and
model compositions are plotted as a function of space time as illustrated in Figure 4.10, a good
fit at the different space velocities evaluated was observed. A similar behavior was obtained for
the other two evaluated scenarios (case 2 and 3) for both feedstocks. Taking a look at the
modeling error for each lump illustrated in Figure 4.11, no significant variation was obtained on
the global average absolute error for case1 (converted DAO 6.72 % and virgin DAO 7.01 %),
case 2 (converted DAO 7.00 % and virgin DAO 6.62 %) and case 3 (converted DAO 7.02 % and
virgin DAO 7.03 %). Thus, the implementation of chemical considerations within the
mathematical modeling did not have an effect on the accuracy and the capability of the model to
represent the experimental data and the modeling error cannot be used as a decisive factor to

72
choose which scenario best represents the chemistry involved in the thermal cracking of these
feedstocks. It was observed that overall the lighter the lump the higher the error, except for the
residual hydrocarbons with boiling points above 560ºC+ that in this case are the DAO (560ºC+)
and the Asp-C5 lumps. This is possibly related to the accuracy of the simulated distillation
method used for determining boiling point distribution of feed and products that is known to
have a higher error for the residual hydrocarbons compared to other lumps.23 Additionally, the
determination of pentane insoluble asphaltenes has a high relative error as in ASTM D893 and
D2007 methods.

0.6
Converted DAO

0.5
Model Composition [weight fraction]

y = 1.0573x - 0.0072 DAO (560 °C+)


R² = 0.9955
0.4 Asp-C5

HVGO
0.3
LVGO

Diesel
0.2
Kerosene

0.1 Naphtha

Gas
0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6
Experimental Composition [weight fraction]
Figure 4.8. Predicted model composition vs. experimental composition for case 1(unconstrained)
converted DAO kinetic model.

73
0.6
Virgin DAO

0.5
y = 1.0397x - 0.0049
Model Composition [weight fraction]

R² = 0.9976
DAO (560 °C+)
0.4
Asp-C5

0.3 HVGO
LVGO

0.2 Diesel
Kerosene
0.1 Naphtha
Gas
0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6
Experimental Composition [weight fraction]
Figure 4.9. Predicted model composition vs. experimental composition for case 1(unconstrained)
virgin DAO kinetic model.

Figure 4.10. Experimental vs. model compositions as a function of space time at 423 °C for case
1 (unconstrained) converted DAO kinetic model.

74
Converted DAO Virgin DAO
20
18
16
14
AAE [%]

12
10
8
6
4
2
0

No sequence Sequence 1 Sequence 2

Figure 4.11. Comparison between the modeling error by lump for converted and virgin DAO
kinetic modeling.
It is important to highlight that, besides the modeling error, there is an experimental error
associated to each pilot plant experiment. Table 4.4 presents an analysis of the variability of the
lump composition determination based on two experimental runs at different operating
conditions performed during the pilot plant commissioning. Three mass balances were analyzed
for each run and the average composition, standard deviation, and relative variance were
reported.

Table 4.4. Statistical analysis of experimental variation for each kinetic model lump
Condition 1 Condition 2 Average
Relative Relative Relative
Lump [% wt.] Standard Standard
Average Variance Average Variance Variance
Deviation Deviation [%]
[%] [%]
Gases 4.3 0.04 1.0 3.7 0.05 1.2 1.1
Naphtha (28 - 190 °C) 8.1 0.36 4.4 6.1 0.73 11.9 8.1
Kerosene (190 - 260 °C) 8.5 0.02 0.2 5.9 0.58 9.9 5.1
Diesel (260 - 343 °C) 12.3 0.21 1.7 9.6 0.41 4.3 3.0
LVGO (343 - 453 °C) 15.0 0.25 1.7 13.1 0.28 2.1 1.9
HVGO (453 - 560 °C) 19.0 0.07 0.4 20.0 0.39 1.9 1.2
DAO (560°C+) 16.8 0.61 3.6 29.1 0.60 2.1 2.8
Asphaltenes-C5 15.6 0.90 5.8 11.7 0.61 5.2 5.5

75
Figure 4.12 presents the average lump composition for both experimental tests with the error
bars calculated using the average relative variance reported in Table 4.4. As can be observed, the
variability in terms of weight fraction is considerably higher for the residual lumps DAO (560
°C+) and the Asp-C5. In addition, the naphtha cut presents a significant variability. On the other
hand, the weight fraction variability of intermediate lumps (kerosene, diesel, LVGO and HVGO)
is lower than DAO (560 C+), Asp-C5 and naphtha. The lowest variability was obtained for the
gases lump.

25
Experimental Compotiion [% wt.]

20

15

10

Figure 4.12. Average lump composition with error bars

The global activation energy for the conversion of DAO (560ºC+) to other lumps is 241
[kJ/mol] and 226 [kJ/mol] for converted DAO and virgin DAO regardless of the evaluated case,
which both fall within the common range for thermal cracking reactions, as well as the activation
energies of conversion of other lumps.9 The most important finding of the type of modeling
adopted in this work (including an asphaltenes production route), is getting to know how fast and
how likely it is to produce an undesired amount of asphaltenes compared to the conversion
reactions of DAO (560 ºC+) to lower boiling point range hydrocarbons that are economically
valuable. As presented in Table 4.5 and Table 4.6 for converted and virgin DAO kinetic
modeling respectively, this reaction is in the same order of magnitude as the conversion

76
reactions, which definitely is a limitation of the thermal cracking process since considerable
amounts of non-virgin asphaltenes are being produced to a similar extent as that of valuable
products. Careful selection of process operation conditions (temperature, residence time) has to
be done in such a way that this reaction step leading to asphaltenes generation is inhibited
keeping the produced asphaltenes content as low as possible while the others pathways leading to
liquid products are favored, producing in this way a higher amount of desirable products
preferably within the gas oil and distillates range. It is very important to indicate how the
chemical composition of the feedstock can impact the velocity of generation of asphaltenes. For
the different evaluated scenarios for converted and virgin DAO kinetic modeling, it can be
observed that the ratio of kinetic constants for asphaltenes production (𝑘1 ) from converted DAO
(560 ºC+) and virgin DAO (560 ºC+) is around 2. This indicates that even though both
feedstocks lack asphaltenes and come from similar reservoirs they have different chemical
characteristics that lead to considerably higher production of asphaltenes from the converted
material compared to the virgin one. This could be explained considering how each feedstock is
produced following the upgrading scheme shown in Figure 4.1. Virgin DAO comes from direct
solvent deasphalting of vacuum residue (560ºC+) using pentane, but without having the thermal
cracking recycling loop. On the other hand, for the converted DAO, deasphalted oil is kept on a
thermal cracking/separation loop in such a way that only the lighter hydrocarbons are produced
(gas, naphtha, kerosene, diesel, LVGO and HVGO), the asphaltenes are rejected along with some
resins for gasification and the DAO (560ºC+) is kept in the thermal cracking loop having always
some proportion of fresh bitumen coming in. In this way, converted DAO (560ºC+) would be
composed of a portion of virgin or unprocessed hydrocarbons and a majority of material that has
been cracked or partially converted several times (that is why this feedstock is named converted
DAO). In this order of ideas, the reactivity of this converted DAO (560ºC+) lump would come
from cracking of heavy aromatic and poly-aromatic hydrocarbons since dealkylation and
removal of easy-to-crack hydrocarbons has been occurring for several cycles. Thermal cracking
of these heavy compounds promotes faster polymerization and condensation reactions to form
asphaltenes that are precursors to form insoluble materials such as coke, compared to a virgin
DAO feedstock which still possesses alkyl moieties able to react initially.

Taking a look at the kinetic parameters presented in Table 4.5 and Table 4.6, the following
remarks can be advanced:

77
Case 1 (No sequence): a decreasing trend can be observed when comparing the activation
energies of conversion of DAO (560ºC+) to heavier compounds compared to lighter ones for
converted DAO, while the opposite behavior was observed for virgin DAO. No clear pattern was
observed for the conversion of the HVGO lump for both feedstocks, whereas LVGO showed
lower activation energies towards the production of heavier lumps for converted DAO and the
opposite behavior was found for virgin DAO. Additionally, activation energies of conversion of
heavier lumps are lower compared to lighter ones for converted DAO with the following values:
241, 278 and 269 [kJ/mol] for DAO (560ºC+), HVGO and LVGO respectively. On the contrary,
for virgin DAO activation energies of conversion of heavier lumps are higher compared to
lighter ones for virgin DAO, these are: 226, 214 and 188 for DAO (560ºC+), HVGO and LVGO
respectively. In case 1, no constraints were placed on the mathematical model for the reactions
leading to gas production and, as can be observed on the kinetics constants at 423 ºC, the heavier
the hydrocarbon the lower its kinetic constant for gas production (𝑘7 < 𝑘12 < 𝑘16 ) for both
feedstocks.

Table 4.7 presents the kinetic parameters for converted DAO thermal cracking without
imposing any constrain in the optimization algorithm (case 1) with and without re-
parameterization of the Arrhenius expression. As observed in the table, no significant changes
were obtained for the pre-exponential factor and activation energies calculated when using the
re-parameterized form of the Arrhenius equation in the optimization algorithm. One possible
explanation for this observation is that, in this work, 𝐸𝑎𝑖 and 𝐴𝑖 are not the independent
variables used in the optimization algorithm to minimize the sum of squared errors (SSE)
between experimental and modeled compositions. Instead, the algorithm finds the reaction rate
constants at each temperature and then Eq. 4.16 is used to estimate the activation energy (𝐸𝑎𝑖 )
and the pre-exponential factor (𝐴𝑖 ) by using linear regression of the Arrhenius equation. If the
minimization algorithm was expressed as a function of the activation energy and the pre-
exponential factor for each reaction, there could be a possibility of having a high correlation
between 𝐸𝑎𝑖 and 𝐴𝑖 . However, as indicated in Section 4.3.4, that approach was not followed in
this work to avoid having an increased number of parameters to be optimized, as well as to avoid
introducing high non-linearities in the system of ODEs with the exponential form of the
Arrhenius expression.

78
Case 2 (Sequence 1): activation energies of conversion of heavier lumps are lower
compared to lighter ones for both feedstocks. The following values were obtained: 241, 277 and
300 for DAO (560ºC+), HVGO and LVGO respectively for converted DAO and 226, 246 and
280 [kJ/mol] for virgin DAO. Therefore, the constraints imposed on the mathematical modeling
to obtain higher activation energies for conversion of lighter lumps were satisfied in both cases.
Likewise, the sequences implemented for conversion of each lump were satisfied since for each
of the reacting lumps the activation energies have a decreasing trend to lighter lumps compared
to heavier ones. Moreover, the constraints on the kinetic constants to gas production
(𝑘7 , 𝑘12 & 𝑘16) were also satisfied and gas production proceeds more readily from heavier lumps.
Analyzing the values of the kinetic constants at 423 ºC, it can be observed that for converted
DAO, DAO (560ºC+) reacts faster towards HVGO, LVGO and the undesired reaction to Asp-C5
compared to any other lumps, so the selectivity towards distillates and naphtha is low. Similarly,
these reactions are also predominant for the fraction DAO (560ºC+) of the virgin material.
However, in the case of virgin DAO the velocity of production of HVGO and LVGO is higher
compared to the reaction to Asp-C5. Also, the generation of naphtha and distillates from virgin
and converted DAO (560ºC+) is at similar levels, while the gas generation is lower for the virgin
material. In the same way, HVGO converts more readily to its subsequent lighter lump LVGO
compared to any other lump, with kinetic constant to LVGO more than 4 times higher than to
diesel, kerosene, naphtha for converted DAO and twice higher for virgin DAO following the
same pattern. Conversely, LVGO thermal cracking has higher selectivity towards kerosene and
naphtha as compared to diesel for both evaluated feedstocks.

Case 3 (Sequence 2): Similar to case 2, activation energies of conversion of DAO (560ºC+),
HVGO and LVGO followed a decreasing trend. The following values were obtained: 241, 243
and 249 [kJ/mol] for DAO (560ºC+), HVGO and LVGO respectively for converted DAO and
226, 230 and 244 [kJ/mol] for virgin DAO. In the same way, the mathematical algorithm
satisfied the constraints imposed. On the other hand, the sequences implemented in the
mathematical modeling for conversion of each lump were also satisfied since for DAO (560ºC+),
HVGO and LVGO the activation energies have an increasing trend to lighter lumps compared to
heavier ones. Similar to case 2, constraints on the kinetic constants for gas production were
satisfied. Overall, selectivities of each lump followed a similar pattern as in case 2, with some
slight variations on the actual values of the kinetic constants.

79
The results obtained in case 1 show that no pattern would align with the thermal cracking
chemistry if no restrictions are placed during the mathematical modeling. The algorithm would
obtain a solution within a reasonable modeling error, but it would not be possible to give a
physical explanation to the results, especially for complex systems involving several lumps and
reactions as those evaluated in this work. For example, kinetic constants for gas production from
lighter compounds are higher than from heavier compounds, which would not be the expected
trend for visbreaking of residual hydrocarbons since gas production proceeds more readily from
heavy fractions.2,30,31 Additionally, even though it would depend on the type of bonds that are
broken in each of the lumps, overall it is expected to have lower activation energies for
conversion of heavier lumps due to the increased crackability of higher boiling point range
hydrocarbons compared to lower ones. As can be seen for case 1, this trend was not obtained for
the three reacting lumps for either of the feedstocks. Therefore, it is necessary to take into
consideration the chemical aspects that can be possibly incorporated into the mathematical
solution of the lumped kinetic model and constrain it in such a way that it fulfills these
requirements. This is in agreement with what was found by Asaee et al.29 for hydrocracking of
heavy residua. The sequence of activation energies in case 2 suggests that the production of
lighter compounds comes from the thermal cracking of bonds that require less energy to break as
compared to the production of heavier lumps. For example, a hydrocarbon with significant
content of asphaltenes that are known to have high presence of alkyl chains would have low
activation energy towards the production of naphtha and gases compared to hydrocarbons within
the HVGO or LVGO range where their production would come from breakage of internal C-C
bonds. It is important to highlight that most poly-aromatic rings with high content of alkyl chains
are commonly present on the asphaltenes fraction that are absent in both feedstock evaluated in
this work. However, asphaltenes generated during thermal cracking would not have the same
characteristic since their structural composition would be considerably lower in alkyl moieties
compared to a virgin asphaltenes. On the contrary, the sequence in case 3, where activation
energies for production of lighter lumps are higher than those for heavier lumps, would be
suitable for example for a feedstock where there is a lower chance of dealkylation reactions
taking place. In this case, cracking of hydrocarbon molecules towards production of light lumps
would require more energy to occur and therefore would become significant at higher
temperatures. Therefore, case 3 would be a more appropriate approach for kinetic modeling of

80
DAO. Taking a closer analysis on the DAO (560°C+) lump, which is the lump of most interest, if
it is assumed that a converted and virgin DAO are composed by only this fraction (perfect
distillation cut at 560 °C), the initial velocity of formation of each lumps at 423 °C can be
calculated using Eq. 4.6 and compared for each lump relative to the global velocity of reaction
for both feedstock. This allows a simple comparison of the preferential pathways using the
modeling results assuming both feedstocks had the same lump composition. As illustrated in
Figure 4.13, production of asphaltenes almost doubles and production of HVGO and LVGO is
considerably lower when comparing converted and virgin DAO. This confirms a strong change
on selectivities for both deasphalted materials with the same boiling point range due to their
differences in chemical nature caused by the sequential cracking of the converted DAO as
explained previously.

Figure 4.13. Relative initial velocities of formation of each lump at 423 °C for DAO (560 °C+) as
percentage of the global initial velocity.
All these aspects are highly dependent on the detailed characterization of the evaluated
material at a molecular level, which is very often not feasible for kinetic modeling based on a
lumping technique where thousands of different compounds are grouped in a single lump.
Nonetheless, this type of models play an important role for simulation and optimization of
industrial processes and still bulk chemical characteristics of the feedstock can be implemented
in the modelling to obtain a solution that makes physical sense for the process under study as
was done in this investigation.

81
Table 4.5. Estimated kinetic parameters for case 1, case 2 and case 3 for converted DAO
-1
Converted DAO Kinetic constants at 423 °C [h ] Case 1 Case 2 Case 3
Ea Ln A Ea Ln A Ea Ln A
2 2 2
Reaction Case 1 Case 2 Case 3 r r r
[kJ/mol] [A in h-1] [kJ/mol] [A in h-1] [kJ/mol] [A in h-1]

r1 DAO (560 C+) → Asp -C5 0.5952 0.5969 0.5975 248 42.27 0.9955 247 42.17 0.9952 246 42.04 0.9953
r2 DAO (560 C+) → HVGO 0.5682 0.4556 0.4003 273 46.53 0.9997 266 45.11 0.9995 235 39.68 0.9994
r3 DAO (560 C+) → LVGO 0.4223 0.3724 0.3686 246 41.74 0.9990 245 41.40 0.9992 239 40.33 0.9992
r4 DAO (560 C+) → Diesel 0.2021 0.1943 0.1985 253 42.25 0.9984 230 38.09 0.9985 240 39.88 0.9988
r5 DAO (560 C+) → Kerosene 0.0525 0.0896 0.0692 230 36.77 0.9998 219 35.40 0.9996 241 38.96 0.9997
r6 DAO (560 C+) → Naphtha 0.0686 0.1333 0.1732 143 22.04 0.9998 215 35.19 0.9988 242 40.03 0.9983
r7 DAO (560 C+) → Gas 0.0965 0.1632 0.1980 177 28.26 0.9998 214 35.09 0.9994 244 40.54 0.9995
r8 HVGO → LVGO 0.5978 0.4202 0.3061 293 50.21 0.9995 307 52.25 0.9996 237 39.85 0.9998
r9 HVGO → Diesel 0.0882 0.1009 0.0933 219 35.43 1.0000 277 45.51 0.9999 240 39.14 0.9999
r10 HVGO → Kerosene 0.1410 0.1272 0.1608 243 39.97 0.9998 259 42.66 0.9998 243 40.21 0.9997
r11 HVGO → Naphtha 0.1789 0.0985 0.0913 307 51.40 0.9995 251 41.10 0.9999 249 40.55 0.9999
r12 HVGO → Gas 0.1523 0.1452 0.1221 284 47.15 0.9999 245 40.46 0.9999 260 42.81 0.9999
r13 LVGO → Diesel 0.1402 0.1692 0.1549 202 33.00 1.0000 342 57.36 0.9999 241 39.76 1.0000
r14 LVGO → Kerosene 0.3707 0.2735 0.2978 217 36.43 0.9998 302 50.86 0.9998 245 41.08 0.9999
r15 LVGO → Naphtha 0.4268 0.2913 0.1476 341 58.18 0.9995 289 48.77 0.9998 255 42.07 1.0000
r16 LVGO → Gas 0.4225 0.1119 0.0564 341 58.16 0.9998 277 45.63 1.0000 299 48.85 1.0000

kglobal 2.0054 2.0054 2.0054 GAAE[%] 6.72 GAAE[%] 7.00 GAAE[%] 7.02

82
Table 4.6. Estimated kinetic parameters for case 1, case 2 and case 3 for virgin DAO

Virgin DAO Kinetic Constants at 423 °C [h-1] Case 1 Case 2 Case 3


Ea Ln A Ea Ln A Ea Ln A
Reaction Case 1 Case 2 Case 3 r2 r2 r2
[kJ/mol] [A in h-1] [kJ/mol] [A in h-1] [kJ/mol] [A in h-1]
r1 DAO (560 C+) → Asp -C5 0.3125 0.3133 0.3118 228 38.20 0.9998 229 38.45 0.9998 226 37.95 0.9998
r2 DAO (560 C+) → HVGO 0.6048 0.6095 0.5757 216 36.87 0.9982 239 40.89 0.9986 214 36.49 0.9984
r3 DAO (560 C+) → LVGO 0.4306 0.4381 0.4461 212 35.74 0.9990 222 37.60 0.9991 224 37.93 0.9991
r4 DAO (560 C+) → Diesel 0.2186 0.2041 0.2073 210 34.80 0.9997 215 35.59 0.9997 228 37.89 0.9997
r5 DAO (560 C+) → Kerosene 0.0780 0.0778 0.0781 240 38.89 0.9999 211 33.92 0.9999 235 38.05 0.9999
r6 DAO (560 C+) → Naphtha 0.0768 0.0643 0.0747 267 43.60 0.9999 208 33.13 0.9999 245 39.66 0.9999
r7 DAO (560 C+) → Gas 0.1163 0.1306 0.1438 342 57.01 0.9999 204 33.20 0.9997 262 43.27 0.9998
r8 HVGO → LVGO 0.5904 0.6198 0.4951 219 37.27 1.0000 301 51.48 1.0000 218 36.97 1.0000
r9 HVGO → Diesel 0.1737 0.2246 0.2036 249 41.20 1.0000 239 39.80 1.0000 223 36.96 1.0000
r10 HVGO → Kerosene 0.2736 0.2884 0.2843 218 36.28 1.0000 217 36.23 1.0000 229 38.31 1.0000
r11 HVGO → Naphtha 0.3210 0.2916 0.3532 238 40.01 1.0000 202 33.71 1.0000 238 40.11 0.9999
r12 HVGO → Gas 0.1401 0.0865 0.0995 118 18.39 1.0000 185 29.51 1.0000 279 45.94 1.0000
r13 LVGO → Diesel 0.2094 0.2509 0.2609 291 48.69 1.0000 346 58.32 1.0000 227 37.92 1.0000
r14 LVGO → Kerosene 0.4335 0.4401 0.4542 187 31.50 1.0000 292 49.57 1.0000 237 40.16 1.0000
r15 LVGO → Naphtha 0.5788 0.7458 0.5751 177 30.11 1.0000 260 44.65 0.9999 252 42.97 1.0000
r16 LVGO → Gas 0.1566 0.0435 0.0480 123 19.39 1.0000 205 32.24 1.0000 314 51.28 1.0000

kglobal 1.8376 GAAE[%] 7.01 GAAE[%] 6.62 GAAE[%] 7.03

83
Table 4.7. Comparison of estimated kinetic parameters with and without using re-
parameterization of the Arrhenius equation for case 1 – converted DAO

Without re-parameterization With re-parameterization


Ea Ln A Ea Ln A
Reaction
[kJ/mol] [A in h-1] [kJ/mol] [A in h-1]
r1 DAO (560 C+) → Asp -C5 248 42.27 248 42.27
r2 DAO (560 C+) → HVGO 273 46.53 273 46.52
r3 DAO (560 C+) → LVGO 246 41.74 246 41.73
r4 DAO (560 C+) → Diesel 253 42.25 253 42.24
r5 DAO (560 C+) → Kerosene 230 36.77 231 36.90
r6 DAO (560 C+) → Naphtha 143 22.04 143 22.05
r7 DAO (560 C+) → Gas 177 28.26 177 28.27
r8 HVGO → LVGO 293 50.21 293 50.17
r9 HVGO → Diesel 219 35.43 219 35.45
r10 HVGO → Kerosene 243 39.97 243 39.96
r11 HVGO → Naphtha 307 51.40 308 51.49
r12 HVGO → Gas 284 47.15 284 47.15
r13 LVGO → Diesel 202 33.00 202 33.01
r14 LVGO → Kerosene 217 36.43 216 36.35
r15 LVGO → Naphtha 341 58.18 341 58.09
r16 LVGO → Gas 341 58.16 342 58.18

GAAE[%] 6.72 GAAE[%] 6.72

Table 4.8. Kinetic parameters for case 3 - converted DAO using a 6-lumps kinetic model

k at 423 °C Ea Ln A
Reaction
[h-1] [kJ/mol] [A in h-1]

r1 DAO (560 °C+) → Asp -C5 0.5940 247 42.08


r2 DAO (560 °C+) → VGO 0.7900 238 40.85
r3 DAO (560 °C+) → Distillates 0.3004 239 40.14
r4 DAO (560 °C+) → Naphtha 0.1112 241 39.20
r5 DAO (560 °C+) → Gas 0.2098 244 40.63
r6 VGO → Distillates 0.2681 239 40.06
r7 VGO → Naphtha 0.2062 244 40.69
r8 VGO → Gases 0.0787 277 45.46

GAAE
8.20
[%]

84
As presented in Table 4.8, where the kinetic parameters for converted DAO thermal
cracking using a 6-lumps kinetic model are reported, all the constraints imposed in the
optimization algorithm were satisfied. This result is similar to the obtained for the 8-lumps
model. In the same way, the values for 𝐸𝑎𝑖 and 𝐴𝑖 are very similar to the obtained for the 8-
lumps model. Taking a look at the kinetic constants at 423 °C for both 8-lumps and 6-lumps
models, if the kinetic constants of formation of HVGO and LVGO from DAO (560 °C+) for the
8-lumps model are grouped (𝑘2 + 𝑘3 = 0.4003 + 0.3686 = 0.7689 ℎ−1 ), the actual value is
very close to the obtained for the production of full-range VGO (343+560) using the 6-lumps
model (𝑘2 = 0.7900 ℎ−1 ). A similar result is obtained if diesel and kerosene are grouped. In this
case, the grouped value of the formation of diesel and kerosene from DAO (560 °C+) for the 8-
lumps model (𝑘4 + 𝑘5 = 0.1985 + 0.0692 = 0.2377ℎ−1 ) is in the same order of magnitude to
the obtained for the production of distillates using the 6-lumps model (𝑘3 = 0.7900 ℎ−1 ).
Formation of gases and asphaltenes from DAO (560 °C+) is also very similar for both models.

0.6
Converted DAO - 6 Lumps
Model Composition [weight fraction]

0.5 y = 1.0503x - 0.0084


R² = 0.9943
0.4
DAO (560 °C+)
Asp-C5
0.3
VGO

0.2 Distillates
Naphtha
0.1 Gases

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6
Experimental Composition [weight fraction]

Figure 4.14. Predicted model compositions vs. experimental compositions for case 3 - converted
DAO for a 6-lumps kinetic model.

85
Based on the previous observations, there is no significant difference in the kinetic
parameters obtained when using a robust 8-lumps model, which represents in more detail the
transformations of the different fractions present in the DAO feedstock. From an error analysis
point of view, a very good fit between the experimental and model data was obtained for the 6-
lumps model as depicted in Figure 4.14. The linear regression of the plotted data has a
coefficient of correlation close to 1. Comparing the global average absolute errors (GAAE), it is
very interesting to see that for the 6-lumps model, the GAAE is higher than the one obtained for
the 8-lumps model (8.20 vs. 7.02 %), which could indicate that using 8-lump represents better
the thermal cracking reactions taking place. It is important to highlight that, contrary to what
could be expected when having a high number of lumps, the quality of the adjustment is better
when the definition of the pseudo components is more precise. This gives an indication that the
system tolerates a higher number of pseudo components to represent the reactions taking place.

4.5. Conclusions

Reactivity of deasphalted vacuum residue from Northern Alberta spanning a relatively wide
range of conversion levels showed a behavior commonly found for upgrading of residual
hydrocarbon fractions via thermal cracking with an enlarged stability limit as it corresponds with
a negligible or lower content of asphaltenes, initially. Total operating pressure and steam partial
pressure did not have a significant effect on the DAO reactivity at the range of operating
conditions evaluated. Despite the absence of asphaltenes on the two evaluated feedstock,
converted DAO and virgin DAO, a considerable yield of these compounds was obtained when
the pre-cracked feedstock was used, which highlights the importance of this undesirable reaction
in the reactions scheme. Very importantly, it was found that significant differences in
asphaltenes and gas oil production are obtained for materials with similar boiling point ranges,
however having differences in their chemical nature with converted feed expectably more
aromatic and less branched than virgin feed. From the kinetic point of view, it was possible to
incorporate the reaction of coke precursors (asphaltenes) generation using the lumping technique
for vacuum residue DAO thermal cracking with a reasonable modeling error around 7 % and for
three different scenarios of sequences on the activation energies and restrictions on kinetic
constants to account for chemical considerations. Kinetic parameters were within common
values found in the literature. Case 3, which incorporate modifications based on chemical

86
reactivity considerations, was found to be the most appropriate for kinetic modeling of
deasphalted oil thermal cracking. It is significant that while the activation energies for the
conversion of heavier lumps are lower than those for conversion of lighter ones, the activation
energies of formation of lighter lumps are higher than those for the formation of heavier lumps.
No relevant differences were observed between the kinetic parameters obtained for the 8-lumps
model compared to the 6-lumps approach for DAO thermal cracking. This indicates that in this
particular case, the quality of the adjustment is better when the definition of the pseudo
components is more precise, having a higher number of lumps to model the thermal cracking
reactions taking place. In the same way, re-parameterization of the Arrhenius equation did not
have any significant effect on the calculated kinetic parameters.

4.6. Acknowledgement

The authors are grateful to the Natural Sciences and Engineering Research Council of
Canada (NSERC), Nexen-CNOOC Ltd, and Alberta Innovates-Energy and Environment
Solutions (AIEES) for the for the financial support provided through
the NSERC/NEXEN/AIEES Industrial Research Chair in Catalysis for Bitumen Upgrading.
Also, the contribution of facilities from the Canada Foundation for Innovation, the Institute for
Sustainable Energy, Environment and Economy, the Schulich School of Engineering and the
Faculty of Science at the University of Calgary are greatly appreciated. Finally, the technical
contributions provided by Mr. Nestor Zerpa, Senior Engineering Advisor at Nexen Energy ULC,
for the development of this investigation is greatly appreciated.

4.7. References

(1) Ancheyta, J.; Trejo, F.; Rana, M. S., Asphaltenes: Chemical Transformation During
Hydroprocessing of Heavy Oils. Taylor & Francis Group: 2009.
(2) Speight, J. G. Visbreaking: A technology of the past and the future. Scientia Iranica 2012, 19
(3), 569-573.
(3) Speight, J. G., Chapter 2 - Thermal Cracking. In Heavy and Extra-heavy Oil Upgrading
Technologies, Gulf Professional Publishing: Boston, 2013; pp 15-38.
(4) Puron, H.; Arcelus-Arrillaga, P.; Chin, K. K.; Pinilla, J. L.; Fidalgo, B.; Millan, M. Kinetic
analysis of vacuum residue hydrocracking in early reaction stages. Fuel 2014, 117, Part A (0),
408-414.
(5) Sánchez, S.; Rodríguez, M. A.; Ancheyta, J. Kinetic Model for Moderate Hydrocracking of
Heavy Oils. Ind. Eng. Chem. Res. 2005, 44 (25), 9409-9413.
(6) Ancheyta, J.; Sánchez, S.; Rodríguez, M. A. Kinetic modeling of hydrocracking of heavy oil
fractions: A review. Catal. Today 2005, 109 (1–4), 76-92.

87
(7) Souza, B. M.; Travalloni, L.; da Silva, M. A. P. Kinetic Modeling of the Thermal Cracking of
a Brazilian Vacuum Residue. Energy Fuels 2015, 29 (5), 3024-3031.
(8) Taghipour, A.; Naderifar, A. Kinetic Modeling of Vacuum Residue Thermal Cracking in the
Visbreaking Process Using Multiobjective Optimization. Energy Technology 2015, 3 (7), 758-
767.
(9) Singh, J.; Kumar, S.; Garg, M. O. Kinetic modelling of thermal cracking of petroleum
residues: A critique. Fuel Process. Technol. 2012, 94 (1), 131-144.
(10) Kataria, K. L.; Kulkarni, R. P.; Pandit, A. B.; Joshi, J. B.; Kumar, M. Kinetic Studies of
Low Severity Visbreaking. Ind. Eng. Chem. Res. 2004, 43 (6), 1373-1387.
(11) Kerr, R.; Birdgeneau, J.; Batt, B.; Yang, P.; Nieuwenburg, G.; Rettger, P.; Arnold, J.;
Bronicki, Y., The Long Lake Project - The First Field Integration of SAGD and Upgrading. In
SPE International Thermal Operations and Heavy Oil Symposium and International Horizontal
Well Technology Conference Calgary, Alberta, Canada, 2002.
(12) Del Bianco, A.; Panariti, N.; Anelli, M.; Beltrame, P. L.; Carniti, P. Thermal cracking of
petroleum residues. Fuel 1993, 72 (1), 75-80.
(13) Singh, J.; Kumar, M. M.; Saxena, A. K.; Kumar, S. Reaction pathways and product yields in
mild thermal cracking of vacuum residues: A multi-lump kinetic model. Chem. Eng. J. 2005, 108
(3), 239-248.
(14) Mohaddecy, S. R. S.; Sadighi, S. Simulation and Kinetic Modeling of Vacuum Residue
Soaker-Visbreaking. Petroleum and Coal 2011, 53 (1), 26-34.
(15) Loria, H.; Trujillo-Ferrer, G.; Sosa-Stull, C.; Pereira-Almao, P. Kinetic Modeling of
Bitumen Hydroprocessing at In-Reservoir Conditions Employing Ultradispersed Catalysts.
Energy Fuels 2011, 25 (4), 1364-1372.
(16) Da Silva De Andrade, F. Kinetic modeling of catalytic in situ upgrading for Athabasca
bitumen, deasphalting pitch and vacuum residue. MSc Thesis, University of Calgary, Calgary,
2014.
(17) Galarraga, C. E.; Scott, C.; Loria, H.; Pereira-Almao, P. Kinetic Models for Upgrading
Athabasca Bitumen Using Unsupported NiWMo Catalysts at Low Severity Conditions. Ind. Eng.
Chem. Res. 2011, 51 (1), 140-146.
(18) Fathi, M. M.; Pereira-Almao, P. Kinetic Modeling of Arab Light Vacuum Residue
Upgrading by Aquaprocessing at High Space Velocities. Ind. Eng. Chem. Res. 2012, 52 (2), 612-
623.
(19) Fathi, M. M. Comparative Upgrading of Arab Light Vacuum Residuum via Aquaprocessing
and Thermal Cracking. PhD Thesis, University of Calgary, Calgary, 2011.
(20) Trujillo-Ferrer, G. Thermal and Catalytic Steam Reactivity Evaluation of Athabasca
Vacuum Gasoil. MSc Thesis, University of Calgary, Calgary, 2008.
(21) Fathi, M. M.; Pereira-Almao, P. Catalytic Aquaprocessing of Arab Light Vacuum Residue
via Short Space Times. Energy Fuels 2011, 25 (11), 4867-4877.
(22) Di Carlo, S.; Janis, B. Composition and visbreakability of petroleum residues. Chem. Eng.
Sci. 1992, 47 (9), 2695-2700.
(23) Carbognani, L.; Lubkowitz, J.; Gonzalez, M. F.; Pereira-Almao, P. High Temperature
Simulated Distillation of Athabasca Vacuum Residue Fractions. Bimodal Distributions and
Evidence for Secondary “On-Column” Cracking of Heavy Hydrocarbons. Energy Fuels 2007, 21
(5), 2831-2839.

88
(24) Hassan, A.; Carbognani, L.; 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.
(25) Carbognani Ortega, L.; Rogel, E.; Vien, J.; Ovalles, C.; Guzman, H.; Lopez-Linares, F.;
Pereira-Almao, P. Effect of Precipitating Conditions on Asphaltene Properties and Aggregation.
Energy Fuels 2015, 29 (6), 3664-3674.
(26) Carbognani, L.; Gonzalez, M. F.; Pereira-Almao, P. Characterization of Athabasca Vacuum
Residue and Its Visbroken Products. Stability and Fast Hydrocarbon Group-Type Distributions.
Energy Fuels 2007, 21 (3), 1631-1639.
(27) Blasetti, A.; de Lasa, H. FCC Riser Unit Operated in the Heat-Transfer Mode:  Kinetic
Modeling. Ind. Eng. Chem. Res. 1997, 36 (8), 3223-3229.
(28) Agarwal, A. K.; Brisk, M. L. Sequential experimental design for precise parameter
estimation. 1. Use of reparameterization. Industrial & Engineering Chemistry Process Design
and Development 1985, 24 (1), 203-207.
(29) Asaee, S. D. S.; Vafajoo, L.; Khorasheh, F. A new approach to estimate parameters of a
lumped kinetic model for hydroconversion of heavy residue. Fuel 2014, 134, 343-353.
(30) Speight, J. G.; Ozum, B., Petroleum Refining Processes. Taylor & Francis: 2001.
(31) Gray., M. R., Upgrading oilsands bitumen and heavy oil. The University of Alberta Press:
Edmonton, 2015.

89
Chapter 5: Catalytic Steam Cracking of Deasphalted Vacuum Residue using a Ni/K
Ultradispersed Catalyst

Fredy A. Cabrales-Navarro* and Pedro Pereira-Almao


Department of Chemical and Petroleum Engineering, Schulich School of Engineering,
University of Calgary, 2500 University Drive N.W., Calgary, AB T2N 1N4 , Canada
* Corresponding author (facabral@ucalgary.ca)

5.1. Abstract

Catalytic Steam Cracking (CSC) of heavy hydrocarbons is seen as an alternative for further
improvement upon conventional thermal cracking performance. In this work, upgrading of an
industrial deasphalted vacuum residue via CSC was assessed in a bench-scale pilot plant
resembling a visbreaking unit. The performance of a 400 ppm Ni and 300 ppm K ultradispersed
catalyst (UDC) formulation previously used for CSC of vacuum residue was evaluated for this
non-asphaltene containing fraction. Reactivity experiments were conducted at temperatures
within 435-445°C and Liquid Hourly Space Velocities (LHSV) of 3-5.5 h-1 and operating
pressure of 300 psig. A preliminary reactivity evaluation using isotopic water spanning
temperatures between 423 °C and 445 °C was carried out to determine the conditions at which
water splitting was occurring. Finally, lumped kinetic modeling including asphaltenes generation
in the process was evaluated and results were compared with previously reported thermal
cracking experiments. Operating variables (T, LHSV) were found to have similar effects on the
reactivity as in thermal cracking for CSC of DAO. Even though water splitting was evidenced at
temperatures above 430 °C, no significant improvement on bulk properties of the liquid products
was obtained for the catalytic experiments using the current formulation. This is attributed to the
high degree of condensation reactions triggered at the range of temperatures evaluated. A global
activation energy for the conversion of DAO (560 °C+) of 175 kJ/mol and a modeling error of
4.23 % were determined. Asphaltenes generation was evidenced at a similar extent as that of
thermal cracking from a kinetic point of view.

Keywords: DAO, Asphaltenes, Lump, Catalytic Steam Cracking.

90
5.2. Introduction

Sustainable exploitation of extra-heavy oil and bitumen via enhanced oil recovery methods
has increased considerably in the last decade to fulfill the supply gap due to the shortage of light
oil reserves that has forced the oil sector to move toward enhanced oil recovery methods for light
oils, such as reservoir fracturing. These heavy materials are characterized by having considerable
amounts of asphaltene compounds that make difficult their processing under conventional
refinery schemes. In this way, the implementation of deasphalting units, where asphaltene
molecules are removed from the hydrocarbon fraction, has become popular in the industry,
demanding processes for upgrading these non-asphaltene containing heavy hydrocarbons, for
which an alternative technology involving innovative ultra-dispersed catalysts and steam
activation, such as catalytic steam cracking, may have a very important role.

Catalytic steam cracking of heavy hydrocarbon fractions is a novel technique developed by


Pereira-Almao et al.1, in which a combination of conventional thermal cracking and mild
hydrocracking processes was implemented. In this technology, a bi-metallic ultradispersed
catalyst, typically combinations of alkali metals (such as K) and transition metals (such as Ni),
are incorporated in the feedstock of interest in the form of unsupported particles at a submicronic
scale. This bimetallic Ni/K catalyst has two functionalities: water splitting carried out by the K
particles present in the form of potassium oxides, and a hydrogenating function provided by the
Ni in the form of nickel oxide. The catalytic feed is then submitted to reaction conditions similar
to a conventional thermal cracking under the presence of steam, which is catalytically dissociated
and serves as the hydrogen donor for saturation of hydrocarbon radicals formed by the thermal
cracking. Hydrogen incorporation is much lower than in a conventional hydroconversion
processing. However, the process conditions are much less severe, thus not only the capital cost
to implement the process is lower, but also the operational expenses, which could make it
economically attractive. In 2001, Pereira-Almao et al.2 carried out a complete evaluation of the
technical benefits and economic potential of this technology for Venezuelan extra-heavy oil from
the Orinoco belt and concluded that this upgrading technology was feasible in both aspects for
the Venezuelan oil industry scenario. One very important advantage of this technology compared
to conventional thermal cracking (visbreaking) is that the conversion level can be considerably

91
increased, which reflects in higher distillates content, while keeping low levels of undesirable
products such as coke, and maintaining the stability of the product at an acceptable value.3-5

The following mechanism of reaction, resulting from a combination of steam reforming and
thermal cracking reactions, has been proposed for the upgrading of heavy oil using ultradispersed
catalysts3,5,6

Thermal Cracking 𝑅 − 𝑅𝑛′ → 𝑅𝑛. + 𝑅𝑛′. Eq. 5.1

Catalytic Dissociation 𝐶𝐴𝑇


𝐻2 𝑂 → 𝐻 ∙ + 𝑂𝐻 ∙ Eq. 5.2
of Water
Free Radicals 𝐶𝐴𝑇
𝑅 ∙ + 𝑅𝑛′∙ + 2𝐻 ∙ → 𝑅 − 𝐻 + 𝑅𝑛′ − 𝐻 Eq. 5.3
Saturation
𝐶𝐴𝑇

Oxidation/Reforming 𝑅𝑛′∙ + 2𝑂𝐻 ∙ → 𝑅𝑛−1 + 𝐶𝑂2 + 𝐻2 Eq. 5.4

Condensation 𝑅𝑛′∙ + 𝑅 ∙ → 𝑅𝑛′ − 𝑅𝑛′ , 𝑅 − 𝑅 Eq. 5.5

From the mechanism of reaction, the effect of the bimetallic catalyst on the different
reactions is observed. The alkali catalyst (K) promotes reaction 2, increasing the production of
hydrogen free radicals that are used in reaction 3 for the saturation of the hydrocarbon free
radicals obtained from the thermal cracking reaction. The OH radicals also react with the
hydrocarbon free radicals by means of a typical reforming process to produce carbon dioxide and
hydrogen. These two steps are mainly promoted by the Ni metal catalyst. Failure of the catalyst
to provide an adequate dissociation of water in reaction No. 2 or to promote reactions 3 and 4
will result in hydrocarbon free radicals reacting with themselves according to the condensation
reaction 5 to form asphaltenes and coke, which is highly undesirable.

Trujillo-Ferrer3 applied the catalytic steam cracking technology for the upgrading of
Athabasca vacuum gas oil (VGO), in order to evaluate the applicability and the performance of
the process to this particular hydrocarbon fraction from the Alberta Oil Sands reserves. With his
work, Trujillo-Ferrer validated previous findings from the evaluations with Venezuelan extra-
heavy oil. It was possible to reach higher conversion levels than with thermal cracking by
increasing the severity of the process, while keeping a steady value of the microcarbon content,

92
which indicates lower tendency of the product to form coke by inhibition of the condensation
reactions taking place during thermal cracking as well as lower gas yields. Fathi and Pereira-
Almao5 studied the upgradability of an Arabian Light Vacuum Residue (ALVR) at short space
times, ranging from 5-10 h-1, using CSC. They were able to obtain a relative increase of 13% in
conversion of the residual fraction with catalytic steam cracking at the asphaltenes stability limit
using the K-Ni ultradispersed catalyst. In this study, the capability of the catalyst to dissociate
water and to provide hydrogen for the process was corroborated using isotopic water.

In the filed patent entitled “Systems and Methods for Catalytic Steam Cracking of Non-
asphaltene Containing Heavy Hydrocarbons” by Nexen Energy ULC6, three different schemes
for field upgrading of heavy and extra-heavy oils using CSC with ultradispersed catalysts are
proposed, all of them aiming to reach Canadian pipeline transportability conditions, which are:
19 API, 350 cSt at the carrier’s reference temperature and 1% wt olefins (1-decene equivalent).
Said schemes have the advantages of using CSC instead of other more expensive processes such
as hydrocracking. One key characteristic of the invention is that the CSC process is applied to
non-asphaltene containing fractions like VGO and deasphalted il (DAO), instead of directly to
the residual fractions such as VR, which allows one to go deeper in terms of conversion while
having a stable final product, based on the fact that even for CSC, processing of the asphaltenes
in the reaction is usually the limiting step of the process in terms of stability. This technology
development has highlighted the advantages of CSC as a feasible process in field upgrading
schemes due to its relative simplicity, which motivates the continuation of research in this area,
especially when non-asphaltene containing hydrocarbons are used as feedstock, since there is
limited information available in the literature related to their reactivity and kinetics.

This work targets the evaluation of the catalytic steam cracking route using a Ni/K catalyst
formulation for upgrading a deasphalted vacuum residue obtained from a bitumen upgrading
facility located in Northern Alberta, Canada. In the first preparation developed by Pereira-
Almao1, the catalyst is incorporated as a microemulsion. In this preparation, the microemulsion
is decomposed to form metal oxides that remain suspended as particles in the hydrocarbon. A
similar preparation procedure was followed by Fathi and Pereira-Almao4,5 with ALVR.
However, unlike their case where the procedure was in a batch mode, a completely continuous
preparation procedure was followed in this work using a skid preparation unit. In addition, static

93
mixers were used in this investigation to generate the precursor micro-emulsions. The main
objective of the work is to assess the capabilities of this process when asphaltenes are not present
in the feedstock as compared to the discussed previous works where the investigated materials
were VRs or VGO. Additionally, this catalytic approach is compared to a baseline case of
thermal cracking upgrading for a similar feedstock carried out in a previous work presented in
Chapter 4, looking forward to highlighting any possible benefits of a catalytic route for DAO
upgrading with non-supported catalysts from reactivity, product quality and kinetic points of
view.

5.3. Experimental Methods

5.3.1. Catalytic Feed Preparation

An industrial deasphalted vacuum residue (converted DAO) obtained from a bitumen


upgrading facility located in Northern Alberta, Canada was used as feedstock. The material
comes from solvent deasphalting of a mixture of unconverted vacuum residue from a thermally
cracked DAO plus virgin vacuum residue. For the incorporation of the ultra-dispersed catalyst
into the hydrocarbon feed, a skid preparation unit available in the Catalysis for Bitumen
Upgrading Group (CBU) in the Schulich School of Engineering at University of Calgary was
used. The first step in the preparation of the catalytic feed was to adjust the viscosity of the
deasphalted vacuum residue by using gasoline as diluent. This dilution was needed since the
equipment does not have the capability of heating up the feedstock, which is immobile at
ambient conditions. Then, diluted DAO was mixed with the aqueous solutions of the
corresponding metals that need to be incorporated, being nickel acetate and potassium hydroxide
(KOH) the precursor salts selected for incorporation of Ni and K particles. As illustrated in
Figure 5.1, the first precursor solution added was the potassium hydroxide and then nickel
acetate. This was to neutralize naphthenic acids present in the oil by reacting with KOH and
producing naphthenic salts that are known to act as natural surfactants that help emulsification of
the water in oil.7 The target concentration for the K and Ni was 400 ppm and 300 ppm
respectively. The water-in-oil transient emulsion was subsequently passed through a high
temperature reactor operating at 370 °C to decompose the metal precursors into metal oxides.
Next, the stream was sent to a hot separator where water and diluent were separated from the oil,
leaving the ultradispersed catalyst particles suspended in the DAO. The gaseous streams was

94
passed through a condenser to liquefy water and diluent and then to a cold separator where the
mixture of diluent and water was collected and gases produced in the decomposition of the
precursor salts were separated. The gas stream was bubbled in an aqueous KOH solution trap to
sweeten it before directing it to ventilation.

Figure 5.1. Catalyst incorporation unit – skid.

5.3.2. Reactivity Test Unit

Reactivity experiments were performed in a bench-scale pilot plant equipped with an up-flow
open tubular reactor. A general scheme of the reactivity test unit is shown in Figure 5.2. Details
of the experimental setup were provided in Section 2.1.

95
Figure 5.2. Schematic of the experimental setup for reactivity tests.

5.3.3. Characterization Techniques

5.3.3.1. Feedstock and Liquid Products

Common heavy oil characterization techniques were used to characterize the feedstock after
catalyst incorporation as well as the liquid product samples collected from each experiment.
Inductively Coupled Plasma (ICP) was used for the determination of metals concentration within
the UDC formulation. DAO samples were prepared for analysis using microwave-assisted
digestion following the ASTM D7455 norm. Subsequently, digested samples were analysed
following the ASTM D7691-05 method in a Thermoelectron IRIS Intrepid II spectrometer.
Product distribution based on boiling point was carried out following the ASTM D7169-05 norm
using an Agilent 6890N chromatograph as described by Carbognani et al.8 The P–value stability
index as described by Di Carlo and Janis9 was used as reference to determine the state of
peptization of asphaltenes. This method has been previously used to analyze samples of VR
CSC5 and DAO thermal cracking as in Chapter 4. Density at a temperature of 15.6 °C was
determined in a digital densitometer, Rudolph Research Analytical model DDM2911, as
explained in more detail elsewhere.10 This density was used later on to determine the API

96
following the definition in standard norm ASTM-D287-292. The Viscosity at 60°C for the
feedstock and 25°C for the liquid products after reaction was measured in a Brookfield
viscometer model DV-II+ assembled with a TC-502 water recirculation system for temperature
control. Hydrocarbon type distribution - Saturates/Aromatics/Resins/Asphaltenes (SARA)
analysis was performed using pentane as the solvent for asphaltenes precipitation following the
modification to the standard norm developed by Carbognani et al.11 for bitumen analysis. Solid
contents were quantified by vacuum filtration of a 1:50 (w/w) solution of DAO after reaction and
CHCl3 (99% Aldrich) in a 47 mm Millipore filter unit provided with 0.45 µm nylon membranes
from Pall Corporation, as explained by Galarraga et al.12 The mass of solids after filtration was
corrected by subtracting the theoretical amount of catalyst incorporated (Ni: as NiO and K as
K2O).

5.3.3.2. Gas Samples

Gas chromatography analysis was conducted inline to quantify the light hydrocarbons and
permanent gases generated for each experiment in an SRI Instruments chromatograph model
8610 C. Details of the specifications of the equipment were provided elsewhere (Section 2.2.3.1).
A Pfeiffer OmniStar quadrupole mass spectrometer (QMS) was used to analyze the gas streams
from thermal and catalytic steam cracking reactions carried out using normal as well as isotopic
water (H2O18 - 50% enriched in O18). As described by Fathi and Pereira-Almao5, this benchtop
equipment consists of a heated inlet system to prevent condensation of water and light
hydrocarbons in the capillary gas inlet tube, a PrismaPlus mass spectrometer, a dry-compressing
diaphragm vacuum pump, and a HiPace turbopump. The software enables scanning of atomic
masses ranging from 1 to 300 amu as well as data processing. As shown in Figure 5.2, the QMS
was connected to the gas outlet after the back-pressure valve for continuous evolved gas
analysis. For this work, the interest was to detect transfer of O18 coming from the splitting of the
isotopic water molecule to the carbon radicals generated by thermal cracking to form isotopic
CO2 as explained in detail in Section 5.3.4.1.

97
5.3.4. Experimental Plan

5.3.4.1. Evidence of Water Splitting with Isotopic Water

In order to corroborate water splitting promoted by the Ni/K UD catalyst, a mixture of 50%
isotopic water H2O18 and 50% demineralized water was used in a continuous catalytic steam
cracking run. As explained in the gas characterization procedure, a QMS was coupled to the gas
stream to monitor the signals of interest corresponding to the following compounds: H2 (m/e=2),
CO2 (m/e=44), CO16O18 (m/e=46) and CO218 (m/e=48). The pilot plant was started using DAO as
feedstock without any catalyst added, at a LHSV of 2 h-1 until a reaction temperature of 423 °C
was reached. LHSV was defined for all experiments at pumping conditions as expressed in Eq.
5.6. Then, a stabilization period of 2 h was allowed. Feedstock was switched to DAO with Ni/K
UDC (DAO+CAT), keeping the same experimental conditions and allowing enough time to
purge all feedlines and displace the DAO present in the reactor (approximately 140 mL in feed
lines and reactor). Subsequently, labeled water was added into the water reservoir and a 30 min
purge time was allowed to evacuate the pure demineralized water from the feedlines. A data
collection time of 30 min was allowed at this temperature. Next, the reaction temperature was
increased up to 430 and 437 °C always ensuring 30 min of data collection once the temperature
was stable. Finally, LHSV was increased to 3 h-1 and temperature up to 440 °C and reaction data
was collected for 30 min. During this last step, labeled water was depleted and regular
demineralized water was poured into the water reservoir to continue the experiment. Finally, the
pilot plant underwent a systematic shutdown procedure at the end of the experiment.

𝐹𝑙𝑜𝑤 𝑜𝑓 𝐷𝐴𝑂 + 𝐶𝐴𝑇 𝑎𝑡 140 °𝐶 𝑎𝑛𝑑 300 𝑝𝑠𝑖𝑔 [𝑚𝑙/ℎ]


LHSV [ℎ−1 ] = Eq. 5.6
𝑅𝑒𝑎𝑐𝑡𝑜𝑟 𝑉𝑜𝑙𝑢𝑚𝑒 [𝑚𝑙]

5.3.4.2. Reactivity Experiments

For the catalytic evaluation of the DAO reactivity, the Ni/K catalyst matrix was incorporated
into the feed using a skid preparation unit as explained in Section 5.3.1, targeting a concentration
of 300 ppm of Ni and 400 ppm of K. As in previous works3,4, a 5 % wt. of water was used. A
total of three different temperatures were evaluated, these are: 435, 440 and 445 C. Three Liquid
Hourly Space Velocities (LHSV) between 3 to 5.5 h-1 defined at pumping conditions (140 °C and
300 psig) were evaluated at each temperature. In this way, a total set of 9 conditions that would

98
allow having enough data to conduct a kinetic study were collected. The pressure of the tests
(300 psig) was the same for all the evaluated conditions.

In order to compare the reactivity at each condition, the conversion of hydrocarbons that boil
above 560 ºC defined in Eq. 5.7 was used:

% 𝑤𝑡. 𝐻𝐶 (560 °C +) 𝐹𝑒𝑒𝑑 − % 𝑤𝑡. 𝐻𝐶 (560 °C +) 𝑃𝑟𝑜𝑑𝑢𝑐𝑡 Eq. 5.7


Conversion = ∗ 100%
%𝑤𝑡. 𝐻𝐶 (560 °C +) 𝐹𝑒𝑒𝑑

5.3.5. Kinetic Modeling

Lump kinetic modeling including asphaltenes generation during CSC process was done
following the procedure discussed in Chapter 4 for thermal cracking of the same feedstock under
similar levels of steam. A schematic of the reaction pathways considered for kinetic modeling is
shown in Figure 5.3. As suggested by the authors, the following sequence on the activation
energies was imposed on the mathematical modeling in order to incorporate chemical
considerations within the complex mathematical algorithm (referred as case 3 in Chapter 4):

Sequence: activation energies of thermal cracking of heavier lumps are lower than lighter
ones. Furthermore, a sequence was imposed on the activation energies in such a way that the
activation energy of conversion of a lump to lighter products is higher than to heavier products.
In addition, constraints on the kinetic constant of gas production were imposed in such a way
that gas production proceeds more readily from heavier lumps (𝑘7 > 𝑘12 > 𝑘16).

Figure 5.3. Reaction pathways for kinetic modeling of DAO catalytic steam cracking.

99
5.4. Results and Discussion

5.4.1. Catalytic Feedstock Preparation

In order to collect the amount of feedstock required to complete the experimental plan, three
different batches of DAO+CAT were prepared in the skid preparation unit. For each batch,
catalyst concentration was measured by ICP analysis to guarantee that an adequate concentration
of particles close to the set target values has been incorporated into the feedstock.

As presented in Table 5.1, where the absolute error (AE) of each metal concentration with
respect to the target value is reported, the majority of the actual measured catalyst concentrations
were close to the expected target values with an average relative error of 6.7 %. This proves the
reliability and reproducibility of the results obtained in the skid unit when operated different
times for the incorporation of UDC particles.

Table 5.1. Catalyst concentration in DAO+CAT

RE - Ni K RE - K
Ni [ppm] [%] [ppm] [%]
Target 300 N/A 400 N/A
Batch 1 282 6.0 402 0.5
Batch 2 312 4.0 433 8.25
Batch 3 254 15.3 377 5.8
*RE: Relative Error
After validation of accurate catalyst concentrations on the different batches, all prepared
materials were mixed together in a mixing tank and the resulting blend was fully characterized.
Table 5.2 presents a summary of the characterization of the feedstock once it was processed in
the skid preparation unit as well as the characterization of unprocessed DAO. As illustrated in
the table, incorporation of the catalyst particles has a slight effect on the bulk properties of the
feedstock. The main reason for the change in properties such as API gravity and Viscosity is
related to a certain amount of the gasoline used as diluent that remains in the feedstock after
separation; SimDist results indicates it is in the order of 2.2%. Increase in MCR (after
subtracting the theoretical amount of catalyst particles), and marginal increase of asphaltenes and
decrease of resins is related to the thermal treatment at 370 °C required to decompose the

100
precursor salts during catalyst incorporation, that even though it was very short in time, had
certain effect on the bulk physical properties.

Table 5.2. DAO feedstock properties before and after catalyst incorporation

DAO+CAT
Analysis DAO
(Ni/K)
API Gravity 6.0 6.7
Microcarbon Residue [%wt] 12.95 14.41
Viscosity @60°C [cP] 5339 3436
Viscosity @100°C [cP] 300 N/A
Saturates [%wt] 4.0 3.6
Aromatics [%wt] 75.0 76.3
SARA
Resins [%wt] 20.2 18.7
Asphaltenes-C5 [%wt] 0.7 1.4
Naphtha (28 - 190 °C) [%wt] 0.2 2.4
Kerosene (190 - 260 °C) [%wt] 2.7 2.6
Diesel (260 - 343 °C) [%wt] 5.3 5.1
Cut Yields
LVGO (343 - 453 °C) [%wt] 7.0 6.6
HVGO (453 - 560 °C) [%wt] 17.2 17.4
Vacuum Residue (560°C+) [%wt] 67.7 66.0

5.4.2. Evaluation of Water Splitting

According to the water dissociation (Eq. 5.2) and oxidation/reforming reactions that are
theoretically taking place during CSC processing (Eq. 5.4), the catalyst would promote formation
of hydrogen free radicals and also CO2 and H2 when compared to a merely thermal non catalytic
reaction used as the bench mark. In addition, when using isotopic water (H2O18), water splitting
can be corroborated by detecting transfer of O18 from the isotopic water to form either CO2
(m/e=46), where only one O18 is present in the newly formed CO2 molecule, or CO2 (m/e=48),
where two O18 are present in the CO2 molecule. In this way, an overall increase in the H2 and
CO2 signals should be observed for the catalytic process as compared to the thermal cracking
reaction if regular demineralized water is used. In the same way, an increase on the relative
signal of CO2 (m/e=46) and CO2 (m/e=48) with respect to normal CO2 (m/e=44) should be
observed when isotopic water is used. A schematic of water splitting and O18 transference from
isotopic water to CO2 is presented in Figure 5.4.

101
Figure 5.4. Mechanism of H2O splitting by K+

Since the experimental setup is pressurized with industrial N2 before reaching the actual
temperature of reaction, in the beginning of the run the concentration of reaction gases was very
low, increasing gradually as they were produced from the CSC process. This increase in
concentration during the experiment was reflected in an increase in the intensity of the signals
measured by the QMS over time. As illustrated in Figure 5.5, where the QMS signals for normal
CO2 (m/e=44), and labelled CO2 (m/e=46 and 48) that are directly related to the concentration of
each species in the gas phase inside the unit, no change of any of the relative concentration of
any of the three species with respect to the others was observed at the initial temperature of 423
°C, using regular demineralized water. As mentioned previously, the increase of all signals at the
same rate is due to the decrease of the N2 concentration in the unit. Similarly, no relative change
was observed when the catalytic feedstock was incorporated either with regular water or with
isotopic water at the same reaction conditions. All signals increase at the same rate up to 8 h of
reaction due to the increasing total concentration of reaction gases and decreasing concentration
of N2 in the unit. When the temperature was increased to 430 °C a noticeable increase in the CO2
(m/e=46) signal was observed. In the same way, the trend of the signal corresponding to CO2
(m/e=48) increased as well at a temperature of 437 °C. When plotting the ratios of CO2 (m/e=46)
and CO2 (m/e=48) with respect to CO2 (m/e=44) as depicted in Figure 5.5, the incorporation of
isotopic oxygen atoms in the formed CO2 molecules can be easily observed due to the sharp
slope change of the plot. This allow us to affirm that for CSC of DAO using a Ni/K UDC
formulation, water splitting starts occurring at temperatures above 430 °C. This is in agreement

102
with previous results reported by Fathi et al.5 for catalytic steam cracking of ALVR where
similar temperature levels were needed for water splitting to take place.

Figure 5.5. Mass spectrometry analysis for catalytic steam cracking of DAO using O18 isotopic
water.

The H2 and CO2 molar concentration calculated from the GC analysis carried out at each
condition allows one to verify the performance of the catalyst in a quantitative manner under the

103
conditions of our tests. As seen in Figure 5.6, there is a constant increase in the hydrogen
concentration in the gas phase as the reaction proceeds compared to thermal cracking. For CO2,
its concentration increases from catalytic conditions C2 (423 °C and LHSV 2 h-1) to C4 (440 °C
and LHSV 3 h-1). This agrees with the QMS results where transference of isotopic oxygen O18
from the isotopic water splitting was observed at T 430. This supports QMS results in a
quantitative manner to affirm that the Ni/K UDC is indeed carrying out water splitting providing
hydrogen to the reaction media. The main finding of these results is that a very specific window
of operating conditions are needed in such a way that the UDC can actively dissociate water
under the CSC processing of DAO.

Figure 5.6. H2 and CO2 comparison for thermal tracking and catalytic steam cracking. Ci:
catalytic conditions.

5.4.3. Reactivity Analysis

Figure 5.7 shows the product distributions vs. conversion for the CSC conditions explored in
this work, compared with the thermal cracking baseline reported in Chapter 4. As observed,
product distribution for CSC followed the same trends as in thermal cracking, with significant
decrease of the heavier cut DAO (560 ºC+), slight increase of distillates and naphtha, and slight
increase of the LVGO cut with respect to the feedstock. The main difference is that severity
ranges attained in CSC are above the 43.3 % conversion level achieved for thermal cracking at

104
the evaluated operating conditions, reaching a maximum conversion of 51.4%. Additionally, a
considerable yield of around 15% of asphaltenes was achieved at the most severe condition. This
behavior suggests that the presence of the UDC does not perform any hydrocarbon cracking
function that would lead to an increase of conversion; this parameter is merely controlled by
thermal cracking reactions. The function of the UDC is solely to provide hydrogen for
hydrocarbon radicals saturation and an eventual enhancement of product quality. In this way, the
attainable conversion level is controlled in the same way as for thermal cracking and the
increased conversions obtained were a result of the severe operating conditions needed to
provide the adequate conditions for the UDC to catalytically dissociate water (T > 430 °C).

Figure 5.7. Product distribution for thermal cracking and catalytic steam cracking.

Table 5.3 summarizes the reactivity results for all the conditions evaluated in this work as
well as one thermal cracking condition at a conversion level of 43.3 % reported in Chapter 4.
The range of conversions of the 560 ºC+ hydrocarbons achieved for the CSC reactions at the
temperature and LHSV combinations explored in this work spanned from 37.1 % to 51.4 %. The

105
limit for further increase on thermal cracking conversion beyond 43.3% is the stability of the
asphaltenes, which was reported to be at the commonly accepted limit of P-value=1.15.
Operation below a P-value of 1.15 is not recommended since asphaltenes deposition and coke
formation inside the reactor and along the connecting lines would start to occur. A higher P-
value for the CSC route would allow the increase of the severity of the process, producing in that
way more valuable products. When carrying out the P-value analysis, a considerable presence of
solid particles was observed in the samples, which made difficult its determination. These
particles are at least in part agglomerates of catalyst particles, which would be removed by water
washing and desalting in an industrial operation. In order to overcome this issue, removal of
CHCl3 insoluble material was carried out using vacuum filtration in order to clean the sample
with further removal of the solvent via rotoevaporation and subsequent removal of remaining
solvent traces by placing the sample in a vacuum oven for 12 h at 50 ºC. Due to the limited
amount of cleaned sample, only the P-value at the limit value of 1.15 was done, as reported in
Figure 5.8. When making a direct comparison of this parameter for thermal and catalytic
processes at similar conversion levels (thermal, C8 and C11), no improvement on the stability of
the product was observed. At the higher levels of conversion reached under CSC processing, the
stability of the liquid product was compromised, with P-values within the range of 1.00 to 1.15.
These results contrast with previous findings, where the stability of the product was improved,
allowing the process to reach higher levels of conversion, and consequently producing more
valuable lighter hydrocarbons.4,5 However, the feedstock used in the previous reports was virgin
vacuum residue from Arabian light oil, substantially different from the partially cracked DAO
feedstock of the present tests. Possibly, the catalytic route has a better performance for
asphaltene containing hydrocarbons, where the nature of this heavy and highly reactive fraction
benefits CSC processing. In the same way, other indicators of the extent of the condensation
reactions taking place where the catalyst could act as an inhibitor, such as asphaltenes content, is
within similar levels at isoconversion. Also, no reduction in gas yield as a function of conversion
was observed for the explored range of operating conditions. Undisclosed results suggest that
UDC can act as seeds for coke formation and asphaltenes agglomeration if not operated at the
appropriate conditions for vacuum residue processing. Nonetheless, no previous works in the
area of CSC of DAO using UD catalyst have been found to confirm the same behavior with this
feedstock. Additionally, even though the UDC is not promoting a significant enhancement of the

106
bulk physical properties of the products, the chemistry of CSC is taking place and does not
promote a detrimental effect.

Table 5.3. Summary of experimental conditions and whole liquid product characterization for
catalytic steam cracking of DAO
Experiment
Analysis
Thermal* C5 C6 C7 C8 C9 C10 C11 C12 C13
Temperature [°C] 423 435 435 435 440 440 440 445 445 445
-1
LHSV [h ] 2 4 3.5 3 4.5 4 3.5 5.5 5 4.5
Pressure [psig] 300 300 300 300 300 300 300 300 300 300
Water content [%wt] 5 5 5 5 5 5 5 5 5 5
HC Mass Balance [%] 98.9 98.0 98.1 96.9 96.0 98.2 97.6 93.7 96.2 99.9
Conversion HC (560 °C+) [%] 43.3 37.1 46.7 48.3 42.7 46.2 51.4 44.2 44.6 46.1
P-value 1.15 Pv>1.15 1<Pv<1.15 1<Pv<1.15 1.15 1<Pv<1.15 1<Pv<1.15 1.15 1<Pv<1.15 1<Pv<1.15
Saturates [%wt] 11.9 11.8 13.2 13.4 13.2 12.8 14.4 14.1 14.3 13.9
Aromatics [%wt] 62.6 67.4 64.7 63.6 65.3 64.5 61.5 59.3 63.1 62.7
SARA
Resins [%wt] 11.4 8.3 7.2 7.2 7.2 7.3 7.1 13.1 7.2 5.6
Asphaltenes-C5 [%wt] 14.1 12.5 14.0 15.9 14.3 15.4 17.0 13.6 15.4 17.8
API Gravity 8.1 9.1 9.4 9.6 8.6 10.0 10.1 9.0 9.5 10.3
MCR [%wt] 16.54 17.40 18.02 18.37 18.38 18.15 18.57 17.20 17.91 18.75
Viscosity @25°C [cP] 2802 1981 1499 1383 1499 1010 832 1034 1036 1096
Solids [%wt] 0.23 0.17 0.35 0.43 0.56 0.35 0.61 0.35 0.48 0.51
*Results for thermal cracking reported in Chapter 4

Pv= 1.0 Pv= 1.0 Pv = 1.15


(No cleaning) (Cleaned) (Cleaned)
440 ºC, LHSV 4.5 h-1

440 ºC, LHSV 4.0 h-1

440 ºC, LHSV 3.5 h-1

Figure 5.8. Microscope images of P-value for products gathered at 440 ºC via catalytic steam
cracking.

107
Based on the discussed results, the evidence indicates that the functionality of the UDC
producing hydrogen from water splitting, attributed to the K particles, is working appropriately.
However, its hydrogenating capabilities to saturate the hydrocarbon free radicals minimizing
condensation reactions, which are attributed to the nickel particles, are not taking place in an
extent that could bring significant enhancement on product quality. Presumably, the submicronic
particles are not in an active state or are rapidly poisoned by the formation of insoluble
compounds due to the high severity needed for the process. As presented in Figure 5.9, where the
content of CHCl3 insoluble materials as a function of conversion are compared for both CSC and
TC, a marked difference on the trends is observed for both processes. At similar conversion
levels, the amounts of solids present in the liquid product for the CSC reactions are more than
twice as high as those produced via TC. This is due to at least two facts. First, the catalyst
particles per se increase the solid content by acting as a solid seeds, getting encapsulated by polar
compounds (resins and asphaltenes), as it can be derived from the increased microcarbon content
of the feed with catalyst, which passes from 12.95% to 14.57%. In total this contribution could
be sufficient for the difference with respect to thermal cracking. Second, the fact that
condensation reactions usually have high activation energies that result in their promotion at high
temperatures. In this order of ideas, the experimental conditions needed for catalytic functions to
take place with regards to water splitting have a counterproductive effect on the overall
performance of the process. Another factor to take into account is the characteristics of this
particular DAO that has a high content of aromatic hydrocarbons and, as explained with more
detail elsewhere in the thermal cracking analysis presented in Chapter 4, it comes industrially
from a sequential thermal cracking/solvent deasphalting recycling loop that makes it much more
prone to produce insoluble compounds and asphaltenes as compared to a virgin DAO. Further
investigations of catalyst formulations to address the challenges of this particular feedstock
should be tasked. An interesting approach would be to develop a catalyst formulation able to
dissociate water and hydrogenate at lower temperatures in a range where thermal cracking still
occurs, but condensation reactions are minimized.

Lastly, it is perhaps interesting to notice in Figure 5.9 that the trend of solid content
increases with severity and seems to turn exponential (expectably for coke formation) whereas
the catalysts test does not seem to provide a definitive trend perhaps indicative of a rather
invariable presence of solids less related to the reaction process. This could be related to the

108
stability of the asphaltenes in the liquid product samples. Whereas all the thermal cracking
products were above the P-value stability limit of 1.15, most of the catalytic products were below
1.15, which indicates that those samples are more prone to precipitate. This increased instability
might have caused agglomeration of solid particles along with some asphaltenes that can
precipitate. Therefore, the samples could be more heterogeneous, which can increase the
variability for the solids analysis.

Figure 5.9. Comparison of solid contents on the whole liquid products from catalytic steam
cracking and thermal cracking.

Comparing the behavior of important properties related to the quality of the whole liquid
product, linear trends were obtained for the plot of logarithm of the measured viscosity as a
function of conversion with coefficients of correlation of 0.9689 and 0.9082 as illustrated in
Figure 5.10, which are fairly good for experimental data collected at a pilot plant level. When
comparing the slopes of the linear correlation obtained for each property as a function of
conversion and with values reported for thermal cracking (viscosity: 0.1046 and MCR=0.0914),
we can see that the slope obtained for the viscosity plot under CSC conditions (slope= -0.1017)
is practically the same as that of thermal cracking. This corroborates that for DAO, the extent of
reduction of viscosity is driven mainly by conversion regardless of the type of properties or
feedstock characteristics. Conversely, the slope obtained for the MCR plot (0.0813) is 11% lower

109
than the one obtained for thermal cracking, which would be an indication of a superiority of CSC
to form products less prone to coking.

Figure 5.10. MCR and viscosity profile vs. conversion of HC (560 °C+).

5.4.4. Kinetic Study

Kinetic modeling of catalytic steam cracking of deasphalted vacuum residue including


asphaltenes generation was successfully conducted following the procedure developed by
Cabrales-Navarro and Pereira-Almao, presented in Chapter 4. A very good fitting between the
experimental and modeled data was obtained as illustrated in Figure 5.11, where a coefficient of
correlation of 0.9841 was achieved for the plot of modeled compositions vs. experimental
results. A global Average Absolute Error (GAAE) of 4.23 % was obtained, which is lower than
the modeling errors obtained for thermal cracking. Considering the modeling errors for each
lump as presented in Figure 5.12, the highest absolute errors were obtained for the lighter lumps
(naphtha and gas) and for the heavier ones DAO (560 °C+) and Asp-C5, which also agrees with
the thermal cracking modeling.

110
0.30

0.25
Model Composition [weight fraction]
y = 0.9572x + 0.0054
R² = 0.9841
0.20
DAO (560 °C+)
0.15 Asp-C5
HVGO
LVGO
0.10 Diesel
Kerosene
0.05 Naphtha
Gas

0.00
0.00 0.05 0.10 0.15 0.20 0.25 0.30
Experimental Composition [weight fraction]

Figure 5.11. Predicted model composition vs. experimental composition for kinetic modeling of
catalytic steam cracking of DAO.

12

10

8
AAE [%]

Figure 5.12. Average absolute error by lump for kinetic modeling of catalytic steam cracking of
DAO.

111
Table 5.4 presents the kinetic parameters for kinetic modeling of CSC as well as kinetic
constants at 435 °C extrapolated for the kinetic modeling of thermal cracking for comparison
purposes. A very good fitting of the Arrhenius equation was obtained with correlation
coefficients close to 1 for all reaction pathways. The global activation energy of conversion of
the DAO (560 °C+), HVGO and LVGO lumps were 175, 178 and 188 kJ/mol. Also, activation
energies of a single lump have an increasing trend for lighter compared to heavier ones. For
example, for DAO (560 °C+) activation energy of reaction 2 is lower than for reaction 3, and
activation energy for reaction 3 is lower than for reaction 4, and so on. Production of the gas
lump proceeds more readily from heavier lumps with k7 > k12 > k16. Under these
circumstances, all the constraints implemented in the mathematical modeling were satisfied. The
reactivity of the DAO (560 °C) lump at the explored operating conditions for CSC showed a
marked selectivity towards production of HVGO, followed by Asp-C5 and gases. The kinetic
constants towards these three lumps are 2 orders of magnitude higher for HVGO and Asp-C5
production and one order of magnitude higher to gas production. This contrasts with the
extrapolated values for thermal cracking without UDC where the selectivity of DAO (560 °C) is
more distributed toward all the other lumps. Very importantly, the kinetic constant for
production of Asp-C5 is lower for CSC. An advantageous catalytic route would, besides
increasing the product quality, inhibit this reaction step while favoring the others, producing in
this way higher amount of desirable products, preferably within the VGO range, while keeping
the produced asphaltenes content as low as possible. Additionally, HVGO and LVGO are much
more reactive towards the production of lighter compounds under CSC processing conditions. In
this order of ideas, taking into consideration that the mechanism of conversion under CSC is
driven by thermal cracking reactions and that there are evident differences between the kinetic
parameters obtained from the extrapolated values of the thermal cracking model developed at
temperatures up to 423°C, it can be implied that there is a change of the reaction mechanism at
high temperatures (T> 430 °C) combined with high LHSV (>3 h-1).

112
Table 5.4. Estimated kinetic parameters for upgrading of DAO via catalytic steam cracking

TC Catalytic Steam Cracking (CSC)


Kinetic Constants [h-1] Ea Ln A r2
Reaction
435* 435 440 445
[kJ/mol] [A in h-1 ]
[°C] [°C] [°C] [°C]

r1 DAO (560 C+) → Asp -C5 1.2435 1.0832 1.4015 1.6890 188 32.00 0.9923

r2 DAO (560 C+) → HVGO 0.7690 2.2328 2.8828 3.3177 168 29.28 0.9740

r3 DAO (560 C+) → LVGO 0.7536 0.0169 0.0207 0.0253 170 24.79 1.0000

r4 DAO (560 C+) → Diesel 0.4157 0.0177 0.0217 0.0266 172 25.25 1.0000

r5 DAO (560 C+) → Kerosene 0.1418 0.0153 0.0188 0.0231 175 25.52 1.0000

r6 DAO (560 C+) → Naphtha 0.3490 0.0144 0.0178 0.0219 177 25.89 1.0000

r7 DAO (560 C+) → Gas 0.4055 0.2556 0.3190 0.3913 180 29.23 0.9996

r8 HVGO → LVGO 0.6229 2.3934 2.9410 3.6028 173 30.24 1.0000

r9 HVGO → Diesel 0.1905 0.7436 0.9195 1.1320 178 29.89 1.0000

r10 HVGO → Kerosene 0.3337 0.6494 0.8054 0.9967 181 30.33 1.0000

r11 HVGO → Naphtha 0.1895 0.4751 0.5915 0.7350 185 30.60 1.0000

r12 HVGO → Gas 0.2626 0.2205 0.2831 0.3621 210 34.11 1.0000

r13 LVGO → Diesel 0.3154 0.5560 0.6875 0.8480 179 29.73 1.0000

r14 LVGO → Kerosene 0.6165 0.4570 0.5685 0.7044 183 30.29 1.0000

r15 LVGO → Naphtha 0.3103 0.9195 1.1492 1.4313 187 31.69 1.0000

r16 LVGO → Gas 0.1354 0.1599 0.2128 0.2818 240 38.89 1.0000

AAE [%] 4.57 3.58 4.52 GAAE[%] 4.23


kglobal 3.6359 4.6823 5.4948

*kinetic constants extrapolated from kinetic parameters reported elsewhere (Chapter 4)


5.5. Conclusions

It was found that the selection of the operating conditions is of paramount importance for
catalytic steam cracking of deasphalted vacuum residue to have catalytic activity. The
capabilities of the Ni/K ultradispersed catalyst as a water dissociation agent at high reaction
temperatures (T>430ºC) were evidenced. However, these severe conditions combined with the
characteristics of the feedstock limit the correct performance of hydrogenating functions due to
the increased production of coke. Nonetheless, no significant enhancement of product quality
was observed with the current formulation at the same conversion levels obtained for thermal
cracking. Product distributions obtained followed the same trend as in thermal cracking. Thus,

113
UDC per se does not induce hydrocarbon cracking. In addition, as for thermal cracking, liquid
product stability seems to be the major constraint to reach increased conversions. Operation
beyond the 43% conversion of HC 560 °C+ would destabilize products from the presence of the
generated asphaltenes and would promote undesirable condensation reactions regardless of
processing with the current catalyst formulation. In this way, the impact that an optimized
catalyst formulation would have on product quality and consequently the added-value this might
add to current thermal cracking technologies motivates the development of new catalyst
formulations at more moderate conditions for this particular feedstock. From the kinetic point of
view, a change to the mechanism was observed under the studied high severity conditions. The
DAO (560 °C+) fraction is more selective toward production of HVGO compared to TC.
Moreover, HVGO and LVGO fractions were determined to be much more reactive under CSC
operating conditions.

5.6. Acknowledgement

The authors are grateful to the Natural Sciences and Engineering Research Council of
Canada (NSERC), Nexen-CNOOC Ltd, and Alberta Innovates-Energy and Environment
Solutions (AIEES) for the financial support provided through the NSERC/NEXEN/AIEES
Industrial Research Chair in Catalysis for Bitumen Upgrading. Also, the contribution of facilities
from the Canada Foundation for Innovation, the Institute for Sustainable Energy, Environment
and Economy, the Schulich School of Engineering and the Faculty of Science at the University
of Calgary are greatly appreciated.

5.7. References

(1) Pereira-Almao, P.; Marzin, R.; Zacarias, L.; Cordova, J.; Carrazza, J.; Marino, M. Steam
conversion process and catalyst. U.S. Patent No. US5885441, 1999.
(2) Pereira-Almao, P.; Flores, C.; Zbinden, H.; Guitian, J.; Solari, R. B.; Feintuch, H.; Gillis, D.
Aquaconversion technology offers added value to E. Venezuela synthetic crude oil production.
Oil Gas J. 2001, 99 (20), 79-85.
(3) Trujillo-Ferrer, G. Thermal and Catalytic Steam Reactivity Evaluation of Athabasca Vacuum
Gasoil. MSc Thesis, University of Calgary, Calgary, 2008.
(4) Fathi, M. M. Comparative Upgrading of Arab Light Vacuum Residuum via Aquaprocessing
and Thermal Cracking. PhD Thesis, University of Calgary, Calgary, 2011.
(5) Fathi, M. M.; Pereira-Almao, P. Catalytic Aquaprocessing of Arab Light Vacuum Residue
via Short Space Times. Energy Fuels 2011, 25 (11), 4867-4877.

114
(6) Pereira-Almao, P.; Trujillo-Ferrer, G.; Peluso, E.; Galarraga, C.; Sosa, C.; Algara, C. S.;
Lopez-Linares, F.; Carbognani-Ortega, L. A.; Zerpa-Reques, N. G. Systems and Methods for
Catalytic Steam Cracking of Non-Asphaltene Containing Heavy Hydrocarbons. U.S. Patent No.
US20130015100A1, 2013.
(7) Poteau, S.; Argillier, J.-F.; Langevin, D.; Pincet, F.; Perez, E. Influence of pH on Stability
and Dynamic Properties of Asphaltenes and Other Amphiphilic Molecules at the Oil−Water
Interface. Energy Fuels 2005, 19 (4), 1337-1341.
(8) Carbognani, L.; Lubkowitz, J.; Gonzalez, M. F.; Pereira-Almao, P. High Temperature
Simulated Distillation of Athabasca Vacuum Residue Fractions. Bimodal Distributions and
Evidence for Secondary “On-Column” Cracking of Heavy Hydrocarbons. Energy Fuels 2007, 21
(5), 2831-2839.
(9) Di Carlo, S.; Janis, B. Composition and visbreakability of petroleum residues. Chem. Eng.
Sci. 1992, 47 (9), 2695-2700.
(10) Carbognani Ortega, L.; Rogel, E.; Vien, J.; Ovalles, C.; Guzman, H.; Lopez-Linares, F.;
Pereira-Almao, P. Effect of Precipitating Conditions on Asphaltene Properties and Aggregation.
Energy Fuels 2015, 29 (6), 3664-3674.
(11) Carbognani, L.; Gonzalez, M. F.; Pereira-Almao, P. Characterization of Athabasca Vacuum
Residue and Its Visbroken Products. Stability and Fast Hydrocarbon Group-Type Distributions.
Energy Fuels 2007, 21 (3), 1631-1639.
(12) Galarraga, C. E.; Pereira-Almao, P. Hydrocracking of Athabasca Bitumen Using
Submicronic Multimetallic Catalysts at Near In-Reservoir Conditions. Energy Fuels 2010, 24
(4), 2383-2389.

115
Chapter 6: Comparative Upgrading of Deasphalted Vacuum Residue via Catalytic
Steam Cracking using Ni/Ce-based Ultradispersed and Fixed-bed Catalyst
Formulations

Fredy A. Cabrales-Navarro* and Pedro Pereira-Almao


Department of Chemical and Petroleum Engineering, Schulich School of Engineering,
University of Calgary, 2500 University Drive N.W., Calgary, AB T2N 1N4, Canada
* Corresponding author (facabral@ucalgary.ca)

6.1. Abstract

Upgrading of deasphalted oil (DAO) via Catalytic Steam Cracking (CSC) has been recently
studied as a feasible alternative to improve the performance of conventional thermal cracking for
implementation on field upgrading applications. In the present study, the reactivity of industrial
vacuum residue DAO via CSC at moderate temperatures was evaluated using Ni/Ce-based
ultradispersed (UDC) and fixed-bed catalyst formulations. Three different 700 ppm Ni/Ce UDC
formulations with Ni/Ce atomic ratios of 2.0, 1.5 and 1.0 were assessed at temperatures between
380-423 °C and Liquid Hourly Space Velocities (LHSV) of 0.25 and 2 h-1. A fixed-bed (FB)
catalyst formulation, in a reduced and an oxidized state, was evaluated for comparison purposes
at temperatures between 360-380 °C and a Weight Hourly Space velocity (WHSV) of 0.25 h-1.
Thermal blanks were conducted for all reactions. A research-scale pilot plant assembled with an
up-flow tubular reactor was used for the experimental evaluations at an operating pressure of 300
psig. Both type of catalysts showed water splitting and hydrogen production capabilities. Unlike
the UDC formulations, the fixed-bed catalyst showed hydrogenating capabilities leading to
enhancement of product quality, as well as increased conversions. The results suggest that the
reduction state of the catalyst active phase has an influence on reaction conversion, but not a
very significant effect on the product quality.

Keywords: Catalytic Steam Cracking, DAO, Ultradispersed Catalyst, Fixed-bed Catalyst

6.2. Introduction

The use of steam as the source of hydrogen for hydrogenation of heavy hydrocarbon
fractions has been seen as a feasible alternative for upgrading of heavy oils for the past two

116
decades. In this steam-based process the catalyst has two main roles, the first one is water
splitting to produce hydrogen free radicals (H•) and the second one is to promote the transfer of
those hydrogen radicals to the hydrocarbon (hydrogenation). Usually, catalysts for steam
conversion processing work at pressures ranging preferably from 100 to 300 psig, considerably
lower than those required for conventional hydroprocessing.1,2 In this way, the development of a
suitable catalytic steam cracking catalyst could bring significant economic benefits if it is able to
outperform a conventional hydroprocessing catalyst.
Catalysts are commonly classified as supported and unsupported. In the first case, the
number of active sites of the catalyst is maximized by dispersing it onto a support surface,
usually characterized by having a very large surface area per unit of mass, increasing in this way
the probability of the catalytic reaction to occur.3 An important advantage of the use of supported
catalysts is that there is no need for a separation stage to separate catalyst from the products. On
the other hand, a huge disadvantage of this type of catalyst is their deactivation over time, which
can be considerably high for extra-heavy feedstocks having high metals, nitrogen, sulphur
contents and also it is very likely to produce solid particles such as coke that may poison the
catalyst.2,3 Also, it is of paramount importance to consider that heavy feedstocks with the
presence of a considerable amount of asphaltene molecules, characterized by having large
diameters and high molecular weights4, may be very difficult to process with supported catalysts
since there could be serious mass transfer limitations due to the difficulty of asphaltene
molecules to mobilize through the micropores and mesopores of the structure.5,6 The second type
of catalyst, the unsupported type, as can be inferred from its name is not dispersed onto a solid
surface or support. In this case, until recently, lower active sites surface area per unit of mass of
catalyst were obtained and a separation process to remove the catalyst from the product was
required, which can be very costly, limiting in this way their applicability. However, there are no
significant mass transfer limitation issues, which makes them attractive for upgrading of heavy
hydrocarbons. Most importantly, with the development of nanotechnology in the last decade,
unsupported catalysts have had major improvements since it has been proven that going into the
nano-scale catalyst particle sizes (<100 nm), the surface area to volume ratio, or the surface area
of active sites per unit of mass of the catalyst, is considerably increased up to values that enable
the unsupported catalyst to outperform the conventional supported catalyst.7

117
During the 80s, Duprez et al.8-10 conducted extensive research work related to reactions of
aromatic hydrocarbons with steam, mainly focusing on the steam reforming reactions that can be
somehow comparable to the catalytic steam cracking reactions. In one work8, the steam
conversion of monoalkyl and dialkyl benzenes on a Rh catalyst with different supports
(Rh/Al2O3, Rh/SiO2, Rh/TiO2) was analyzed, classifying the products depending on their
producing reaction to: dealkylation, dehydrogenation, and degradation. Benzene, which comes
mainly from the dealkylation reaction, was found to be the most abundant product for the
reaction using Rh/Al2O3. The selectivity of the catalyst towards this component decreased as the
conversion increased, while the conversion towards toluene increased due to increase of C-C
splitting in the middle or end of the side chain. For the dehydrogenation products, a quick
decrease of the selectivity with conversion was found. Unlike the dehydrogenation reactions, the
support was found to have a strong effect on the selectivity of the catalysts towards benzene,
with differences of more than 10% on the selectivity. TiO2 was found to be the best support for
this type of reactions, followed by Al2O3, and SiO2. Continuing with the research in this area,
Duprez et al.9 extended the evaluation to Pt-based and Ni-based catalysts, where they found a
higher activity towards dehydrogenation than dealkylation for the Pt-based catalysts, and an
opposite behavior with nickel, this metal being much less selective in general terms than
platinum, irrespective of the feedstock used.
Based on an analogy with the reactivity observed in Aquathermolysis during huff-and-puff
steam injection in reservoirs, Clark et al.11 investigated the applicability of transition-metal
catalysts with the presence of steam for the upgrading of Canadian bitumen. A catalyst solution
of water with iron and ruthenium was employed, as they suggested those metals are known to
promote steam reforming and water gas shift reactions, looking forward to determining the
chemical effects, if any, of the catalyst on the bitumen. It was found that unlike water/ruthenium,
the use of water/iron catalysts showed a reasonably good performance in terms of the reduction
of insoluble compounds (mainly coke). The poor performance of the water/ruthenium was
attributed to the promoting of coupling reactions of reactive hydrocarbon fragments that
triggered the formation of solids, or also to the lack of capabilities of the catalyst to provide
hydrogen supply from the water, since tests carried out later on with the same catalyst showed a
fairly good performance if hydrogen was supplied to the reaction along with water.

118
Fumoto et al.12 investigated the upgradability of an atmospheric residue via catalytic steam
cracking using two iron-oxide catalysts, the first one being a hematite-based (α-Fe2O3) and the
second one a goethite-based catalyst (FeOx). It was found that FeOx yielded a larger amount of
distillates, lower amounts of gas, and no carbonaceous residue, compared to the α-Fe2O3 catalyst,
which would make it more suitable for upgrading applications. The author attributed this
difference in performance to the difference in the porous structure of both catalysts, having
practically a non-porous structure for α-Fe2O3, while for the FeOx a structure with pore diameters
within the 6 nm to 10 nm range was obtained, which finally reflects higher surface area and
active sites available for the surface reaction to occur. In the same study, in order to promote the
dissociation of water and consequently the production of active hydrogen and oxygen, ZrO2 was
used as the support structure for the FeOx, since ZrO2 is known to promote water dissociation.
The enhancing of the catalytic steam cracking reaction when adding ZrO2 was noted up to a
certain support concentration after which the catalytic activity of the catalyst decreases, a
phenomenon attributed to a change in the pore volume distribution of the catalyst, denominated
as pore plugging. In this way, it was concluded that it was possible to obtain valuable products
from residual hydrocarbons by using this type of ZrO2 supported FeOx catalyst. In 2005, Fumoto
et al.13 continued the study of this type of catalyst and carried out a regeneration test where it was
found that the sequence of reaction/regeneration was considerably deactivating the catalyst,
attributed to changes in the iron oxide phase. Using transmission electron microscopy, the
authors were able to see how the pores in the FeOx phase disappeared and also the ZrO2 support
peeled off from the FeOx structure with the increase of the number of sequences of
reaction/regeneration stages. This was caused by the phase transfer among the hematite and
magnetite which induced shrinkage and expansion in the FeOx phase. In order to avoid the
deactivation of the catalyst, a new formulation including alumina was used, which is stable at the
reaction/regeneration conditions. It was found that the addition of alumina in the catalyst
formulation enhanced its performance and durability even when several sequences of
reaction/regeneration were used. In a subsequent work14, it was found that the addition of
alumina also promoted the cracking reaction of the heavy hydrocarbon fractions.
Junaid et al.15 approached the catalytic steam cracking reaction in a different way, proposing
an integrated extraction-upgrading method. They investigated the upgradability of bitumen using
natural zeolites (clinoptilotites and chabazites) as catalysts, with particle sizes below 44 μm,

119
mixed with water, being able to demonstrate the enhancement of the cracking reactions
compared to uncatalyzed reactions in terms of product quality and reduced production of gas and
insoluble materials.
Pinilla et al.16 recently studied the catalytic steam cracking reaction of anthracene over a
nickel based catalyst, both undoped and doped with either sodium, calcium, or potassium,
supported on mesoporous alumina (γ-Al2O3). They used anthracene as a model molecule since
polycyclic aromatic hydrocarbons like this compound represent a large fraction of the heavy oil
components including asphaltenes, gaining in this way valuable understanding of the complex
steam-based reactions of heavy hydrocarbon fractions. The targeting of these catalytic reactions
is to increase the rupture of the internal rings of the aromatic molecules, instead of the external
ones, increasing in this way the selectivity towards intermediate products with higher amounts of
monocyclic components, but keeping a low consumption of hydrogen in the reaction. According
to the results of the study, a slight difference in terms of conversion between the use of undoped
Ni/Al2O3 compared to the catalyst doped with the alkali metals (NiNa/Al2O3, NiCa/Al2O3,
NiK/Al2O3) was found. However, an increase on selectivity to liquid products using the catalyst
doped with K and Na was found, which suggests effective ruptures of the internal bonds of the
anthracene were taking place. Based on this finding, Pinilla et al.17 decided to further study the
catalytic steam cracking of anthracene on NiK/Al2O3 catalyst, obtaining more detailed
information about the pathways of the reaction via the internal or external rings of the molecule.
Pereira-Almao and co-workers have been working for several years on the development of
catalysts for catalytic steam cracking reactions of heavy hydrocarbon fractions. In 1997, they
investigated the use of a combination of one metal from Group VIII of the periodic table (Fe, Co,
Ni, etc) and one alkali metal, preferably potassium or calcium, supported on a mesoporous
aluminosilicate, or as a water-in-oil emulsion, as feasible catalysts for catalytic steam cracking.18
They found very good results in terms of conversion and products quality. Based on their
findings, they continued the investigation using the catalytic emulsion route and patented the
novel catalytic steam conversion process.1 More recently, the catalytic process has been
improved with the aid of nanotechnology, having now a bimetallic ultra-dispersed catalyst
obtained from decomposition of the water-in-oil bimetallic microemulsions, obtaining particle
sizes in the nanometers order of magnitude, which significantly improves the activity of the
catalyst. For example, Fathi and Pereira-Almao, studied the catalytic steam cracking of an

120
Arabian Light Vacuum Residue (ALVR) at LHSVs spanning 5-10 h-1 using the developed Ni/K
ultradispersed catalyst formulation, and it was feasible to dissociate water and to provide
hydrogen for the process to improve product quality.19
A very interesting plot presenting the different transformations of the main molecular groups
present in hydrocarbons is shown in Figure 6.1 for the upgrading of a vacuum residue (VR) and
vacuum gas oil (VGO) via conventional thermal cracking (visbreaking), and using catalytic
steam cracking as presented by Pereira-Almao et al.20 It is seen how catalytic steam cracking
outperforms thermal cracking on the production of lighter fractions from the aromatics and resins
fractions, and very advantageously decreases the transformation of resins to asphaltenes, as well
as the generation of more asphaltene molecules. Resins help to stabilize asphaltene molecules,
thus, an increase on the production of asphaltenes from resins would really threaten the stability
of the final product. A key factor to highlight, at this point, is the possibility of a partial
upgrading of asphaltenes to light products, which is not present in conventional thermal
cracking.

Figure 6.1. Transformation of molecular groups for thermal cracking and catalytic steam
cracking of vacuum residue and vacuum gas oil.20

Recently, field upgrading of bitumen to produce partially upgraded bitumen that meets
Canadian transportability conditions (API: 20; viscosity 350 cSt at carrier reference temperature;
olefins < 1% wt 1-decene equivalent), has been seen as an alternative to maximize the economic
potential of bitumen commercialization. Solvent Deasphalting (SDA), where a paraffinic solvent
(nC5, nC6, etc.) is used to remove the heavy asphaltene molecules responsible to a large extent
for the high viscosity of bitumen, along with thermal cracking of the deasphalted oil fraction, are

121
among the most prominent technologies for field upgrading applications.21,22 Catalytic steam
cracking might play a very important role in the development of field upgrading technologies to
improve product quality and liquid yields upon the performance of conventional thermal
cracking of DAO. This directed catalytic steam cracking research towards the study of non-
asphaltene containing hydrocarbons (i.e. DAO, VGO, etc.), and possible upgrading schemes
involving this technology have already been patented.20 Trujillo-Ferrer23 applied catalytic steam
cracking for the upgrading of Athabasca vacuum gas oil (VGO) using water-in-oil emulsions of
Ni/K catalysts and satisfactory results in terms of microcarbon (MCR) reduction were obtained.
Additionally, Garcia-Hubner24 assessed the applicability of performing catalytic steam cracking
directly on a pentane (nC5) deasphalted bitumen on a fixed-bed reactor. In his work, the
capabilities of water dissociation and hydrocarbon cracking and hydrogenation were evidenced
when processing a heavy feedstock (DAO) in a fixed-bed reactor. The same feedstock used in
the present work was previously used to evaluate catalytic steam cracking processing of a Ni/K
UDC formulation as presented in Chapter 5. As indicated by the authors, high temperatures
(above 430 °C) were needed in order to promote a catalytic effect on water dissociation, which
directed their work towards the evaluation of high temperature/low residence time operating
conditions as represented in Figure 6.2. However, even though water splitting and hydrogen
production were evidenced, no significant hydrogenation leading to eventual enhancement on the
bulk product quality was observed. This phenomenon was attributed to triggering of the
condensation reactions, leading to high solid formation and rapid deactivation of the
ultradispersed particles. This phenomenon was ascribed to the high content of highly aromatic
partially-cracked hydrocarbons present in the evaluated feedstock. Those results motivated the
research in catalytic steam cracking of DAO towards the low temperature/high residence time
operating conditions, where hydrocarbon condensation is expected to be reduced.
Previous investigations conducted using Ni/Ce-based catalysts for the model Water Gas
Shift (WGS) reaction as well as heavy hydrocarbon feedstocks under catalytic steam cracking
processing using supported catalysts, highlighted the capabilities of ceria as a water dissociation
agent at relatively low temperature as presented by Garcia-Hubner.24 These findings encouraged
the design and preparation of a Ni/Ce-based catalyst in ultradispersed fashion following the pre-
established method of preparation previously used for Ni/K catalyst formulations to compare its
performance with a conventional supported catalyst of the same nature.

122
Figure 6.2. Effects of operating variables and condensation reactions when choosing a catalytic
steam cracking catalyst.

In this way, the objective of this work is to evaluate the capabilities of a Ni/Ce
ultradispersed catalyst for catalytic steam cracking of a partially cracked deasphalted vacuum
residue (DAO) at low temperatures. The effects of Ni/Ce atomic ratio in the catalyst formulation,
the effect of reaction temperature, as well as the effect of the oxidation state of the catalyst active
phase on the performance of the CSC process were assessed. In addition, all results were
compared with merely thermal cracking blanks. The oxidation state of the ultradispersed catalyst
is difficult to control, since very high temperatures would be needed to obtain completely
reduced metal particles in the microemulsion decomposition step. In this way, a fixed-bed
catalyst formulation had to be prepared and tested under similar conditions. Fixed-bed catalysts
can be reduced using hydrogen, which would facilitate the comparison with the results obtained
with the oxide-form UDC.

6.3. Experimental Methods

6.3.1. Catalytic Feedstock Preparation

An industrial deasphalted vacuum residue (DAO) with the characteristic of having a


considerable portion of pre-cracked hydrocarbons due to a sequence of thermal cracking/nC5-

123
SDA steps required for its preparation was used as feedstock and organic matrix for the
suspension of the Ni/Ce ultradispersed catalyst particles. Details about thermal reactivity and
characteristics of the industrial processing of this feedstock are provided elsewhere (Chapter 4).
In order to incorporate the UDC into the feedstock, a completely continuous skid preparation
unit was used. This bench-scale setup had been previously used for the incorporation of a Ni/K
UDC formulation for the same feedstock and proved to be very reliable for this kind of
procedures. A schematic of the apparatus is presented in Figure 6.3. For the preparation, DAO
was pumped at a temperature 140 °C to make it mobile. The first precursor solution
corresponding to the cerium nitrate heptahydrate (99 % Aldrich) was added at a rate that
corresponds to approximately 1 % wt. of the feedstock flowrate. The solution was passed
through a set of two static mixers connected in series to create a microemulsion. Next, the second
precursor solution that corresponds to nickel nitrate hexahydrate (99 % Aldrich) was added using
a similar volume percentage basis. Then, the stream was passed through a second set of static
mixers connected in series. In order to guarantee that water was in the liquid state during mixing,
the preparations were conducted at a pressure of 100 psig using nitrogen to pressurize the unit.
The amounts of precursor salts were calculated to target a total composition of 700 ppm of
catalyst in the feedstock at three different Ni/Ce atomic ratios (AR): 2.0, 1.5 and 1.0. In this way,
a total of three different formulations were assessed. The water-in-oil microemulsion was
submitted to a decomposition step at 370 °C to generate the corresponding Ce and Ni metal
oxide catalyst particles. The next step was to separate the water from the DAO matrix in a hot
separator operated at a temperature of 200 °C. The gaseous stream composed by water and gases
generated during the decomposition were passed through a condenser to liquefy water, which
was collected in a cold vessel. Non-condensable gases passed through a back pressure valve
(BPV) to control the operating pressure of the unit and then the gaseous stream was bubbled in
an aqueous potassium hydroxide solution trap (1N concentration) to sweeten it before directing it
to ventilation.

124
Figure 6.3. Ni/Ce ultradispersed catalyst preparation unit – skid.

6.3.2. Supported Catalyst Preparation

An in-house prepared supported catalyst was used for evaluating the effect of the oxidation
state of the catalyst active phase on the CSC reaction. For this preparation, 50 g of a Ni-Ce-
doped hydrotalcite (Ni/Ce-HDT) material, previously calcined at 450 C for 18 hours, were
blended with 50 g of kaolin (Sigma-Aldrich) and the mixture was ground until obtaining a
homogeneous textured powder without agglomerates. Then, 15 g of ludox AS40 (Sigma-
Aldrich) were added as a binder along with 50 g of water. The mixture was continuously
kneaded adding extra water as needed until getting a dough easy to extrudate to 1.7mm diameter
extrudates using an in-house prepared extruder based on a metallic syringe mounted on a press.
Next, extrudates were dried overnight at room conditions, and then, further dried and calcined in
an oven using the following conditions: i) heating to 100 C (for 6 h) at a heating ramp of 10
C/min, ii) heating to 550 °C (for 6 h) at the same heating ramp of 10 C/min. Finally, extrudates
were cooled down to room temperature. Afterwards, extrudates were manually cut to an average
length of 3 mm and placed in a sealed container for further use.

6.3.3. Reactivity Tests

A bench-scale experimental setup with capabilities of performing catalytic steam cracking


reactions in an open tubular or fixed-bed reactor available at the Catalysis for Bitumen
Upgrading group (CBU) at the University of Calgary was used to conduct all reactivity

125
experiments. Details about the description and operation of the unit were provided in Section 2.1.
As presented in Figure 6.4, a 103 mL empty-tube reactor was used when the Ni/Ce UDC
formulations were evaluated. This reactor consists of a ¾” O.D. – 0.049 W.T. with 50 cm length
stainless steel tubing and is assembled with a 1/8” diameter thermowell with 7 temperature
sensing points (5 points inside, 1 before the inlet and 1 after the outlet of the reactor) for
temperature monitoring. For the fixed-bed reactions, 53 g of catalyst extrudates were packed in
80% of the inner section of the reactor. The remaining 20% was filled with inert Silicon Carbide
(commercially known as carborundum) supplied by Panadyne. This material had to be cleaned to
remove impurities (mainly iron) following an in-house method24 based on stirring inside a
rotatory evaporator, washing batches of 50 g and 100 mL of 10% wt. HNO3 solution in deionized
water during 3 hours, followed by filtration and subsequent washing with excess of deionized
water. When no iron impurities leached from the material by checking with potassium
thiocyanate (turns red when iron is present), the material was dried up in a vacuum oven at 80 °C
for 24 hours. A fine metal mesh was added at the top and bottom of the reactor to hold the
catalyst/inert bed. Glass fiber wool was also added on top and bottom to prevent leaching of
catalyst solid particles from the reactor, as well as in between the inert material and the catalyst
extrudates to clearly separate the reaction from the inert zone.

Figure 6.4. Schematic of packed-bed and empty tube reactors. Reactor drawing from Swagelok
database.

126
6.3.4. Characterization Techniques

6.3.4.1. UDC- Characterization

 Metal Analysis: the metal content on the ultradispersed catalyst particles in the DAO
feedstock was obtained via Inductively Coupled Plasma (ICP), following the ASTM
D7691-05 norm, using a Thermoelectron Spectrometer model IRIS Intrepid II.
Microwave assisted digestion (ASTM D7455 norm) was used to prepare the sample
before the ICP analysis in order to decompose the organic feedstock matrix and solubilize
the metals presented in the catalysts as nitrates soluble in aqueous phase (a MARS6
equipment from CEM Corporation was used for this purpose).
 Particle Size: Nanoparticle Tracking Analysis (NTA) method developed by Rodriguez-
DeVecchis, et al.25 was used to determine the UDC particle diameter using a NanoSight
model NS300 equipment for two selected samples, which correspond to the lowest and
highest Ni/Ce atomic ratio.

6.3.4.2. Supported Catalyst Characterization

 XRD: powder XRD technique was used to analyze the crystalline structure of the solid
catalyst in a Rigaku ULTIMA III X-ray diffractometer with CuKα radiation. Scans were
done in the range of 3 to 90 2 , with a 0.02 step and a counting time of 0.2 /min at
40 kV and 44 mA. The sample was placed in a glass top-loaded sample holder, having a
0.5 mm depth cavity, for the analysis. The sample was distributed evenly on the cavity of
the sample holder with the help of a microscope slide. The crystalline domain sizes were
measured using the Scherrer equation26, as implemented in the commercial software
JADE27 (provided with the diffractometer) by fitting the experimental profile to a pseudo-
Voigt profile function, and then, calculating the full width at half maximum (FWHM) of
the peaks.
 Metal Content: metal contents of the sample, after calcination, were determined by
inductively coupled plasma (ICP) spectroscopy. The sample was added into proper
Teflon vessels and a mixture of hydrochloric and nitric acids was added to the samples.
The vessels were sealed and heated in a MARS6 microwave digestion apparatus
manufactured by CEM for a full cycle to extract the metals contained in the solids and
dissolve them in the acid solution. Then ICP analysis was carried out, on the obtained

127
solutions, in an Iris Intrepid II XDL spectrometer manufactured by Thermo Electron
Corporation.
 Surface Area: catalyst surface area was determined following ASTM D3663-84 method
in a Micrometrics TriStar 3000 analyzer.

6.3.4.3. Liquid Products

 Products Distribution: liquid product distribution by pseudo components based on


boiling point was calculated using High Temperature Simulated Distillation (SimDist)
according to an in-house modified version of the ASTM norm D-7169-05 developed by
Carbognani et. al.28
 Hydrocarbon Group-Type Distribution: Saturates-Aromatics-Resins-Asphaltenes
(SARA) hydrocarbon group type distribution was done using the method developed by
Carbognani et al.29, which consists on asphaltenes precipitation by pentane (nC5)
microdeasphalting followed by analysis of the deasphalted oil (maltenes) by Thin Liquid
Chromatography (TLC).
 Microcarbon content (MCR): MCR or Conradson carbon content was determined
following the muffle furnace method developed by Hassan et al.30
 Viscosity: viscosity was determined using a Brookfield viscometer model DV-II+ Pro
provided with a Brookfield TC-502 temperature bath. Measurements were conducted at
25 °C.
 API Gravity: API gravity was determined following the definition in ASTM D287-92
norm. For this, specific gravity was estimated using a digital densitometer Rudolph
Research Analytical model DDM2911 via dilution with toluene as described with more
detail elsewhere.31
 Solid contents: percentage of solids defined as CHCl3 (99% Aldrich) insoluble material
was quantified by vacuum filtration of a 1:50 (w/w) solution of sample/CHCl 3 in a 47
mm Millipore filter unit provided with 0.45 µm nylon membranes from Pall Corporation,
as explained by Galarraga et al.32 Mass of solids after filtration was corrected for the CSC
products with UDC by subtracting the theoretical amount of catalyst incorporated.

128
 Asphaltenes Stability: asphaltenes Stability Index (P-value) was used as a measure of the
state of peptization of the asphaltenes present in the sample following the titration
method developed by Di Carlo and Janis.33
6.3.4.4. Gas Products

 Gas Chromatography (GC): GC analysis was used to determine the distribution of


hydrocarbon and permanent gases of the produced gaseous stream using a SRI
Instruments chromatograph model 8610 C. Details of the specifications of the equipment
are provided in a previous work (Section 2.2.3.1).
 Mass Spectrometry (MS): in order to conduct a continuous monitoring of changes in the
gaseous stream composition, a Pfeiffer OmniStar Quadrople Mass Spectrometer (QMS)
was connected inline for all experimental runs. This allowed scanning atomic masses in
the gas stream ranging from 1 to 300 amu. Details about the equipment hardware and
functioning are provided elsewhere.19

6.3.5. Experimental Procedures for Catalyst Evaluations

6.3.5.1. Ni/Ce Ultradispersed Catalysts

Catalytic steam cracking of DAO using the three different prepared formulations (AR 2.0,
1.5 and 1.0) was conducted at Liquid Hourly Space Velocities (LHSV) between 0.25-2 h-1
defined at pumping conditions (140 °C and 300 psig), at reaction temperatures of 380 °C and 423
°C and at an operating pressure of 300 psig. In addition, a thermal blank (DAO without catalyst)
was also evaluated as a baseline to compare the actual effect of the catalyst incorporation on the
reaction performance. The reactivity unit was started up using DAO without catalyst at an LHSV
of 0.25 h-1 and 380 °C. When reaction conditions were achieved, the unit was allowed to stabilize
for a period of 12 h (3 times the space time of 4h) before collecting any mass balance. After this
step, two 8 h mass balances were collected. Then, feedstock was switched to DAO with Ni/Ce
UDC (AR:2.0), keeping the same experimental conditions and allowing the same 12 h
stabilization period followed by the two 8 h mass balances. The run was continued in the same
way for the remaining two DAO+CAT formulations (AR: 1.5 and 1.0). In a separate run, an
extra condition was tested using the remaining catalytic feedstock for the AR:1.5 formulation at
a temperature of 423 °C and a LHSV of 2.0 h-1. In this case, a stabilization period of 1.5 (3 times
the space time of 0.5 h) was allowed before collecting the two corresponding 1 hour mass

129
balances. This condition was found to be the conversion limit in terms of stability under thermal
cracking processing for the same feedstock in a previous work (Chapter 4). In this way, the
catalytic activity could be compared against the thermal baseline at a rather moderate
temperature (423 °C).

6.3.5.2. Ni/Ce Fixed-bed Catalysts

The first step in the catalytic steam cracking of DAO using the in-house prepared supported
catalyst was to carry out the catalyst reduction step. For this, 60 sccm of ultra-high purity
hydrogen were flowed through the reactor bed with simultaneous temperature increase ramp of
10 °C/min until reaching 500 °C at atmospheric pressure. Reduction at this temperature
continued for 6 h. After this step, hydrogen flow was stopped and replaced by industrial nitrogen
until passing three times the volume of the pilot plant to guarantee removal of remaining
hydrogen in the unit. Simultaneously, reactor temperature was decreased to 360 °C. Next, a
catalyst steam pre-treatment aiming to saturate the catalyst porous structure with water was
started at a flow of 0.3 g/min until an amount equivalent to 4 times the mass of catalyst was
injected. Then, pumping of DAO was started and reactor set-points were adjusted in order to
have a constant temperature profile at 360 °C along the reactor. Reaction was conducted at a
Weighty Hourly Space velocity (WHSV) of 0.25 h-1 and at an operating pressure of 300 psig.
Similar to the UDC evaluations, a stabilization period of 12 h was allowed for the reaction, after
which two 8 h mass balances were collected. Subsequently, reaction temperature was increased
to 380 °C and the same stabilization and mass balance period was allowed. In order to have an
idea about the extent of catalyst deactivation, the temperature was decreased back to 360 °C and
the first reaction was repeated. Aiming to evaluate the effect of the reduction step on the catalytic
performance, the catalytic bed was regenerated by oxidation at 430 °C using ultra high purity air
at a flow of 100 sccm with continuous GC analysis until no CO2 was present in the gaseous
stream. Then, a similar steam pre-treatment was conducted and the CSC reaction was carried out
at 380 °C and WHSV of 0.25 h-1 allowing in this case a stabilization period of 24 hours. As a
thermal baseline to compare with the performance of the catalytic process, the catalyst bed was
replaced by only carborundum and the same conditions were tested, those are: 360 °C – 0.25 h-1
and 380 °C – 0.25 h-1.

130
6.4. Results

6.4.1. Ultradispersed Catalyst Preparation

Catalyst concentrations in the DAO feedstock matrix measured by ICP analysis are
presented in Table 6.1. As observed, the actual nominal metal concentrations were close to the
target value with an average relative error of 12.3 % for the three prepared catalyst formulations.
On the other hand, the mean particle diameter measured using the NTA technique confirm the
incorporated particles are in a submicronic scale close to the nanometric range (<100 nm). No
significant change was observed on the reported particle diameters, which was expected since
this parameter is mainly controlled by the mixing method used during the preparation of the
precursor solutions microemulsions. In this case, all preparations were performed using the same
system based on static mixers connected in series as depicted in Figure 6.3.

Table 6.1. Catalyst concentration and particle diameter in DAO+CAT

Ni Ce Mean
Particle
Formulation Target Measured Target Measured
% RE* % RE* Diameter
(ppm) (ppm) (ppm) (ppm) [nm]
DAO-Ni/Ce
319 267 16.4 381 330 13.5 170±31
(AR = 2.0)
DAO-Ni/Ce
270 229 15.4 430 376 12.6 N/A
(AR = 1.5)
DAO-Ni/Ce
207 198 4.4 493 437 11.4 185±7
(AR = 1.0)
*RE: Relative Error

6.4.2. Supported Catalyst Characterization

Textural properties for the prepared supported catalyst, as well as the spent and regenerated
catalysts are presented in Table 6.2. Metal concentrations in the fresh catalyst are presented in
Table 6.3 . The adsorption-desorption isotherms and the pore width distribution on the supported
catalyst are presented in Figure 6.5. Based on the actual metal concentrations in Table 6.3, the
atomic ratio of Ni to Ce is 1.7, which is within the same range as those targeted for the UDC
preparations. As seen in Table 6.2, there is a small contribution of surface area due to
microporosity (about 11 %); however, the main area is not attributed to the micropores (meso-
macro), and thus, the heavy molecules can interact with the active sites exposed on this surface

131
area. Comparing the textural properties of the spent and regenerated catalyst with the fresh
catalyst, a considerable 49% of the BET surface area is lost after reaction is conducted, probably
caused by deposition of organic matter (mainly coke) inside the porous structure of the catalyst.
After calcination of the spent catalyst (catalyst regeneration), where a mass loss of 16.6%
(organic matter) was burned out, an increase of 56% of surface area (compared to the spent
catalyst) was obtained. The contribution of surface area due to microporosity in the spent catalyst
(about 15%) increases after reaction compared to the 11% for the fresh catalyst and decreases to
9% after regeneration.
Table 6.2. Textural properties of supported catalyst

Catalyst
Property
Fresh Spent Regenerated
BET Surface Area [m2/g] 81 41 64
t-Plot Micro Area [m2/g] 9 6 6
t-Plot External Area [m2/g] 72 35 59
BJH desorption Pore Volume [cm3/g] 0.302 0.154 0.316
BJH desorption Average Pore Width [Å] 124 119 84

Table 6.3. Metals content for supported (fresh) catalyst

Metal [%wt]
Aluminum 11.5
Cerium 12.1
Iron 0.3
Magnesium 9.2
Sodium 0.3
Nickel 8.4

The adsorption-desorption isotherms presented in Figure 6.5 are Type II isotherms,


according to the IUPAC classification.34 It is the normal form of isotherm obtained with a non-
porous or macroporous adsorbent. The Type II isotherm represents unrestricted monolayer-
multilayer adsorption. However, the isotherm shows a hysteresis loop indicating some
mesoporosity in the prepared material. The Type H3 loop (IUPAC classification of hysteresis

132
loops34) seems to be the most similar to the observed experimental one; this loop does not exhibit
any limiting adsorption at high p/p°, and accordingly, it is observed with aggregates of plate-like
particles giving rise to slit-shaped pores. The original NiCe-HDT and kaolin used have a plate-
like morphology which should be giving origin to the type of hysteresis loop observed in the
obtained isotherms. The pore width distribution is also presented in Figure 6.5 and as can be
inferred, the pore distribution in the mesoporous range is very broad with a maximum around
150 Å (15 nm) for the fresh catalyst. Comparing the fresh vs. regenerated catalyst, a drastic
change of the isotherms’ shape (the hysteresis loop), which affected the pore distribution, was
observed. The original abundant pores in the fresh catalyst near 200 Å (20 nm) were almost lost
and shifted to 300 Å (30 nm). Moreover, pores have a very narrow and centered distribution
around 35 Å was observed after regeneration. Comparing the fresh vs. spent catalyst, a similar
drastic change in the isotherms’ shape was evidenced, as well as the appearance of pores near 32
Å. These pores were produced under reaction and were stable after the regeneration step. The
near 300 Å pores were also generated during the reaction and were found to be stable at
regeneration conditions.

The XRD diffractograms for the prepared fixed-bed catalyst and the spent and regenerated
catalysts are depicted in Figure 6.6. Identification of the crystalline phases indicated that the
signals corresponded to NiO (PDF#01-075-0269) and CeO2 (PDF#01-073-6318). As expected,
the kaolin was transformed into the amorphous meta-kaolin after calcination, and thus, only the
signals coming from NiO and CeO2 were the ones detected. Additionally, TiO2 (PDF#01-089-
4921) signal, which is an impurity present in the kaolin used for the preparation of the catalyst
signals, was also detected in the diffractograms. The observed XRD signals are very broad
indicating that the NiO and CeO2 phases present in the prepared catalyst are in the nano-domain
range which is important from the point of view of active sites dispersion and activity. For the
two crystalline phases the obtained average crystalline domain sizes are 5±1 nm. A slight change
of the crystalline domain size after reaction and regeneration to 6-7 nm was obtained.

The XRD of the spent catalyst presents some amorphous materials compared to the fresh
one as depicted in Figure 6.7. Also, the diffractogram’s signals of the regenerated catalyst show
lower intensity than for the fresh catalyst, indicating some amorphisation of the catalyst that
occurred during the activation/reaction/regeneration steps. However, when comparing the

133
regenerated catalyst with the spent one, it can be concluded that the evolution occurred during
the activation/reaction and not during the regeneration step.

Fresh Catalyst

Spent Catalyst

Regenerated Catalyst

Figure 6.5. Textural properties for the supported catalyst: a) adsorption-desorption isotherms and
b) pore width distribution.

134
Fresh Catalyst

300 CeO2 NiO

250

Intensity(CPS)
200
TiO2

150

100

50

0
10 20 30 40 50 60 70 80 90
Two-Theta (deg)

Spent Catalyst
NiO
CeO2

200
Intensity(CPS)

150
TiO2

100

50

0
10 20 30 40 50 60 70 80 90
Two-Theta (deg)

Regenerated Catalyst

250
CeO2
NiO

200
Intensity(CPS)

150
TiO2

100

50

0
10 20 30 40 50 60 70 80 90
Two-Theta (deg)

Figure 6.6. XRD diffractograms for the supported catalyst (fresh, spent and regenerated).
Vertical red lines indicate the CeO2 phase (PDF#01-073-6318); vertical blue lines indicate the
NiO phase (PDF#01-075-0269); Vertical green lines indicate the TiO2 phase (PDF#01-089-
4921).

135
Fresh vs. Spent Catalyst

300 Fresh Catalyst

250

Intensity(CPS)
200

Spent Catalyst
150

100

50

0
10 20 30 40 50 60 70 80 90
Two-Theta (deg)

Fresh vs. Regenerated Catalyst

Fresh Catalyst
300

250 Regenerated Catalyst


Intensity(CPS)

200

150

100

50

0
10 20 30 40 50 60 70 80 90
Two-Theta (deg)

Spent vs. Regenerated Catalyst

Regenerated Catalyst

200
Intensity(CPS)

150
Spent Catalyst

100

50

0
10 20 30 40 50 60 70 80 90
Two-Theta (deg)

Figure 6.7. XRD diffractograms comparison for fresh, spent and regenerated catalyst.

136
6.4.3. Catalytic Steam Cracking with Ultradispersed Catalyst

To verify the performance of the catalysts under the conditions of our tests, hydrogen and
carbon dioxide molar concentrations in the gas product were quantitatively analyzed via gas
chromatography for each evaluated condition. In the same way, as indicated previously, a QMS
was connected along the experimental run for continuous qualitative analysis of the chemical
species in the outlet gas stream. According to the catalytic water dissociation and
oxidation/reforming reactions that are theoretically taking place (Eq. 6.1 and Eq. 6.2)19,20, the
catalyst would promote formation of hydrogen free radicals and also CO2 and H2 and the ratio of
H2 to CO2 should increase compared to the non-catalytic route (thermal cracking).

Catalytic Dissociation 𝐶𝐴𝑇


𝐻2 𝑂 → 𝐻 ∙ + 𝑂𝐻 ∙ Eq. 6.1
of Water

𝐶𝐴𝑇

Oxidation/Reforming 𝑅𝑛′∙ + 2𝑂𝐻 ∙ → 𝑅𝑛−1 + 𝐶𝑂2 + 𝐻2 Eq. 6.2

As seen in Figure 6.8, where GC results are presented, there is an increase from
approximately 10 to 12.4 % v/v of H2 in the gas phase when comparing the results for the
thermal reaction with the catalytic ones. No change in H2 concentration was evidenced when
comparing the three different catalyst formulations. However, comparing the CO2 concentration
for thermal and catalytic reactions, it can be observed that it remained practically unchanged.
Similar results were obtained from the QSM analysis presented in Figure 6.9; this analysis is
more sensitive to small changes in concentration of the gas phase. An increase on the H2 signal
and constant CO2 signal was evidenced. Taking into consideration the H2 to CO2 ratios presented
in both figures, there is an increase for catalytic processing attributed to the slight increase on H2.
These results would suggest that the catalyst is indeed carrying out water splitting providing
hydrogen radicals to the reaction media, but it is not performing an adequate catalytic function in
the oxidation/reforming reaction considering no increase on CO2 was observed, which is critical
to promote eventual enhancement of the liquid and gas product properties.

137
Figure 6.8. Comparison of H2 and CO2 GC results for thermal and catalytic steam cracking using
Ni/Ce UDC. AR: Atomic Ratio (Molar).

Figure 6.9 Comparison of H2 and CO2 QMS results for thermal and catalytic steam cracking
using Ni/Ce UDC.
As illustrated in Figure 6.10, conversion of the residue fraction (560°C+) for thermal and
catalytic reactions carried out at 380 °C using UDC is at the same level with values close to the
30% range. A similar behavior was found for the reactions at 430 °C, with conversion levels

138
within the 40-43% range. This result was expected as the UDC functionality is not to promote
hydrocarbons cracking, but to promote improvement of the product quality via water dissociation
and subsequent hydrogenation. For the same reasons, the liquid product distributions were
similar for thermal and catalytic cracking.

Figure 6.10. Conversion and liquid product distribution for thermal and catalytic steam cracking
of DAO using Ni/Ce UDC.

A summary of the properties of the whole liquid products from the reactions is presented in
Table 6.4. When comparing the catalytic results against the thermal cracking blank, the
following observations can be made: 1) the stability of the liquid products measured by the P-
value stability index is the same for both processes. This indicates that the catalytic function does
not have any effect on the degree of peptization of the asphaltene molecules that would allow the
process to reach a higher level of conversion; 2) MCR content slightly increased for catalytic
processing, which could be related to the thermal treatment at 370 °C during the incorporation of
the catalyst particles to the feedstock. However, the reaction at 423 °C does not follow the same

139
trend. In this case, the MCR content of the liquid product from catalytic processing had a lower
value than that of the thermal cracking reaction under the same conditions; 3) Hydrocarbon
group type analysis (SARA) analysis shows a very similar distribution profile, with an increase
of the saturates fraction, slightly reduced aromatics content, depletion of resins and increase of
the asphaltenes content in relation to the feedstock. Depletion of resins is a disadvantage of both
processes, considering they help to maintain stable asphaltenes micelles in the product. Once
they are thermally broken, exposed asphaltenes tend to agglomerate and precipitate very easily.
In general, properties of the whole liquid products obtained from the UDC reactions are similar
or slightly worse when compared with the non-catalytic route; 4) API Gravity is similar for the
liquid products of thermal and catalytic processing at 380 °C, except for C1-C where a decrease
below the ±0.5 error of the method was observed. Conversely, an enhanced API gravity was
obtained for the catalytic sample at 423°C. These results support our previous affirmation that
suggest that even though the catalyst is promoting an increase on H2 concentration, no significant
enhancement on the liquid properties is evidenced if catalytic oxidation/reforming reaction does
not take place. These findings lead us to suggest that Ni particles, present in the form of NiO,
seem to have a dehydrogenating effect or act as coke promoting seeds, which could be related to
their state of oxidation; Ni would have to be completely reduced in order to promote a
hydrogenating catalytic effect, while Ce particles would be able to carry out water splitting in an
oxide form. Unlike ultradispersed particles, the oxidation state of the catalyst active phase can be
controlled with more flexibility when using a supported catalyst.

140
Table 6.4. Summary of experimental conditions and whole liquid product characterization for
thermal and catalytic steam cracking of DAO using Ni/Ce UDC
Reaction Conditions
ULTRADISPERSED CATALYST -1 -1
380 ºC-0.25 h 423 °C-2 h
RUN DATA FEED* C1-T C1-C C2-C C3-C C2-T* C4-C
Ni/Ce Atomic Ratio N/A 2 1.5 1 N/A 1.5
Conversion HC (560 °C+) [%] 29.9 27.5 29.5 29.8 43.3 40.0
Saturates [%wt] 3.8 10.6 10.4 10.2 10.7 11.9 10.2
Aromatics [%wt] 76.6 73.1 72.5 76.3 77.2 62.6 63.2
SARA
Resins [%wt] 18.9 7.5 7.5 3.9 3.9 11.4 11.9
Asphaltenes-C5 [%wt] 0.7 8.9 9.6 9.6 8.1 14.1 14.7
P-Value N/A 1.35 1.35 1.35 1.35 1.15 1.15
API Gravity 6.0 8.2 7.5 8.0 8.2 8.1 8.9
MCR [%wt] 12.95 14.60 15.86 15.49 14.66 17.18 16.45
Viscosity @25°C [cP] 5339 (60 °C) 4695 6286 4747 4609 2802 1389
Sediments [%wt] 0.11 0.08 0.11 0.12 0.23 0.20
*Experimental data reported in Chapter 4

6.4.4. Catalytic Steam Cracking in a Fixed-bed Reactor

A similar gas analysis as the one conducted for the Ni/Ce UDC reactions was carried out for
the fixed-bed reactions. Figure 6.11 presents the H2 (amu = 2) and CO2 (amu = 44) signals
obtained from the QMS analysis for the catalytic reactions using the reduced Ni/Ce-based
catalyst, as well as the thermal blanks performed with a reaction bed of inert material. As
observed, unlike the UDC results, in this case both H2 and CO2 signals increased for the catalytic
runs at the two studied temperatures (360 and 380 °C). Very importantly, the intensity of the
signals is approximately one order of magnitude higher for the catalytic runs compared to the
thermal blanks. This clearly indicates that both, catalytic water splitting and oxidation/reforming
reactions, are taking place in the process.

Results from QMS analysis were quantitatively corroborated via GC analysis, where the
actual volume percentage of each component was obtained. These results presented in Figure
6.12 show that there is a considerable increase of H2 production from around 6 to 20% v/v at 360
°C and 4 to 18 % v/v at 380 °C. In the same way, CO2 concentration increased from
approximately 4 to 10% v/v at both temperatures when the catalyst was reduced before
conducting the set of reactions. The H2/CO2 ratio practically doubles for the catalytic steam
processing. After the catalyst bed was regenerated and the reaction conducted at 380 °C without

141
the reduction step (oxidized catalyst active phase), the increase of CO2 levels was considerably
lower, approximately 4 to 6 % v/v as observed in Figure 6.12. This result is in agreement with
the results found for the Ni/Ce UDC reactions from the gas analysis point of view.

Figure 6.11. Comparison of H2 and CO2 QMS results for thermal and catalytic steam cracking
using Ni/Ce-based fixed-bed catalyst.

142
Figure 6.12. GC analysis of H2 and CO2 and H2/CO2 ratios for fixed-bed reactions.

As illustrated in Table 6.5, where experimental conditions and bulk properties of the whole
liquid products are presented for the two evaluated temperatures (360 and 380 °C), an increase in
the conversion levels was observed for the catalytic process compared to the thermal blank at
both temperatures, which suggests that, besides water splitting and hydrogenation functionalities,
the fixed-bed catalyst has a hydrocarbon cracking effect. For example, at 360 °C (Section A of
Table 6.5), an increase of 10 conversion points was obtained. This increased conversion is
reflected in a liquid product with a reduced level of viscosity and slightly higher API gravity and

143
a higher content of saturated hydrocarbons. A similar behavior was observed at 380 °C as
presented in Section B of Table 6.5. Comparing C1-FB-CAT and C3-FB-CAT, which are both
catalytic runs at the same conditions (360 °C-0.25 h-1) with the reduced catalyst, a decay on the
level of conversion from 22.3 to 18.3 is evidenced. This decrease in conversion produces an
increase in the viscosity of the liquid product, as well as a decrease in the API gravity, however
still within the error of the technique (0.5 API). This is a clear evidence of catalyst deactivation
taking place in the reaction bed, possibly caused by coke deposition on the catalyst surface. Now,
taking into consideration properties that have a detrimental effect on the quality of the liquid
products, such as increased MCR and nC5-asphaltenes content, an increase of these properties is
observed in the same way for the catalytic process, which was somehow expected considering
they have a linear increasing trend with conversion. In this way, in order to assess the net effect
of the water chemistry taking place on the catalytic steam cracking process and quantifying the
actual effect on the bulk properties of the products, upgrading via thermal cracking, catalytic
steam cracking with UDC and catalytic steam cracking with fixed-bed catalyst had to be
compared at similar conversion levels (isoconversion). This comparison is presented in Section
6.4.5.

As mentioned previously, the main reason for conducting the catalytic steam cracking study
via fixed-bed catalyst was to be able to evaluate the effect of the oxidation state of the catalyst
active phase on the performance of the process. In this way, it could be determined whether the
oxidation state of the ultradispersed catalyst particles was the main cause for their poor
hydrogenation performance for improving product quality. It is important to highlight that the
oxidation state of the UDC particles cannot be easily controlled with the current method of
preparation. Analyzing the results presented in Section B of Table 6.5 for the reduced (C2-FB-
CAT) and oxidized (C4-FB-CAT), a significant 10% relative decrease on the conversion level
was evidenced when the catalyst was oxidized. However, this decrease did not have any effect
on the liquid product properties since similar levels of viscosity, API gravity, MCR and
asphaltenes were obtained. In addition, the stability of the liquid product was not compromised
when that catalyst active phase was not in a reduced state. This contrasts with the gas analysis
results where a clear decrease of the CO2 level from the oxidation/reforming reaction was
observed.

144
Table 6.5. Summary of experimental conditions and whole liquid product characterization for
thermal and catalytic steam cracking of DAO using Ni/Ce-based fixed-bed catalyst. Section A:
reactions at 360 °C. Section B: reactions at 380 °C

-1
Section A T=360 °C and WHSV = 0.25 h
Thermal Catalytic
Experiment C1-FB-T C1-FB-CAT C3-FB-CAT
Conversion (560 °C+) 12.3 22.3 18.3
Saturates [%wt] 5.7 7.5 6.9
Aromatics [%wt] 74.9 70.4 76.6
SARA
Resins [%wt] 14.7 15.5 10.5
Asphaltenes-C5 [%wt] 4.7 6.6 5.9
P-Value >2.0 >2.0 >2.0
API Gravity 7.2 7.6 7.4
MCR [%wt] 13.75 14.28 14.43
Viscosity @25°C [cP] 59806 13748 25890

-1
T= 380 °C and WHSV = 0.25 h
Section B Catalytic
Thermal
Reduced Oxidized
Experiment C2-FB-T C2-FB-CAT C4-FB-CAT
Conversion (560 °C+) 31.0 42.9 38.6
Saturates [%wt] 9.5 12.9 14.0
Aromatics [%wt] 68.3 68.2 66.5
SARA
Resins [%wt] 12.1 6.4 6.5
Asphaltenes-C5 [%wt] 10.1 12.5 13.0
P-Value 1.65 1.30 1.35
API Gravity 9.0 9.6 9.57
MCR [%wt] 14.90 16.52 16.88
Viscosity @25°C [cP] 4626 1304 1199

6.4.5. Comparison of Thermal vs. UDC vs. FB Processing

Table 6.6 gathers the product properties for thermal cracking in an empty-tube and catalytic
steam cracking using ultradispersed and fixed-bed catalysts at similar conversion levels, within
the 40-43 % range approximately. As can be deduced from the results, the performance of
thermal cracking and the catalytic process with UDC is very similar in terms of product quality.
However, there is a slight increase on API and decrease on MCR and viscosity, which gives

145
added-value to the final product. Conversely, when using the fixed-bed catalyst there is a more
marked differentiation between the catalytic route compared to the thermal baseline mainly in
terms of API gravity, asphaltenes content, and viscosity. Very importantly, the P-value of the
liquid product is considerably higher, which would allow to increase the severity of the process
to obtain a higher conversion level until reaching the commonly acceptable value of 1.15;
thermal and UDC processing are already at the limit value. Preliminary estimations based on
thermal cracking results in a previous work suggest that conversion could be increased by 10%
relative when P-value is reduced from 1.3 to 1.15. This would provide a lighter liquid product.

Table 6.6. Whole liquid product characterization comparison for thermal and catalytic runs at
isoconversion

PRODUCT QUALITY
Empty Fixed
Property UDC
Tube Bed
Conversion HC (560 °C+) [%] 43.3 40.0 42.9
Saturates [%wt] 11.9 10.2 12.9
Aromatics [%wt] 62.6 63.2 68.2
SARA
Resins [%wt] 11.4 11.9 6.4
Asphaltenes-C5 [%wt] 14.1 14.7 12.5
P-Value 1.15 1.15 1.3
API Gravity 8.1 8.9 9.6
MCR [%wt] 17.18 16.47 16.52
Viscosity @25°C [cP] 2802 1389 1304

Comparing the liquid product distributions for the three processes as depicted in Figure 6.13,
it can be observed that there is no marked difference in the distribution of products by boiling,
which suggests these distributions are controlled by the level of conversion achieved and not the
nature of the process. Nontheless, as indicated previously, the bulk properties of the liquid
product vary depending on the upgrading route (thermal or catalytic).

146
100
Gas

80 Naphtha

Kerosene
60
%wt

Diesel
40 LVGO

HVGO
20
DAO (560°C+)
0 Asphaltenes-C5
Empty Tube UDC Fixed Bed

Figure 6.13. Liquid product distributions at isoconversion.

6.5. Conclusions

It was possible to prepare Ni/Ce-based UDC formulations for low temperature catalytic
steam cracking of deasphalted vacuum residue, with actual concentrations within a reasonable
error compared to the target. Similarly, a Ni/Ce-based supported catalyst formulation was
prepared for comparison purposes. Even though it was possible to produce H2 via water
splitting/dissociation using the UDC formulation, no improvement on the quality of the liquid
products was evidenced. However, changes on the gas phase induced by the catalytic process
were evidenced. On the other hand, catalytic steam cracking of DAO using a pre-reduced fixed-
bed catalyst presented improved conversions and an enhanced product quality when compared to
conventional thermal cracking and catalytic steam cracking using Ni/Ce UDC. This hydrocarbon
catalytic cracking effect is attributed to the presence of acid sites in the catalyst. Also, even
though the oxidation state of the catalyst active phase has a direct effect on the gas product
properties, bulk liquid product quality presented very similar characteristics as those of
processing with the reduced catalyst. Based on the results, ceria particles can perform the water
splitting functionality in the form of oxides. However, Ni particles have to be completely
reduced to effectively carry out its function in the partial oxidation/reforming reaction, which
leads to product improvement. Considering the analysis of the supported catalyst after reaction,

147
considerable changes in textural properties due to deposition of organic matter (16.6% wt) in the
catalyst surface were observed, having reduced surface area and drastic changes of the
isotherms’ shape, which affected the pore distribution of the catalyst after reaction/regeneration.
Also, amorphisation of the catalyst crystalline structure during the activation/reaction step and
slight increase on the crystalline domain sizes after reaction/regeneration were evidenced.

6.6. Acknowledgement

The authors are grateful to the Natural Sciences and Engineering Research Council of
Canada (NSERC), Nexen-CNOOC Ltd, and Alberta Innovates-Energy and Environment
Solutions (AIEES) for the financial support provided through the NSERC/NEXEN/AIEES
Industrial Research Chair in Catalysis for Bitumen Upgrading. Also, the contribution of facilities
from the Canada Foundation for Innovation, the Institute for Sustainable Energy, Environment
and Economy, the Schulich School of Engineering and the Faculty of Science at the University
of Calgary are greatly appreciated. Finally, special thanks to Dr. Gerardo Vitale, Research
Associate at the Catalysis for Bitumen Upgrading Group, for his technical contributions for the
development of this work.

6.7. References

(1) Pereira-Almao, P.; Marzin, R.; Zacarias, L.; Cordova, J.; Carrazza, J.; Marino, M. Steam
conversion process and catalyst. U.S. Patent No. US5885441, 1999.
(2) Ancheyta, J., Modeling and Simulation of Catalytic Reactors for Petroleum Refining. Wiley:
2011.
(3) Bartholomew, C. H.; Farrauto, R. J., Fundamentals of Industrial Catalytic Processes. Wiley:
2011.
(4) Strausz, O. P.; Lown, E. M.; Institute, A. E. R., The Chemistry of Alberta Oil Sands,
Bitumens and Heavy Oils. Alberta Energy Research Institute: 2003.
(5) Gray., M. R., Upgrading oilsands bitumen and heavy oil. The University of Alberta Press:
Edmonton, 2015.
(6) Ancheyta, J.; Trejo, F.; Rana, M. S., Asphaltenes: Chemical Transformation During
Hydroprocessing of Heavy Oils. Taylor & Francis Group: 2009.
(7) Pereira-Almao, P.; Ali-Marcano, V. A.; Lopez-Linares, F.; Vasquez, A. Ultradispersed
catalyst compositions and methods of preparation. U.S. Patent No. US7897537B2, 2011.
(8) Duprez, D.; Miloudi, A.; Delahay, G.; Maurel, R. Selective steam reforming of aromatic
hydrocarbons: IV. Steam conversion and hydroconversion of selected monoalkyl- and dialkyl-
benzenes on Rh catalysts. J. Catal. 1984, 90 (2), 292-304.
(9) Duprez, D.; Miloudi, A.; Delahay, G.; Maurel, R. Selective steam reforming of aromatic
hydrocarbons: V. Steam conversion and hydroconversion of selected monoalkyl- and
dialkylbenzenes on Pt and Ni catalysts. J. Catal. 1986, 101 (1), 56-66.

148
(10) Duprez, D. Selective steam reforming of aromatic compounds on metal catalysts. Appl.
Catal., A 1992, 82 (2), 111-157.
(11) Clark, P. D.; Kirk, M. J. Studies on the Upgrading of Bituminous Oils with Water and
Transition Metal Catalysts. Energy Fuels 1994, 8 (2), 380-387.
(12) Fumoto, E.; Tago, T.; Tsuji, T.; Masuda, T. Recovery of Useful Hydrocarbons from
Petroleum Residual Oil by Catalytic Cracking with Steam over Zirconia-Supporting Iron Oxide
Catalyst. Energy Fuels 2004, 18 (6), 1770-1774.
(13) Fumoto, E.; Tago, T.; Masuda, T. Production of Lighter Fuels by Cracking Petroleum
Residual Oils with Steam over Zirconia-Supporting Iron Oxide Catalysts. Energy Fuels 2005, 20
(1), 1-6.
(14) Fumoto, E.; Matsumura, A.; Sato, S.; Takanohashi, T. Recovery of Lighter Fuels by
Cracking Heavy Oil with Zirconia−Alumina−Iron Oxide Catalysts in a Steam Atmosphere.
Energy Fuels 2009, 23 (3), 1338-1341.
(15) Junaid, A. S. M.; Street, C.; Wang, W.; Rahman, M. M.; An, W.; McCaffrey, W. C.;
Kuznicki, S. M. Integrated extraction and low severity upgrading of oilsands bitumen by
activated natural zeolite catalysts. Fuel 2012, 94 (0), 457-464.
(16) Pinilla, J. L.; Arcelus-Arrillaga, P.; Puron, H.; Millan, M. Selective Catalytic Steam
Cracking of anthracene using mesoporous Al2O3 supported Ni-based catalysts doped with Na,
Ca or K. Appl. Catal., A 2013, 459 (0), 17-25.
(17) Pinilla, J. L.; Arcelus-Arrillaga, P.; Puron, H.; Millan, M. Reaction pathways of anthracene
selective catalytic steam cracking using a NiK/Al2O3 catalyst. Fuel 2013, 109 (0), 303-308.
(18) Carrazza, J.; Martinez, N.; Pereira, P. Process and catalyst for upgrading heavy
hydrocarbon. U.S. Patent No. US5688395, 1997.
(19) Fathi, M. M.; Pereira-Almao, P. Catalytic Aquaprocessing of Arab Light Vacuum Residue
via Short Space Times. Energy Fuels 2011, 25 (11), 4867-4877.
(20) Pereira-Almao, P.; Trujillo-Ferrer, G.; Peluso, E.; Galarraga, C.; Sosa, C.; Algara, C. S.;
Lopez-Linares, F.; Carbognani-Ortega, L. A.; Zerpa-Reques, N. G. Systems and Methods for
Catalytic Steam Cracking of Non-Asphaltene Containing Heavy Hydrocarbons. U.S. Patent No.
US20130015100A1, 2013.
(21) Zerpa-Reques, N. G.; De Clerk, A.; Xia, Y.; Omer, A. A. Upgrading of hydrocarbon
material. U.S. Patent No. US20130043033, 2016.
(22) De Klerk, A.; Zerpa-Reques, N. G.; Xia, Y.; Omer, A. A. Integrated central processing
facility (cpf) in oil field upgrading (ofu). U.S. Patent No. US20140138287A1, 2014.
(23) Trujillo-Ferrer, G. Thermal and Catalytic Steam Reactivity Evaluation of Athabasca
Vacuum Gasoil. MSc Thesis, University of Calgary, Calgary, 2008.
(24) Garcia Hubner, E. Catalysts Evaluation for Catalytic Steam Cracking of De-asphalted Oil in
a Fixed Bed Reactor. MSc Thesis, University of Calgary, Calgary, 2015.
(25) Rodriguez-DeVecchis, V. M.; Carbognani Ortega, L.; Scott, C. E.; Pereira-Almao, P. Use of
Nanoparticle Tracking Analysis for Particle Size Determination of Dispersed Catalyst in
Bitumen and Heavy Oil Fractions. Ind. Eng. Chem. Res. 2015, 54 (40), 9877-9886.
(26) Scherrer, P. Bestimmung der Größe und der inneren Struktur von Kolloidteilchen mittels
Röntgenstrahlen. Göttinger Nachrichten Math. Phys 1918, 2, 98-100.
(27) JADE v 7.5.1 XRD Pattern Processing Identification & Quantification, Materials Data Inc.:
2005.
(28) Carbognani, L.; Lubkowitz, J.; Gonzalez, M. F.; Pereira-Almao, P. High Temperature
Simulated Distillation of Athabasca Vacuum Residue Fractions. Bimodal Distributions and

149
Evidence for Secondary “On-Column” Cracking of Heavy Hydrocarbons. Energy Fuels 2007, 21
(5), 2831-2839.
(29) Carbognani, L.; Gonzalez, M. F.; Pereira-Almao, P. Characterization of Athabasca Vacuum
Residue and Its Visbroken Products. Stability and Fast Hydrocarbon Group-Type Distributions.
Energy Fuels 2007, 21 (3), 1631-1639.
(30) Hassan, A.; Carbognani, L.; 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.
(31) Carbognani Ortega, L.; Rogel, E.; Vien, J.; Ovalles, C.; Guzman, H.; Lopez-Linares, F.;
Pereira-Almao, P. Effect of Precipitating Conditions on Asphaltene Properties and Aggregation.
Energy Fuels 2015, 29 (6), 3664-3674.
(32) Galarraga, C. E.; Pereira-Almao, P. Hydrocracking of Athabasca Bitumen Using
Submicronic Multimetallic Catalysts at Near In-Reservoir Conditions. Energy Fuels 2010, 24
(4), 2383-2389.
(33) Di Carlo, S.; Janis, B. Composition and visbreakability of petroleum residues. Chem. Eng.
Sci. 1992, 47 (9), 2695-2700.
(34) Sing, K. S. W. Reporting physisorption data for gas/solid systems with special reference to
the determination of surface area and porosity (Recommendations 1984). Pure Appl. Chem.
1985, 57 (4), 603-619.

150
Chapter 7: Conclusions and Recommendations

Each one of the paper-based chapters presented in this work contained its own detailed
conclusions. In this way, a summary of the key concluding remarks is presented in this section
along with some recommendations for directing a possible the next stage of investigations in this
area.

7.1. Conclusions

Pilot Plant Procurement, Design and Construction

A Reactivity Test Unit (RTU-1) that resembles a conventional industrial thermal cracking
unit was successfully procured, designed and constructed. The design incorporated an operator-
friendly vertical type of layout and a wide range of operability conditions and type of reactions,
in such a way that this versatility allows the use of the experimental setup not only for the
purpose of this work, but for most areas of interest of the Catalysis for Bitumen Upgrading
(CGU) group (i.e. hydrocracking evaluations or bitumen in-situ upgrading evaluations).

The performance of the RTU-1 in terms of stability, safety, reliability and control was
successfully assessed and met the expectations established in the proposed design. It was
possible to carry out experimental tests at different reaction conditions for both thermal and
catalytic processes with good results in terms of operation of the reactivity unit for three different
feedstocks. The bench-scale setup showed a consistent behavior for the evaluation of more than
50 experimental conditions for different types of processing.

Upgrading of DAO via Thermal Cracking

Even though thermal cracking is a heavily researched area, the thermal cracking behavior of
deasphalted hydrocarbons and previously cracked hydrocarbons, is an area relatively unexplored.
Reactivity of a pre-cracked deasphalted vacuum residue (converted DAO) and a virgin
deasphalted vacuum residue (virgin DAO) were evaluated spanning a wide range of operating
conditions and conversion levels. Experimental conditions were chosen to resemble an industrial
thermal cracking operation. This allowed us to create correlations to estimate the properties of
the whole liquid product as a function of conversion. Effects of LHSV, temperature and pressure
on the performance of the process were found to have a similar behavior as for upgrading of

151
residual hydrocarbons via thermal cracking. However, the absence of asphaltenes on the
feedstock resulted in an enlarged limit stability that allowed the process to reach higher
conversion levels compared to thermal cracking of asphaltenes containing hydrocarbon fractions
such as vacuum residue. Very importantly, despite the absence of asphaltenes in the evaluated
feedstocks, a considerable yield of this fraction was obtained, especially for the pre-cracked
DAO.

A comprehensive new approach for kinetic modeling not reported in the open literature thus
far was successfully developed based on a multi-parameter non-linear optimization programmed
in Matlab. Reasonable modeling errors around 7% were found. The main novelties of the model
were the inclusion of an asphaltenes generation reaction and differentiating the reactivity of
vacuum gas oil by splitting it into heavy vacuum gas oil (HVGO) and light vacuum gas oils
(LVGO) and the distillates lump split into diesel and kerosene instead of the simplified lumping
as distillates. This innovative 8-lumps kinetic modeling approach will be instrumental for the
development of DAO thermal cracking aspects in bitumen upgrading. This kinetic model allows
modeling of thermal cracking to predict liquid product yields in either a once-through
configuration of virgin DAO thermal cracking or also in an upgrading scheme including a
recycle configuration.

Kinetic parameters were within common values found in the literature. In addition,
significant differences were found between the kinetic parameters for a pre-cracked and a virgin
DAO. This was attributed to the chemical nature of each feedstock, with the pre-cracked being
more aromatic and less branched than the virgin feedstock. Additionally, three different
scenarios of sequences on the activation energies were evaluated, in an attempt to incorporate
chemical considerations related to the nature of the feedstock and the process into the kinetic
modeling. The case where activation energies of conversion of heavier lumps are lower than
those for the conversion of lighter ones and the activation energies of formation of lighter lumps
are higher than those for the formation of heavier lumps was found to be the most appropriate for
kinetic modeling of DAO thermal cracking.

152
Upgrading of DAO via Catalytic Steam Cracking

When evaluating the upgrading of pre-cracked deasphalted vacuum residue using a Ni/K
UDC formulation, it was found that the selection of the operating conditions were critical in
order to have a catalytic activity for water dissociation and hydrogen production. The capabilities
of the Ni/K were evidence at high reaction temperatures (T>430ºC). Nonetheless, no significant
enhancement of product quality was observed with this Ni/K catalyst formulation at the same
conversion levels obtained for thermal cracking. The severe conditions combined with the high
aromaticity of the feedstock that makes it prone to produce insoluble compounds limit the correct
performance of hydrogenating functions due to the increased production of coke. Product
distribution obtained followed the same trend as in thermal cracking. Thus, a UDC does not
induce hydrocarbon cracking per se. In addition, as for thermal cracking, liquid product stability
seems to be the major constraint to reach increased conversions. From the kinetic point of view,
a change of the mechanism was observed under the studied high-severity conditions. The DAO
(560 °C+) fraction is more selective towards production of HVGO compared to TC. Moreover,
HVGO and LVGO fractions are much more reactive under CSC operating conditions.

New catalyst formulations were proposed to address the issue of the high operating
temperature needed for the Ni/K UDC and, considering the impact that an optimized catalyst
formulation could have on product quality and consequently the added-value this might add to
current thermal cracking technologies, the development of new catalyst formulations for more
moderate conditions for this particular feedstock was pursued. In this way, new Ni/Ce-based
catalyst formulations aiming to have a catalytic steam cracking activity at low temperature
(T<380°C) were investigated.

The ability of Ni/Ce-based ultradispersed and fixed-bed catalysts to produce H2 via water
splitting/dissociation was evidenced. Even though changes in the gas phase induced by the
catalytic process were noticed when using UDC formulations, no improvement on the quality of
the liquid products was observed. Conversely, catalytic steam cracking of DAO using a reduced
fixed-bed catalyst presented improved conversions and an enhanced product quality when
compared to conventional thermal cracking. In addition, the oxidation state of the catalyst active
phase was found to have an effect on conversion and on the composition of the gas phase, having
a decreased conversion as that of a pre-reduced catalyst and also lower CO2 production from

153
oxidation/reforming reactions. Advantages in terms of yields for the catalytic bed processing
were evidenced, however, no significant differences on the liquid product quality were observed.

7.2. Recommendations

The main target of this work was to study the catalytic steam cracking of DAO using
ultradispersed catalyst formulation in order to minimize effects of catalyst deactivation present
when using supporting catalyst with a heavy hydrocarbon feedstock. However, in order to have a
comparable activity as that of a supported catalyst, the particle diameter has to be preferably in
the nano-scale (<100 nm) so that the surface area of active sites per unit of mass of catalyst is
maximized. As observed in the results of this doctoral investigation, particle sizes obtained with
the current skid-preparation unit are close to 180 nm. Even though this number is at the sub-
micronic scale, it is still far from the nanometric range. In this way, it is recommended to
evaluate alternatives for reduction of the catalyst particle size to values lower than 100 nm to
compare the activity of the catalyst as a function of the particle size. This can be accomplished
by changing the method of preparation of the micro-emulsions by using a high efficiency mixing
system (i.e. high shear mixing), instead of the current static mixers method.

Another factor to take into consideration is the amount of catalyst incorporated in the
feedstock. All experimental evaluations in this doctoral work using UDC formulations were
conducted using a nominal catalyst amount of 700 ppm. In this way, it would be interesting to
evaluate higher catalyst concentrations in order to discard that the low activity for hydrogenation
is not being caused by an inadequate amount of catalyst in the reaction media. Additionally,
tracking of catalyst particles at different sections of the experimental setup is recommended in
order to evaluate the extent of catalyst loses in pipelines and connections and in that way
determine the actual amount of catalyst being fed to the reactor.

Moreover, looking forward to improving the hydrogenating function of the UDC, it is


recommended to investigate possibilities to produce reduced nickel UDC particles to assess its
effect on the performance of the process. The oxidation state was found to have an effect on the
catalytic activity when using a fixed-bed catalyst. Controlling the oxidation state with the current
method of preparation to produce reduced nickel is very challenging. However, the reaction
environment has combined reducing and oxidizing effects that if somehow can be controlled to

154
have reduced catalyst active sites, the catalytic functionality for hydrogenation could be
maximized.

The issue of olefins reduction without the use of hydroprocessing is of relevance for
bitumen field upgrading applications. Olefins are concentrated in the lighter hydrocarbon
fractions, such as naphtha and distillates, where catalytic steam cracking could be applied mainly
to target olefins reductions instead of conversion to lighter hydrocarbons. One investigation in
this is already taking place at the Catalysis for Bitumen Upgrading (CBU) group, but due to the
potential application of CSC in a field upgrader scheme due to its relative simplicity and the
absence of hydrogen use, further research in this area is encouraged.

155
Appendixes

Appendix A – Standard Operating Procedure (SOP) – RTU-1

University of Calgary
HEALTH, SAFETY AND ENVIRONMENTAL
STANDARD OPERATING PROCEDURE Issued: Fredy Cabrales
FOR: Reactivity Test Unit 1 (RTU-1) Revised: Alejandro Coy
(Heavy Feedstock-based) Reviewed:
Pages

1. Purpose / Background

The Reactivity Test Unit No. 1 (RTU-1) was built to simulate and evaluate the reactivity of
heavy oils under Thermal Cracking (TC), Catalytic Steam Cracking (CSC), or Hydroprocessing
in a research-scale setup under a wide range of variables and conditions (temperature, pressure,
feed flow, steam flow, residence time, catalytic fixed-bed, ultradispersed catalyst, etc.) using
high throughput experimentation. Gustavo Trujillo built a first version of the RTU-1 in 2008.
This new version includes a re-design and complete restructuration of the previous unit.

This procedure is written to comply with all of the U of C HS&E regulatory requirements,
and to teach new users, the standard operative protocols of the RTU-1

2. Scope

This procedure is intended for all those workers and/or students/interns working in the
Upgrading and Refining Lab that will be operating the Reactivity Test Unit No. 1 (RTU-1). The
procedure covers specifics about starting up of the unit, pressurization, warming up, stabilization
period, mass balance procedures, shutting down, and learning experiences.

156
3. Prerequisites and Safety Issues

Specific information related to the feed to be processed, catalytic formulation, operational


conditions, test length, have to be provided in the experimental plan. Typical machine shop tools
are required in case of any repair, change or adjustment in the system.

Successful completion of the University of Calgary generic WHMIS course and H2S Alive
course are also required.

Protocol:

 Do not work alone in the laboratory, if required follow University of Calgary procedures for
working alone (http://www.ucalgary.ca/safety/workingalone)
 Be aware of the inherent risks associated to, and the nature of, the processes and materials
used in the laboratory. This is not limited to the risks and hazards in one own equipment, but
extended to include those of every other person in the laboratory.
 Wear appropriate laboratory attire (lab coat, safety glasses, and closed shoes).
 Maintain a neat and organized working environment.
 If leaving an experiment unattended, ensure others are aware and understand the situation
including how to deal with an emergency situation. Leave a telephone number at the
experiment for contact purposes.
 Dispose of waste in a safe and environmentally friendly manner.
Campus Security: 403 220 5333

Hazard Identification (Hazardous Chemicals or Processes):

 Hazardous Chemicals
 Products and by-products expected from the system are:
o Hot heavy oils from the feed section, reactor and heavy product tanks.
o Nitrogen and/or Hydrogen compressed gas
o Light liquid hydrocarbons
o Hydrocarbon gases
o H2S in concentrations lower than 200 ppm

157
Others:

o Toluene
Processes:

Heavy oils flowing under moderate pressure and temperature conditions.

Hazard Assessment:

Products:

Reaction Gas is sweetened in a KOH or NaOH solution at the outlet gases stream.

Personnel must wear personal protective equipment (PPE) all the time when working with
the unit.

Use of toluene should be done discretely and only when strictly needed.

Processes:

Heat insulation is used in all process lines under high temperature conditions.

Pressure relief device is located in the feed section, pump outlet line, to protect the system in
the case of an overpressure caused by plugging of process line.

Engineering/Ventilation Controls:

The RTU-1 was built inside of an enclosure with an extraction system, which was specially
designed to suit the needs of the process. All venting lines (release and exhaust gas lines from the
unit) are also disposed to this extraction.

Personal Protective Equipment:

Personal protective equipment (PPE) required for this SOP includes but is not limited to:

Nitrile gloves, safety glasses and laboratory coats are standard safety equipment for all
employees, students and interns in the lab.

158
Respirators with adequate filters/cartridges for organic vapours when cleaning any oil or
emulsion spill from the plant using any type of solvent like toluene, acetone, gasoline, etc.

Quartz gloves are needed for handling high temperature objects.

4. Procedures

4.1. Plant Preparation

4.1.1. General Preparation

1. Verify the plant is not energized, by checking the electrical box, whose location can be
seen in Figure 1. At the beginning of each week or after a long period of time without
using the plant all switches must be turned off.

Figure 1. Electrical box RTU-1.


2. Make sure ALL valves in the plant are closed.

3. Energize the plant by turning on the switches in the power supply box.

4. Open Software according to the Annex 1. RTU-1 Control Software.

5. Check the instrument air pressure (Figure 2); it should be within the 90-100 psig range.

159
Figure 2. Instruments air supply.
6. Check Nitrogen cylinder pressure, it should not be below the operating pressure plus
200 psig. In this case, if the operating pressure is 300 psig, it should not be below 500
psig.

7. Set the regulator valve at the Nitrogen cylinder so the supply is the operating pressure
plus 100 psig. In this case, if the operating pressure is 300 psig, it should be set to 400
psig.

8. If Hydrogen is used in the reaction, check the Hydrogen cylinder pressure, it should not
be below the operating pressure plus 200 psig. In this case, if the operating pressure is
300 psig, it should not be below 500 psig. Please note that this recommended value
might vary if the Hydrogen flow in the reaction is relatively high, since the 200 psig of
Hydrogen might not be enough for the duration of the run.

9. If Hydrogen is used in the reaction, set the regulator valve at the Hydrogen cylinder so
the supply is at least the operating pressure plus 50 psig. In this case, if the operating
pressure is 300 psig, it should not be set below 350 psig.

10. Check the supply pressure of the feed tank pressure regulator (Figure 3); it should be set
to 70 psig.

160
Figure 3. Feed tank pressure regulator.
11. Verify there is no pressure in the plant by checking the pressure gauges and indicators
and the pressure readings of the feed pumps and water pumps. In case the pressure
readings are above 10 psig, open the drain valve on the pressurized section and closed it
after the section is completely depressurized.

12. Verify there is no liquid in the stability, mass balance, and liquid trap tanks. If there is
any, drain everything and remember to close ALL valves.

4.1.2. Pressurization Procedure

During the operation of the RTU-1, it is often required to pressurize some sections of the
plant. The pressurization can be done on four different locations of the plant, the heavy products
stability tank, the heavy products mass balance tank, the light products stability tank, and the
light products mass balance tank. The following steps are intended to carry out this practice in an
efficient and safe manner.

Heavy products section pressurization.

1. Go to the “Pressurization System” panel on the heavy products section (Figure 4) and
align the tank selection 3-way valve to the tank to be pressurized (stability tank or mass
balance tank).

2. Align the gas selection 3-way valve to Nitrogen.

161
3. VERY SLOWLY turn the Nitrogen regulator valve and increase the pressure at a pace of
50 psi per minute.

4. When reaching the desire pressure, close the Nitrogen regulator valve.

5. Align the gas selection 3-way valve to the OFF position.

6. Align the tank selection 3-way valve to the OFF position.

Figure 4. Pressurization system for heavy products section.


Light products section pressurization.

1. Go to the “Pressurization System” panel on the light products section (Figure 5) and
align the tank selection 3-way valve to tank to be pressurized (stability tank or mass
balance tank)

2. Align the gas selection 3-way valve to Nitrogen.

3. VERY SLOWLY turns the Nitrogen regulator valve and increase the pressure at a pace
of 50 psi per minute.

4. When reaching the desire pressure, close the Nitrogen regulator valve.

5. Align the gas selection 3-way valve to the OFF position.

162
6. Align the tank selection 3-way valve to the OFF position.

Figure 5. Pressurization System for light products section.


4.1.3. Depressurization Procedure

During the operation of the RTU-1 often is required to depressurize some sections of the
plant. The depressurization can be done on four different locations of the plant, the heavy
products stability tank, the heavy products mass balance tank, the light products stability tank
and the light products mass balance tank. The following steps are intended to carry out this
practice in an efficient and safe manner.

Heavy products section depressurization.

1. Go to the “Depressurization System” panel on the heavy products section (Figure 6) and
align the tank selection 3-way valve to the tank to be depressurized (stability tank or
mass balance tank)

2. VERY SLOWLY turns the depressurization regulator valve to reduce the pressure at a
pace of 50psi/min approx.

3. When reaching the desire pressure, close the depressurization regulator valve

4. Align the tank selection 3-way valve to the OFF position.

163
Figure 6. Depressurization system for heavy products section.
Light products section depressurization.

1. Go to the “Depressurization System” panel on the light products section (Figure 7) and
align the tank selection 3-way valve to the tank to be depressurized (stability tank or
mass balance tank)

2. VERY SLOWLY turn the depressurization regulator valve to reduce the pressure at a
pace of 50psi/min approx.

3. When reaching the desire pressure, close the depressurization regulator valve.

4. Align the tank selection 3-way valve to the OFF position.

164
Figure 7. Depressurization system for light products section.
4.2.Start Up
4.2.1. Heating Ramps

During the start up and stabilization of the plant, some areas require to be heated using the
heating tapes. When carrying out this heating, the following temperature increase ramps shall be
used:

Temperature Range [˚C] Ramp [˚C / min]


0 - 200 4
200 - 300 3
300 - 400 2
400 - 440 1
Table 1. Temperature ramps.

4.2.2. Feed Loading

1. In order to measure the amount of feed loaded to the feed tank, make sure to weight the
can with the feed before and after the it is loaded in the feed tank.

165
2. Preheat the feed tank to receive the load using TC-101 to 140 ˚C if the feedstock is very
heavy (vacuum residue, DAO or similar), or to 60 ˚C if the feedstock is a VGO or
similar, or simply ambient temperature if the feedstock is lighter (i.g. gasoline, toluene,
etc).

Note: These pre-heating temperatures might vary depending on the


characteristics of the feedstock used. The idea is to have the feedstock viscous
enough so the feed pump cylinders can be successfully filled up (lower than 400
cP at pumping temperature).

3. If the feed is not mobile at ambient temperature, preheat it using the temperatures in the
previous paragraph as reference and load it into the feed tank.

4. When feed loading is finished, close the feed tank and pressurize it at 70 psig with
Nitrogen opening the tank pressurization valve (Figure 8).

5. In case the pressure in the feed tank gets over 70 psig, SLOWLY open the tank
depressurization valve (Figure 8) until reaching the desire pressure.

Figure 8. Pressurization/depressurization valves for feed tank.


4.2.3. Feed Pumps Loading

1. Preheat the lines to the pumps and the pumps’ cylinders according to the following
Table 2, using the ramps previously established.

166
Set Point Temperature [˚C]
Controller
DAO/VR Feed VGO Feed
TC-101 Feed Tank 140 60
TC-102 Line to Pumps 140 60
TC-103 Pump B Cylinder 140 60
TC-104 Pump A Cylinder 140 60
TC-105 Pump B Valves 140 60
TC-106 Pump A Valves 140 60
TC-107 Pumps Discharge 140
TC-108 Pressure Relief Valve 140
Table 2. Feed section heating temperatures.

2. Once the temperature of the feed in TI-101 is close to the set point and to ensure there is
no air bubbles inside the pumps or the feed lines, a suction test must be carried out. Open
the valve to allow flow from the feedstock tank to the pumps (Figure 9) making sure the
valve from the auxiliary tank used for cleaning is closed. Load each pump with
approximately 100 cc of feed and then pump the content out using the drain valves on
each pump following the guidelines in Annex A.

Figure 9. Tank to pump valve.

167
3. After the suction test is finished, load at least one of the feed pump cylinders and set it
for continuous operation mode according to the Annex A.

4.2.4. Steam and/or Hydrogen Injection

1. Align the 3-way valves of the back-pressure selection (Figure 10) to the appropriate
back-pressure valve (BPV), depending on the pressure of the experiment.

Figure 10. Back-pressure valves.


2. Open the back-pressure valve (BPV).

3. If steam is needed, check the water feed reservoir levels and make sure they are filled
with demineralized water and open the valve to allow flow of water to the plant (Figure
11).

Figure 11. Water reservoir.

168
4. Open the input and output utility water supply valves (Figure 12).

Figure 12. Utility water supply valves.


5. Open the valves to supply cooling water to all the condensers (Figure 13).

Figure 13. Condensers cooling water supply valves.


6.Align the heavy and light products section 3-way valves (Figure 14) to the stability
tanks if the product collection is semi-continuous or to the mass balance tank if it is
continuous.

Figure 14. Heavy and light products section 3-way valves.

169
7. Start preheating the rest of the plant according to the following Table 3, using the ramps
previously established.

Set Point Temperature [˚C]


Controller Hydrogen
Steam
DAO/VR VGO

TC-109 Line to Reactor 1 200


TC-201 Steam Generator/Hydrogen Heater* Tb+70 200
TC-202 Line to Furnace 200
TC-203 Furnace 200
TC-204 Reactor Bottom Section 200
TC-205 Reactor Middle Section 200
TC-206 Reactor Top Section 200
TC-301 Products Outlet 200 100 60
TC-302 Line to Stability Tank 200 100 60
TC-303 Line to Mass Balance Tank 200 100 60
TC-304 Stability Tank 200 100 60
TC-305 Mass Balance Tank 1 200 100 60
TC-306 Line from Stability Tank to Light 200 100 60
Products Outlet
TC-307 Line from Mass Balance Tank 2 to 200 100 60
Light Products Outlet
TC-308 Light Products Outlet 200 100 60
TC-309 Intermediate Vessel 140 100 60
TC-310 Mass Balance Tank 2 140 100 60
*The set point of TC-201 is set as the water boiling temperature at the operation pressure plus 70 ˚C.
At the typical operating pressure of 300 psig for Catalytic Steam Cracking, the set point for TC-201 is 230 ˚C.
Table 3. Pilot plant pre-heating parameters.

8. Start the Hydrogen flow at a rate according to the operational conditions established in
the experimental plan.

9. Purge the water pump.

170
10. Once TC-201 has reached 150 ˚C, start pumping water at a rate according to the
operational conditions established in the experimental plan.

4.2.5. Feed Injection

1. Pressurize the plant at 75 psi using the light products Stability Tank for semi-continuous
product’s collection or Mass Balance Tank for continuous product collection, following
the pressurization/depressurization procedure on either the hot separation or cold
separation section. Remember that the BPV has to be closed to adjust the pressure at the
required set point.

2. Open the valve to allow flow to the reactor (Figure 15).

Figure 15. Feed section to reaction section valve.


3. Start pumping feed at the operational conditions established in the experimental plan.

4. If the product collection is to be carried out in continuous mode, start pneumatic valves
cycling and Nitrogen flow through hot separator following the guidelines in Annex B.

5. Simultaneously increase the temperature according to the following Table 4, using the
ramps previously established.

171
Set Point Temperature [˚C]
Controller Hydrogen
Steam
DAO/VR VGO

TC-202 Line to Furnace 300


TC-203 Furnace Trxn-20 (up to 400)
TC-204 Reactor Bottom Section 400
TC-205 Reactor Middle Section 400
TC-206 Reactor Top Section 400
TC-301 Products Outlet 220
TC-302 Line to Stability Tank HS1 220
TC-303 Line to Mass Balance Tank HS2 220
TC-304 Stability Tank* 250
TC-305 Mass Balance Tank 1* 250 N/A
TC-306 Line from Stability Tank to Light 220
Products Outlet
TC-307 Line from Mass Balance Tank 1 to 220
Light Products Outlet
TC-308 Light Products Outlet 220
*This temperature changes depending on the operating pressure of the unit.
It has to be set to guarantee separation of water from the oil.
Table 4. Pilot plant heating parameters.

4.2.6. Operating Conditions Adjustment

1. When the reactor temperature is 300 ˚C close the BPV and increase the plant pressure
with Nitrogen up to the pressure indicated in the experimental plan using the
pressurization procedure explained previously. The Nitrogen injection for pressurization
should be done from the light section in order to avoid heavy product carryover to the
light section.

2. Adjust the back-pressure valve (BPV) in order to maintain a constant pressure in the
unit.

172
3. Once reached 400 ˚C in the reactor, increase the reactor temperatures (TC-204/205/206)
at a pace of 1 ˚C/min until reaching the desired set point established in the experimental
plan.

4. Also, with the same pace adjust the furnace (TC-203) temperature until reaching 20 ˚C
below the reactor temperature.

4.3.Stabilization

1. When all the conditions (temperatures, pressure and flow) reach the desire set point, the
stabilization period is started. Wait the time established in the experimental plan (as a
rule of thumb, the time required will be the minimum time to pass 3.5 times the volume
of the reactor throughout the unit considering the feed flow at pumping conditions).

2. If the products collection is in semi-continuous mode, approximately 20 min before the


stabilization period is finished proceed to take a gas flow measurement and carry out a
Gas Chromatography (GC) according to the procedure explained below. If the
collection is in continuous mode, gas flow measurement is and GC analysis is
conducted during the mass balance or as needed between the cycles of the pneumatic
valves.

4.4.Gas Flow Measurement and Gas Chromatography (GC) Analysis


4.4.1. Gas flow measurement
1. Align the valve for gas stream (Figure 16) to the Vent/WTM section.

Figure 16. Gas stream valve for GC WTM/Vent.


2. Align the Vent/WTM selection (Figure 17) valve to the WTM section.

173
Figure 17. WTM/Vent selection valve.
3. Measure the gas flow during a pre-established time using the Wet Test Meter shown in

Figure 18. To do so, read the initial volume (Vi) in the equipment and immediately start

timing. After a pre-established period of time, the volume of gas in the equipment is

measured again (Vf). The difference between the final and initial volume (Vf-Vi)

corresponds to the volume of gas, which divided by time gives the estimation of the gas

flow rate. Note: It is very important to have a constant pressure in the unit, otherwise the

measured volume is not representative of the steady state flow.

Figure 18. Wet Test Meter.

174
4.4.2. Gas Chromatography (GC) Analysis

In order to carry out the GC analysis, align the gas stream selection vale in Figure 16 to GC,
let the gas flow for a couple of minutes through the GC in order to purge any contamination and
then start the test by pressing the space bar at the GC controlling computer.

4.5.Mass Balance Period -Semi-continuous collection


4.5.1. Before opening the mass balance

1. Following the procedure described in the pressurization/depressurization section,


increase the pressure in the heavy products mass balance tank 1 and in the light product
mass balance tank 3, until they are 2 psig above the process condition.

2. Weight two containers, one to collect the heavy product and another one for the light
product. Both should be identified with name, operational conditions of the run, and
initial and final weight.

3. Write down the time of mass balance opening.

4.5.2. Opening the mass balance

1. Turn the heavy products 3-way valves (Figure 14) to the heavy products mass balance
tank 1, both at the same time.

2. IMMEDIATLY turn the light products 3-way valves (Figure 14) to the light products
mass balance tank 3, both at the same time.

3. While the system is working in the mass balance section, the stability tanks can be
depressurized and drained (IF NEEDED).

4. For draining the stability tanks, open the manual valves SLOWLY and drain the tanks.
Sometimes it will be necessary the use of a heat gun in order to increase the
temperature in the valve to melt the product and allow the draining.

5. Once the tanks are drained, close the respective valve for sample collection.

6. Proceed to pressurize the stability tanks following once again the


pressurization/depressurization procedure, until the pressure is 2 psig above the mass
balance section.

175
4.5.3. Closing the mass balance

1. Turn the heavy products 3-way valves (Figure 14) to the heavy products stability tank,
both at the same time.

2. Immediately, turn the light products 3-way valves (Figure 14) to the light products
stability tank, both at the same time.

3. Depressurize the mass balance tanks, ONE AT A TIME, following the


pressurization/depressurization procedure.

4. SLOWLY open the drain valves and drain the mass balance tanks into the corresponding
weighted container. In the heavy products mass balance tank HS2, sometimes it will be
necessary the use of a heat gun in order to increase the temperature in the valve to melt
the product and allow the draining.

5. Drain the depressurization liquid trap (Figure 19) and pour the collected liquid into light
product mass balance container.

Figure 19. Depressurization liquid trap.


6. Weight the two containers.

4.6. Mass Balance Period -continuous collection

4.6.1. Before opening the mass balance

1. Weight two containers, one to collect the heavy product and another one for the light
product. Both should be identified with name, operational conditions of the run, and
initial and final weight.

176
2. Write down the time of mass balance opening.

4.6.2. Mass balance tanks drainage procedure


1. Make sure heavy product Mass Balance Tank 2 does not have any pressure build up. If
required, use the depressurization valve in Figure 20 to depressurize it.
2. Drain the liquid products from the heavy product Mass Balance Tank 2. Open the
manual drainage valve in Figure 20 very SLOWLY and drain the tank. When no liquid
product is coming out, flush with Nitrogen using the flushing valve to ensure that there
is not any remaining liquid product in the mass balance tank 2.

Figure 20. Mass balance collector drainage, depressurization and flushing valves for continuous
collection system – Heavy products.

3. Drain the light liquid products light products Mass Balance Tank 4. Open the manual
drainage valve in Figure 21 very SLOWLY and drain the tank.

177
Figure 21. Mass balance collector drainage for continuous collection system – Light products.

4. Drain the depressurization liquid trap (Figure 19) and pour the collected liquid into light
product container.

5. Once the tanks are drained, close the respective valves for sample collection.

6. Weight the two containers.

4.6.3. Opening the mass balance

7. Right after on cycle of the pneumatic valves of the collection system is finished and the
stability period is done, the mass balance can be started. As explained in Annex B,
timing of the pneumatic valves is done in such a way that each mass balance the
cycling takes place an exact number of times. In this way, the mass balance finishes
right after the last cycle takes place.

8. Collect all the liquid products corresponding to start-up and stability period following
the procedure explained in the previous section 4.6.2.

178
4.6.4. Closing the mass balance
9. Right after the last cycle of the pneumatic valves corresponding to the mass balance is
finished, collect all the liquid products corresponding to the pre-established mass
balance period following the procedure explained in the previous Section 4.6.2.
4.7. Shutdown

4.7.1. Conventional Shutdown

1. Decrease the temperature according to the following Table 5. This temperature can be
adjusted depending on the pressure of the run. The idea is to have a temperature at which
there is steam production that is useful for cleaning of the unit.

Set Point Temperature [˚C]


Controller Hydrogen
Steam
DAO/VR VGO

TC-201 Steam Generator/Hydrogen Heater* Tb+15 200

TC-202 Line to Furnace Tb+15 200


TC-203 Furnace Tb+15 200
TC-204 Reactor Bottom Section Tb+15 200
TC-205 Reactor Middle Section Tb+15 200
TC-206 Reactor Top Section Tb+15 200
TC-301 Products Outlet 200
TC-302 Line to Stability Tank HS1 200
TC-303 Line to Mass Balance Tank HS2 200
TC-304 Stability Tank HS1 200
TC-305 Mass Balance Tank HS2 200 N/A

TC-306 Line from Stability Tank HS1 to 200


Light Products Outlet
TC-307 Line from Mass Balance Tank HS2 200
to Light Products Outlet
TC-308 Light Products Outlet 200

Table 5. Pilot plant cooling temperatures.

179
2. When the temperature in the reactor reaches 250 ˚C, stop pumping feed.

3. Set the ISCO pump to independent mode and pump the content of the pump cylinders out
using the drain valves on each pump according to the Annex 2. Independent Mode Pump
Operation.

4. Set the water pump at 2 cc/min in order to increase the steam flowing through the plant.

5. Wait at least 20 minutes.

6. Load VGO into the auxiliary feed tank for cleaning. Make sure the tank is depressurized
before opening it. If needed, depressurize it using the depressurization valve in Figure 22.

Figure 22. Auxiliary feed tank configuration.

7. Once loaded, close the auxiliary feed tank and pressurize it using the pressurization valve
in Figure 22.

8. Close the valve to allow flow from the main feedstock tank to the pumps and open the
valve from the auxiliary tank (Figure 9).

9. Refill any of the pump cylinders and start pumping VGO at a rate close to 5 cc/min flor
cleaning.

180
10. Close the feed tank pressurization valve, depressurize the tank using the depressurization
valve until reaching 10 psig, and then drain all the remaining feedstock. Store it in an
appropriate container for reuse.

11. Depressurize the plant following the pressurization/depressurization procedure and using
the heavy section depressurization system to avoid carrying over of dirt to the light
products section.

12. Open the back-pressure valve (BPV).

13. Stop pumping water and/or injecting Hydrogen.

14. Stop pumping cleaning VGO.

15. Drain all the collector tanks opening VERY SLOWLY the drainage valves.

16. Decrease the whole plant temperature to room temperature.

17. In order to avoid vacuum, make sure the drain valves in both stability and mass balance
tanks remain opened.

18. Depressurize the auxiliary feed tank opening the depressurization valve and drain any
remaining cleaning feedstock.

19. Stop the LabVIEW software by clicking the red button and then close it.

20. Turn off the plant electrical switches.

21. Close the Nitrogen and/or Hydrogen supply valve.

22. Close condenser valves and water supply valves.

23. When the whole plant is cool, make sure ALL valves in the plant are closed.

4.7.2. Emergency Shutdown

If any emergency situation is presented that is required to shut down the unit immediately,
follow the procedure below:

1. Decrease the whole plant temperature to room temperature.

2. Release the pressure of the unit following the depressurization procedure and open the
back-pressure valve (BPV).

181
3. Turn off the feed pump, water pump, and Hydrogen flow controller.

4. If it is a high risk emergency, turn off the breakers of the power source located next to the
water pump. It will de-energize all heating tapes and turn off the pumps.

5. Another way to de-energize the unit is pushing down the power lever located on the wall
behind the RTU-1 Unit.

5. Roles & Responsibilities

Key Personnel

Main operator and responsible of the system:

Fredy Cabrales - Luis Alejandro Coy

Alternate operators: N/A

6. Training

It is recommended that a person to operate the RTU-1 be educated or pursuing


studies in the area of chemical or petroleum engineering and/or chemistry.

Any new operator of the RTU-1 has to follow an operational training conducted by
either the main operator or alternate operator. The operational training would be no
shorter than three whole experimental runs which include previous preparation of the
unit, starting up, stability condition, mass balances and shutting down of the plant.

As per the pre-requisites of entering the lab, The University of Calgary Generic
WHIMS course must be successfully completed.

7. Monitoring Requirements

Not applicable

8. Record Management

Each SOP shall be reviewed within 12 months of the date of issuance or date of last,
review to ensure the SOP is up-to-date.
This first SOP for the RTU-1 unit was submitted on September, 2013.

182
9. References

-STANDARD OPERATING PROCEDURE FOR: DAO Soaker-Stripper Unit by Gustavo


Trujillo, David Gutierrez (2012).
-STANDARD OPERATING PROCEDURE FOR: Reactivity Test Unit 1 by Gustavo Trujillo
(2008).

10. Definitions

Other abbreviations as stated in the SOP.

11. Emergency:

Contact 911

University Emergency: 403 220-5333

Immediately supervisor: Luis Alejandro Coy 403-210-9781 / 3956

Principal Investigator: Dr. Pedro Pereira-Almao 403- 220-4799

12. University Notification

Indicate if there are any concerns that EH&S should be made aware of prior to the start of a
particularly hazardous protocol, such as a disruption to a water source, loss of power, loss of gas
detection systems etc.

*Indicates elements which may be required for particularly hazardous substances or


processes.

Signed:_________________________ Date:__________________________

Principal Investigator/Manager Signature

183
Annex A: CBU operation procedure for ISCO feed pumps.

1. Press the on switch for each of the pumps, as well as for the controller
2. Before starting the operation, at least one cylinder and 25% of the other have to be filled.
In order to do this, the pumps should be set up to work in the INDEPENDENT MODE.
To do this:
o Press the MENU button and press the MORE option with soft key A
o Press the MULTI PUMP OPTION
o Press the INDEPENDENT MODE. If the number besides the option starts to
blink, it means that this options has been selected.
o Press Previous/Return button with soft key D until the Initial menu is display.
(this menu always has as a first item the word UNIT)
3. Go to the main menu and press the MORE option with soft key A and press option 6:
VALVE. Make sure that the valve is in ACTIVE mode and go to main menu pressing the
previous button.
4. Once in independent mode we can select the REFILL option, to start filling the cylinders
o Select Pump A and set up the flow rate in option 3.
o Select Pump B and set up the flow rate in option 3.
o Go back to the Initial menu by pressing the Previous/Return button or the MENU
key.
o Go to the option SELECT PUMP, go to Pump A or B and press the REFILL
button, wait until the chosen cylinder is filled.
o Do the same for the other pump.
o NOTE: In Independent mode one pump could be refilling its cylinder while the
other pump is discharging. To avoid having void space or gas bubbles inside the
cylinder, we have establish a maximum refilling rate of 20 ml/min (for the
feedstock viscosity commonly handled). Also, both pumps SHOULD NOT refill
at the same time. The manual recommends the refilling rate to be at least twice the
pumping rate.
5. Once both pumps are filled, the continuous flow operation can be started.

184
6. Go to the main menu and hit the MORE option with soft key A and press option 6
VALVE. Make sure that the valve is in ACTIVE mode. Then, go to main menu by
pressing the Previous button.
7. Click the MORE option, and got to the MULTI PUMP in this case, the CONTINUOS
FLOW MODE has to be selected. Make sure that NORMAL PRESSURE is display.
After the CONTINUES FLOW is turned on, go to LIMITS MAX/MIN and set the
remaining volume to 5%.
8. Go to SELECT PUMP, and set up the flow rate for the chosen pump. It is not necessary
to set up the flowrate for the other pump in continuous mode.
9. When ready, click the run button and the operation will start.
o When starting to equilibrate the discharge pressure of the pumps, both pumps will
deliver flow at half the flow rate.
o Once the equilibration is finished, one pump will continue to release flow at the
flow rate specified, while the other will start to refill. If there is a pressure in the
discharge of both cylinders, the pump that is off-line will finish refilling, and
when is done it will set the pressure inside its cylinder to half the pressure of the
other pump while compressing the fluid.
o When the volume of the running pump is at 5% of the total volume of the
cylinder, the flow for each pump is half the set flow rate until both pumps are
stable and then the other pump starts to refill while the other one starts in the
discharge mode.
10. If necessary, the flow in the continuous mode can be changeg while the pump is running
going to the main MENU, SELECT PUMP, and simply changing the flow rate. Notice
that this rate should be at least half of the pumping rate.
11. When the operation is done. The pumps have to be set back to independent mode in order
to drain the cylinders.

Note: for the RTU-1, this procedure can be done without the manual input of some
parameters in the Digital control box of the ISCO pump, since the LabVIEW program
incorporated a small section for controlling the ISCO pumps.

185
Annex B. Product collection: semi-continuous and continuous

The RTU-1 has two different ways of collecting liquid products: semi-continuous and
continuous.

Semi-continuous product collection

In the semi-continuous mode of operation, both stability and mass balance tank 1 act as a hot
separator. During start-up and stability period, the stream coming out of the reactor is aligned to
the stability tanks in both heavy and light products section. Once everything is ready for starting
a mass balance, the stream is directed to the mass balance tanks for a given time. This redirecting
of the stream coming out of the reactor is made with a pair of manually actuated 3-way valves.
After the mass balance time is finished, the stream is aligned back to the stability tank and
sample is collected from the mass balance tank after depressurization. Then, the mass balance
tanks can be repressurized and the procedure can be repeated as much as needed to collect the
required amount of liquid product.

Continuous product collection

In this mode of operation, small amounts are drained from the hot separator through the use
of a set of two computer-controlled pneumatic valves with a small 40 mL Swagelok vessel in
between. The set of valves is placed at the bottom of the hot separator as observed in Figure 1.
The pneumatic valves are timed in such a way that when upper valve (V-301) opens,
approximately 90% of the small vessel fills up with liquid product and the rest with reaction
gases and nitrogen. The 10 % remaining volume acts as leverage for changes in the density of the
liquid product and thermal expansion at the operating temperature. This causes a small pressure
drop in that does not induce any type of instability in the experimental run. The upper valve is
left open for 60 s and then it closes. Then, after 10 s, the lower valve (V-302) is opened for 60 s
and the product is released to the atmospheric collector. After this point, the cycle starts again.
The time between the closure of V-302 and the opening of V-301 is called Repeat Time and is
calculated following the procedure in next page. This series of events is called a cycle and is
repeated an exact number of times during a mass balance. Cycle parameters are presented in
Table 1. In order to recover the pressure set point, a constant amount of nitrogen is injected at the

186
bottom of the hot separator. This stream also helps stripping water and carrying it to the cold
separator.

Figure 1. RTU-1 continuous collection configuration in heavy products section.

Event Event Duration [s] Parameter

Start of Cycle Open V-301


60 V-301 Open Time
Close V-301

Delay 10 Delay Time

Open V-302
60 V-302 Open Time
Close V-302

Waiting Time Calculation Procedure 1 Repeat Time

End of Cycle

Table 1. Cycle parameters for pneumatic valves actuation.

187
For the light products collection, a similar approach is followed. The only difference is that
the size of the vessels is smaller, as observed in Figure X. The same timing of the valves used for
the heavy products explained in detail in this section are used in the light product collection
system.

Figure 2. RTU-1 continuous collection configuration in light products section.

Repeat time calculation procedure:

1. Estimate the net volume available in the intermediate collector assuming 90% of it will

be occupied by the liquid.

𝑁𝑒𝑡 𝑉𝑜𝑙𝑢𝑚𝑒 𝑖𝑛 𝐼𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 𝑐𝑜𝑙𝑙𝑒𝑐𝑡𝑜𝑟 [𝑐𝑐] = 40 𝑚𝐿 ∗ 0.90 = 36𝑚𝐿

2. Calculate the feed flow at the operating conditions of the separator assuming a percentage

of expansion. For DAO, a 15% volumetric expansion was assumed at 250 °C (separator

temperature) and 300 psig operating pressure.

𝐹𝑒𝑒𝑑 𝐹𝑙𝑜𝑤 𝑖𝑛 𝑆𝑒𝑝𝑎𝑟𝑎𝑡𝑜𝑟 [𝑐𝑐/𝑚𝑖𝑛] = 𝐹𝑒𝑒𝑑 𝐹𝑙𝑜𝑤 ∗ [1 + (% 𝐸𝑥𝑝/100)]

3. Calculate the number of cycles per mass balance. This number has to be a rounded-up

integer.

188
𝑀𝐵 𝑇𝑖𝑚𝑒 [𝑚𝑖𝑛]
𝐶𝑦𝑐𝑙𝑒𝑠 𝑝𝑒𝑟 𝑀𝑎𝑠𝑠 𝐵𝑎𝑙𝑎𝑛𝑐𝑒 (𝐶𝑃𝑀) = 𝑅𝑂𝑈𝑁𝐷 𝑈𝑃 [ ]
36 [𝑐𝑐]
( )
𝐹𝑒𝑒𝑑 𝐹𝑙𝑜𝑤 𝑖𝑛 𝑆𝑒𝑝𝑎𝑟𝑎𝑡𝑜𝑟 [𝑐𝑐/𝑚𝑖𝑛]

4. Calculate the V-301 Repeat Time

𝑅𝑒𝑝𝑒𝑎𝑡 𝑇𝑖𝑚𝑒 [𝑠]


[𝑀𝐵 𝑇𝑖𝑚𝑒 [𝑠] − 𝐶𝑃𝑀 ∗ (𝑉 − 301 𝑂𝑝𝑒𝑛 𝑡𝑖𝑚𝑒 [𝑠] + 𝐷𝑒𝑙𝑎𝑦 𝑇𝑖𝑚𝑒[𝑠] + 𝑉 − 302 𝑂𝑝𝑒𝑛 𝑇𝑖𝑚𝑒[𝑠])]
=
𝐶𝑃𝑀

Calculation of Nitrogen flow:

The flow of nitrogen used to recover the pressure set point and also as stripping agent for
water is calculated based on the following assumptions. First, the gas phase in the pilot plant
cannot be too diluted with nitrogen because the GC analysis cannot be performed properly
considering the hydrocarbon signals would be too noisy. Based on rule of thumbs of the CBU
group, 30% reaction gases and 70% N2 can be perfectly analyzed in the GC. In this way, a target
content of 70% N2 was established. Since the amount of the produced reaction gases change
among the different evaluated conditions, a conservative 5% wt. yield of gases and an average
reaction gases density of 0.0018 g/cc at standard conditions were assumed. In this way, the flow
of nitrogen is calculated as follows:

𝑐𝑐 𝑔
70 𝐹𝑒𝑒𝑑 𝐹𝑙𝑜𝑤 [𝑚𝑖𝑛] ∗ 𝐹𝑒𝑒𝑑 𝐷𝑒𝑛𝑠𝑖𝑡𝑦 [𝑐𝑐 ] ∗ 5
𝑁𝑖𝑡𝑟𝑜𝑔𝑒𝑛 [𝑠𝑐𝑐𝑚] = [ ]
30 100

189
Appendix B – Chapter 3 Operational Data and Experimental Results
Table 1. Pilot plant data and operational conditions for different thermal cracking runs

THERMAL CRACKING RUNS


RUN DATA
C1 C2 C3 C4 C4
Mass Balance MB1 MB1 MB1 MB1 MB2
Feed Type DAO DAO DAO DAO DAO
Pressure [psig] 300 300 300 300 300
Reaction Temperature [°C] 415 420 425 423 423
LHSV [h^-1] 2 2 2 2 2
Residence Time [min] 30 30 30 30 30
Mass Balance Time [min] 60 60 60 60 60
Hydrocarbon Feed Flow [cc/min] 3.4487 3.4487 3.4487 3.4487 3.4487
Water Feed [%wt] 5 5 5 5 5
Water Feed Flow [cc/min] 0.1655 0.1655 0.1655 0.1655 0.1655
Water in Feed [%wt] 0 0 0 0 0
Total Liquid Hydrocarbon Feed [g] 193.68 193.68 193.68 193.68 193.68
Total Water Feed [g] 9.91 9.91 9.91 9.91 9.91
Total Hydrocarbon Feed [g] 193.68 193.68 193.68 193.68 193.68
Total Feed [g] 203.59 203.59 203.59 203.59 203.59
Catalyst Matrix None None None None None
Catalyst Concentration [wt.ppm] N/A N/A N/A N/A N/A
LIQUID PRODUCTS
Light Product - Light Collector [g] 10.80 12.71 15.66 13.35 13.03
Heavy Product [g] 175.70 172.37 170.91 172.67 171.53
Water in Light Product [g] 8.52 9.52 10.28 9.60 9.36
Water in Heavy Product [%] 0.00 0.00 0.00 0.00 0.00
Water in Heavy Product [g] 0.00 0.00 0.00 0.00 0.00
Total Water Collected [g] 8.52 9.52 10.28 9.60 9.36
Light Hydrocarbon Product [g] 2.28 3.19 5.38 3.75 3.67
Heavy Hydrocarbon Product [g] 175.70 172.37 170.91 172.67 171.53
Total Liquid Hydrocarbon Product [g] 177.98 175.56 176.29 176.42 175.20
% Light HC Product 1.28 1.82 3.05 2.13 2.09
% Heavy HC Product 98.72 98.18 96.95 97.87 97.91
GAS PRODUCTS
Measured Gas Flow [sccm] 111.17 109.05 146.65 148.15 148.15
Gas Hydrocarbon Product [g] 10.08 8.06 9.60 9.48 9.48
Total Hydrocarbon Product [g] 188.06 183.62 185.89 185.90 184.68
USEFUL CALCULATIONS
Overall Mass Balance [%] 96.6 94.9 96.4 96.0 95.3
Hydrocarbon Mass Balance [%] 97.1 94.8 96.0 96.0 95.4
Water Mass Balance [%] 86.0 96.0 103.7 96.9 94.4
Liquid Yield [%] 91.6 90.9 91.6 91.4 90.7
Liquid HC Yield [%] 94.6 95.6 94.8 94.9 94.9
Naphtha Yield (28 - 190 °C) [%] 3.5 3.9 4.2 4.3 4.1
Kerosene Yield (190 - 260 °C) [%] 3.1 4.6 6.0 5.9 5.3
Diesel Yield(260 - 343 °C) [%] 15.7 16.7 17.4 16.8 16.4
LVGO Yield (343 - 453 °C) [%] 26.3 26.2 25.3 25.4 26.3
HVGO Yield (453 - 560 °C) [%] 18.0 17.2 15.8 16.3 17.6
Vacuum Residue Yield (560°C+) [%] 28.0 27.0 26.1 26.3 25.0
Gas Yield [%] 5.4 4.4 5.2 5.1 5.1
Conversion HC (350 °C+) [%] 17.8 18.0 21.8 21.0 19.8

190
Table 2. Pilot plant data and operational conditions for different catalytic steam cracking runs

CATALYTIC STEAM CRACKING RUNS


RUN DATA
C5 C6 C7 C8 C9 C9
Mass Balance MB1 MB1 MB1 MB1 MB1 MB2
Run Name C5 C6 C7 C8 C9 C9
Feed Type DAO DAO DAO DAO DAO DAO
Pressure [psig] 300 300 300 300 300 300
Reaction Temperature [°C] 423 427 425 430 435 435
LHSV [h^-1] 2 2 2 2.5 2 2
Residence Time [min] 30 30 30 24 30 30
Mass Balance Time [min] 60 60 60 60 60 62
Hydrocarbon Feed Flow [cc/min] 3.4487 3.4487 3.4487 4.3108 3.4487 3.4487
Water Feed [%wt] 5 5 5 5 5 5
Water Feed Flow [cc/min] 0.1655 0.1655 0.1655 0.2112 0.1655 0.1655
Water in Feed [%wt] 0 0 0 0 0 0
Total Liquid Hydrocarbon Feed [g] 193.68 193.68 193.68 242.10 193.68 200.13
Total Water Feed [g] 9.91 9.91 9.91 12.65 9.91 10.24
Total Hydrocarbon Feed [g] 193.68 193.68 193.68 242.10 193.68 200.13
Total Feed [g] 203.59 203.59 203.59 254.75 203.59 210.38
Catalyst Matrix K-Ni
Catalyst Concentration [wt.ppm] 303/260
K-Ni
LIQUID PRODUCTS 303/260
K-Ni
Light Product - Light Collector [g] 16.17 20.84 17.23303/260
K-Ni24.20 23.34 23.90
Heavy Product [g] 170.36 165.60 172.90
303/260
K-Ni212.65 159.25 167.29
Water in Light Product [g] 9.27 10.05 8.78303/260
K-Ni12.42 9.78 10.06
Water in Heavy Product [%] 0.00 0.00 0.00303/260
0.00 0.00 0.00
Water in Heavy Product [g] 0.00 0.00 0.00 0.00 0.00 0.00
Total Water Collected [g] 9.27 10.05 8.78 12.42 9.78 10.06
Light Hydrocarbon Product [g] 6.90 10.79 8.45 11.78 13.56 13.84
Heavy Hydrocarbon Product [g] 170.36 165.60 172.90 212.65 159.25 167.29
Total Liquid Hydrocarbon Product [g] 177.26 176.39 181.35 224.43 172.81 181.13
% Light HC Product 3.89 6.12 4.66 5.25 7.85 7.64
% Heavy HC Product 96.11 93.88 95.34 94.75 92.15 92.36
GAS PRODUCTS
Measured Gas Flow [sccm] 176.70 184.97 171.22 198.61 221.66 221.66
Gas Hydrocarbon Product [g] 12.72 13.10 11.67 13.32 15.31 15.82
Total Hydrocarbon Product [g] 189.98 189.49 193.02 237.75 188.12 196.95
USEFUL CALCULATIONS
Overall Mass Balance [%] 97.9 98.0 99.1 98.2 97.2 98.4
Hydrocarbon Mass Balance [%] 98.1 97.8 99.7 98.2 97.1 98.4
Water Mass Balance [%] 93.5 101.4 88.6 98.2 98.7 98.2
Liquid Yield [%] 91.6 91.6 93.4 93.0 89.7 90.9
Liquid HC Yield [%] 93.3 93.1 94.0 94.4 91.9 92.0
Naphtha Yield (28 - 190 °C) [%] 4.5 5.8 5.8 6.1 7.3 7.7
Kerosene Yield (190 - 260 °C) [%] 5.1 6.4 6.1 6.2 8.3 8.0
Diesel Yield(260 - 343 °C) [%] 15.3 15.9 14.8 15.7 17.4 16.7
LVGO Yield (343 - 453 °C) [%] 26.7 25.9 25.4 26.3 23.9 24.1
HVGO Yield (453 - 560 °C) [%] 18.3 17.2 16.9 17.3 14.0 13.9
Vacuum Residue Yield (560°C+) [%] 23.4 22.0 24.9 22.7 21.1 21.7
Gas Yield [%] 6.7 6.9 6.0 5.6 8.1 8.0
Conversion HC (350 °C+) [%] 23.3 27.1 24.6 25.6 33.9 33.1

191
Table 3. Products characterization results for thermal cracking runs
THERMAL CRACKING RUNS
ANALYSIS FEED
C1 C2 C3 C4 C4
Saturates [%wt] 20.8 21.1 23.2 24.1 22.2 20.7
Aromatics [%wt] 55.1 35.4 37.1 46.4 47.6 49.3
SARA
Resins [%wt] 17.4 35.2 31.3 18.3 19.6 19.9
Asphaltenes -C5 [%wt] 6.7 8.3 8.4 11.2 10.6 10.1
P-Value N/A >1.15 >1.15 <1.15 1.15 1.15
API Gravity 11.2 10.7 11.0 12.0 11.7 11.2
MCR [%wt] 9.34 10.28 11.69 12.62 12.54 11.71
Viscosity @25°C [cP] 19297 3176 1864 936 1184 1060
Sediments [%wt] N/A 0.02 0.04 0.19 0.18 0.15
GAS ANALYSIS (% v/v)
H2 1.1 0.2 1.4 1.7 1.7
H2S 0.0 0.0 0.0 0.0 0.0
SO2 0.0 0.0 0.0 0.0 0.0
Methane 41.4 42.1 52.6 53.4 53.4
CO2 3.5 3.0 3.5 3.6 3.6
Ethylene 1.5 0.8 0.7 0.7 0.7
N/A
Ethane 0.0 27.2 22.7 23.5 23.5
Propylene 5.3 4.3 3.4 3.3 3.3
Propane 14.1 12.0 9.2 8.5 8.5
i-Butane/1 Butene 12.0 5.2 3.6 3.2 3.2
n-Butane 19.9 4.9 2.7 2.1 2.1
i-Pentane 1.2 0.3 0.1 0.1 0.1
SimDist (% wt. off)
0 231.10 159.8 143.8 150.3 125.6 130.5
5 294.60 252.5 241.3 235.4 227.2 233.2
10 323.30 289.6 279.5 272.4 267.9 272.3
15 348.30 311.6 303.3 297.1 295.5 299.6
20 370.10 332.7 323.1 316.1 315.4 320.3
25 391.00 352.2 343.5 336 336 340.3
30 411.50 370.1 360.9 354.2 354.6 359.2
35 428.70 389.5 379.4 371.7 372.5 377.8
40 448.30 409.3 399.4 391.5 392.5 396.5
45 469.00 425.5 417.7 411.8 412.8 415.9
50 490.90 449.6 436.3 428.6 429.5 435.2
55 512.10 475.9 464.7 456.8 458 457.2
60 536.30 502.3 492.5 486.1 486.8 481.3
65 563.00 529.5 520.5 515.8 516 508
70 589.80 560.1 551.3 548.4 547.9 538.6
75 617.10 592.4 585.3 585.2 583.9 573.6
80 644.40 627.2 622.8 625.8 623.4 612.1
85 672.10 663.3 664.7 672.2 668.2 656.6
90 699.70 702.4 712.3
95
100

192
Table 4. Products characterization results for catalytic steam cracking runs
CATALYTIC STEAM CRACKING RUNS
ANALYSIS C9 C9
C5 C6 C7 C8
MB1 MB2
Saturates [%wt] 24.7 23.2 22.3 22.0 23.3 23.5
Aromatics [%wt] 47.1 47.7 50.1 59.1 53.6 54.3
SARA
Resins [%wt] 18.2 14.4 16.1 7.2 8.6 8.5
Asphaltenes-C5 [%wt] 10.1 14.8 11.5 11.7 14.5 13.7
P-Value N/A N/A N/A N/A N/A N/A
API Gravity 10.5 11.3 10.6 11.5 12.7 12.4
MCR [%wt] 12.38 12.35 12.22 12.52 12.76 13.51
Viscosity @25°C [cP] 1181 720 1665 720 238 311
Sediments [%wt] 0.14 0.21 0.23 0.37 0.59 0.63
GAS ANALYSIS (% v/v)
H2 5.3 3.2 3.1 3.0 2.4 2.4
H2S 5.2 1.2 0.9 0.7 0.1 0.1
SO2 0.2 0.4 0.4 0.4 0.0 0.0
Methane 38.9 44.0 47.6 49.8 49.7 49.7
CO2 2.5 3.7 3.7 4.4 4.4 4.4
Ethylene 1.3 1.7 1.7 1.7 1.5 1.5
Ethane 20.1 20.5 19.8 18.7 17.4 17.4
Propylene 4.2 4.6 4.3 4.1 4.8 4.8
Propane 11.2 11.4 10.4 9.8 9.6 9.6
i-Butane/1 Butene 6.8 5.3 4.6 4.3 6.0 6.0
n-Butane 4.3 3.9 3.3 3.0 4.0 4.0
i-Pentane 0.0 0.1 0.0 0.0 0.0 0.0
SimDist (% wt. off)
0 111.7 133.7 100.9 112.5 124.2 124.6
5 237 233.3 220.2 224.8 215 215
10 280.4 272.8 268 269 252.4 253.5
15 306.9 300.2 299.3 298.4 281.3 283.1
20 327.9 320.8 321.8 319.7 303.1 305.1
25 347.5 340.2 342.7 339.5 322.2 324.4
30 365.5 358.3 361.7 357.9 340.6 343.2
35 382.7 376 380.2 375.8 358.3 360.8
40 401.2 393.9 398.9 393.4 376.2 379
45 419.4 412.7 418.3 412.3 394.3 396.4
50 438 431.2 438 430.8 414.2 416.5
55 458.9 452.1 460.4 451.8 435 437.4
60 481.5 474.5 485.2 474.8 459.6 462.2
65 506.3 499.6 512.2 500.7 487.6 490.9
70 534.6 527.5 543.8 529.3 520.2 523.9
75 566.8 561 579.6 562.8 559 564.1
80 601.6 597.7 620.1 599.3 603.7 609.3
85 641.2 640.4 667.4 642.1 656 665.3
90 689.7 694.1 695.6
95
100

193
Table 5. Pilot plant data and operational conditions for continuous collecting system tests
RUN DATA Condition1 Condition2
Mass Balance MB1 Condition1
MB2 MB3 MB1 Condition2
MB2 MB3
Date 10/22/2015
Condition1 10/24/2015
Condition2
Run Name Bam-Bam
10/22/2015
Test 1 Bam-Bam
10/24/2015
Test 2
Feed Type Bam-Bam
10/22/2015
DAOTest 1 Bam-Bam
10/24/2015
DAOTest 2
Pressure [psig] Bam-Bam
DAO
300 Test 1 Bam-Bam
DAO
300 Test 2
Reaction Temperature [°C] DAO
300
423 DAO
300
380
LHSV [h^-1] 300
423
2 0.25
300
380
Residence Time [min] 423
30
2 0.25
380
240
Mass Balance Time [min] 30
60
2 0.25
240
480
Hydrocarbon Feed Flow [cc/min] 3.4487
30
60 0.4311
240
480
Water Feed [%wt] 3.4487
60
5 0.4311
480
5
Water Feed Flow [cc/min] 3.4487
0.1778
5 0.4311
0.0222
5
Water in Feed [%wt] 0.1778
5
0 0.0222
5
0
Hydrogen Feed Flow [sccm] 0.1778
0 0.0222
0
Total Liquid Hydrocarbon Feed [g] 198.64
0 198.64
0
Total Water Feed [g] 10.65
0 196.57
10.65
0
Total Hydrocarbon Feed [g] 196.57
198.64
10.65 196.57
198.64
10.64
Total Feed [g] 196.57
209.29
10.65 196.57
209.29
10.64
Catalyst Matrix 196.57
207.22
None 196.57
207.21
None
Catalyst Concentration [wt.ppm] 207.22
None
N/A 207.21
None
N/A
LIQUID PRODUCTS None
N/A None
N/A
Light Product - Light Collector [g] 22.15 21.56
N/A 21.18 27.16 26.32
N/A 18.49
Heavy Product [g] 174.01 177.22 178.22 175.31 177.77 184.89
Water in Light Product [g] 10.15 9.93 9.92 10.95 10.57 6.89
Water in Heavy Product [%] 0.00 0.00 0.00 0.00 0.00 0.00
Water in Heavy Product [g] 0.00 0.00 0.00 0.00 0.00 0.00
Total Water Collected [g] 10.15 9.93 9.92 10.95 10.57 6.89
Light Hydrocarbon Product [g] 12.00 11.63 11.26 16.21 15.75 11.60
Heavy Hydrocarbon Product [g] 174.01 177.22 178.22 175.31 177.77 184.89
Total Liquid Hydrocarbon Product [g] 186.01 188.85 189.48 191.52 193.52 196.49
% Light HC Product 6.45 6.16 5.94 8.46 8.14 5.91
% Heavy HC Product 93.55 93.84 94.06 91.54 91.86 94.09
GAS PRODUCTS
Measured Gas Flow [sccm] 122.00 122.00 122.00 13.50 13.50 13.50
Gas Hydrocarbon Product [g] 8.50 8.50 8.50 7.53 7.53 7.53
Total Hydrocarbon Product [g] 194.51 197.35 197.98 199.05 201.04 204.02
MASS BALANCES
Overall Mass Balance [%] 97.8 99.0 99.3 100.3 101.1 100.8
Hydrocarbon Mass Balance [%] 97.9 99.3 99.7 100.2 101.2 102.7
Water Mass Balance [%] 95.3 93.2 93.1 102.9 99.4 64.8

194
Table 6. Simulated distillation results for continuous collecting system tests
Condition1 Condition2
SimDist FEED
MB1Condition1
MB2 MB3 MB1Condition2
MB2 MB3
0 185.5 104.6Condition1
87.1 117.8 162.4Condition2
164.3 138.1
5 297.7 202.7 199.1 206.7 257.8 258.2 238.3
10 367.1 245.1 243.9 247.8 303.6 303.5 286
15 450.7 278.1 277.2 280.9 341.7 341.2 325
20 510.2 308.8 308.7 311.3 377.7 376.9 362.4
25 535.6 338.8 339.3 341.5 415.8 414.6 401
30 553.0 368.8 370.1 371.8 453.5 451.9 440.3
35 567.6 401.5 403.7 404.5 487.3 485.4 477.8
40 580.6 435.1 438.2 437.5 513.9 511.9 507.4
45 594.0 469.7 473 469.7 534.6 532.7 530.2
50 607.7 499.9 502.8 497.6 551.8 549.7 548.1
55 622.3 524.9 527.3 521.2 567.3 565.2 564.4
60 637.6 545.6 547.4 541.3 581.6 579.3 578.9
65 654.1 565 566.1 560.1 596.7 594.1 594.1
70 673.0 582.9 583.5 577.1 613.1 610 610.4
75 693.0 602.3 602.3 595.1 631.1 627.5 628.5
80 712.8 624.7 623.9 615.2 650.4 646.2 647.7
85 649.3 647.7 637.6 672.8 667.3 669.9
90
95
100

195
Appendix C – Chapter 4 Operational Data and Experimental Results
Table 1. Pilot plant data and operational conditions – effect of total pressure and steam partial
pressure on converted DAO thermal Cracking
RUN DATA C1 C2 C3 C4 C5
Mass Balance MB2 MB1 MB1 MB1 MB1
Feed Type C-DAO C-DAO C-DAO C-DAO C-DAO
Pressure [psig] 300 300 300 150 225
Reaction Temperature [°C] 423 423 423 423 423
LHSV [h^-1] 2 2 2 2 2
Residence Time [min] 30 30 30 30 30
Mass Balance Time [min] 60 60 60 60 60
Hydrocarbon Feed Flow [cc/min] 3.4487 3.4487 3.4487 3.4487 3.4487
Water Feed [%wt] 7.5 5 2.5 5 5
Water Feed Flow [cc/min] 0.2667 0.1778 0.0889 0.1778 0.1778
N2 flow at rxn conditions [cc/min] 0 0 0 0 0
Total Liquid Hydrocarbon Feed [g] 198.64 198.64 198.64 198.64 198.64
Total Water Feed [g] 15.97 10.65 5.32 10.65 10.65
Total Hydrocarbon Feed [g] 198.64 198.64 198.64 198.64 198.64
Total Feed [g] 214.62 209.29 203.97 209.29 209.29
LIQUID PRODUCTS
Light Product - Light Collector [g] 30.42 24.26 14.26 24.06 25.64
Heavy Product [g] 172.55 174.00 176.27 172.13 172.40
Water in Light Product [g] 16.60 12.00 4.87 10.78 12.56
Water in Heavy Product [%] 0.00 0.00 0.00 0.00 0.00
Water in Heavy Product [g] 0.00 0.00 0.00 0.00 0.00
Total Water Collected [g] 16.60 12.00 4.87 10.78 12.56
Light Hydrocarbon Product [g] 13.82 12.26 9.39 13.27 13.08
Heavy Hydrocarbon Product [g] 172.55 174.00 176.27 172.13 172.40
Total Liquid Hydrocarbon Product [g] 186.37 186.26 185.66 185.41 185.48
% Light HC Product 7.42 6.58 5.06 7.16 7.05
% Heavy HC Product 92.58 93.42 94.94 92.84 92.95
GAS PRODUCTS
Measured Gas Flow [sccm] 133.34 131.50 110.93 105.45 140.00
Gas Hydrocarbon Product [g] 9.57 8.39 8.02 7.46 9.10
Total Hydrocarbon Product [g] 195.94 194.65 193.68 192.86 194.58
USEFUL CALCULATIONS
Overall Mass Balance [%] 99.0 98.7 97.3 97.3 99.0
Hydrocarbon Mass Balance [%] 98.6 98.0 97.5 97.1 98.0
Water Mass Balance [%] 103.9 112.7 91.5 101.3 117.9
Liquid Yield [%] 94.6 94.7 93.4 93.7 94.6
Liquid HC Yield [%] 95.1 95.7 95.9 96.1 95.3
Naphtha (28 - 190 °C) [%] 6.5 6.0 5.0 6.0 5.8
Kerosene (190 - 260 °C) [%] 7.2 7.0 7.0 7.2 6.9
Diesel (260 - 343 °C) [%] 10.5 10.6 10.5 11.1 10.5
LVGO (343 - 453 °C) [%] 12.9 13.6 13.5 13.6 12.8
HVGO (453 - 560 °C) [%] 17.4 18.5 18.8 18.2 17.5
DAO (560°C+) [%] 28.6 28.2 27.7 26.0 27.7
Asphaltenes-C5 [%] 12.0 11.7 13.4 14.0 14.2
Vacuum Residue (560°C+) [%] 40.6 40.0 41.1 40.1 41.9
Gas Yield [%] 4.9 4.3 4.1 3.9 4.7
Conversion DAO (560°C+) [%] 57.3 57.9 58.7 61.1 58.7
Conversion Residue (560 °C+) [%] 40.0 41.0 39.3 40.9 38.2

196
Table 2. Pilot plant data and operational conditions for kinetic modeling of thermal cracking at
different conditions for converted DAO
RUN DATA C10 C11 C12 C9
Mass Balance MB1 MB2 MB1 MB1
Feed Type C-DAO C-DAO C-DAO C-DAO
Pressure [psig] 300 300 300 300
Reaction Temperature [°C] 409 409 409 416
LHSV [h^-1] 2 1.5 1 2.5
Residence Time [min] 30 40 60 24
Mass Balance Time [min] 60 60 60 60
Hydrocarbon Feed Flow [cc/min] 3.4487 2.5865 1.7243 4.3108
Water Feed [%wt] 5 5 5 5
Water Feed Flow [cc/min] 0.1778 0.1333 0.0889 0.2222
Total Liquid Hydrocarbon Feed [g] 198.64 148.98 99.32 248.30
Total Water Feed [g] 10.65 7.98 5.32 13.31
Total Hydrocarbon Feed [g] 198.64 148.98 99.32 248.30
Total Feed [g] 209.29 156.97 104.65 261.61
LIQUID PRODUCTS
Light Product - Light Collector [g] 15.64 14.33 11.15 21.57
Heavy Product [g] 186.94 136.87 86.57 226.58
Water in Light Product [g] 10.31 9.07 5.91 13.61
Water in Heavy Product [%] 0.00 0.00 0.00 0.00
Water in Heavy Product [g] 0.00 0.00 0.00 0.00
Total Water Collected [g] 10.31 9.07 5.91 13.61
Light Hydrocarbon Product [g] 5.34 5.27 5.24 7.97
Heavy Hydrocarbon Product [g] 186.94 136.87 86.57 226.58
Total Liquid Hydrocarbon Product [g] 192.27 142.14 91.81 234.54
% Light HC Product 2.77 3.71 5.71 3.40
% Heavy HC Product 97.23 96.29 94.29 96.60
GAS PRODUCTS
Measured Gas Flow [sccm] 110.41 80.00 62.57 113.40
Gas Hydrocarbon Product [g] 7.51 6.18 4.00 8.60
Total Hydrocarbon Product [g] 199.78 148.32 95.80 243.14
USEFUL CALCULATIONS
Overall Mass Balance [%] 100.4 100.3 97.2 98.1
Hydrocarbon Mass Balance [%] 100.6 99.6 96.5 97.9
Water Mass Balance [%] 96.8 113.6 111.0 102.2
Liquid Yield [%] 96.8 96.3 93.4 94.9
Liquid HC Yield [%] 96.2 95.8 95.8 96.5
Naphtha (28 - 190 °C) [%] 3.1 3.3 4.3 4.1
Kerosene (190 - 260 °C) [%] 5.2 5.4 6.7 5.8
Diesel (260 - 343 °C) [%] 9.2 9.4 10.7 9.7
LVGO (343 - 453 °C) [%] 12.1 12.4 13.7 12.5
HVGO (453 - 560 °C) [%] 19.0 19.1 19.2 18.8
DAO (560°C+) [%] 40.6 37.2 29.9 35.5
Asphaltenes-C5 [%] 7.1 9.0 11.3 10.1
Vacuum Residue (560°C+) [%] 47.7 46.2 41.3 45.6
Gas [%] 3.8 4.2 4.2 3.5
Conversion DAO (560°C+) [%] 39.3 44.5 55.3 47.0
Conversion Residue (560 °C+) [%] 29.6 31.8 39.1 32.7

197
Table 3. Pilot plant data and operational conditions for kinetic modeling of thermal cracking at
different conditions for converted DAO
RUN DATA C7 C8 C13 C14
Mass Balance MB2 MB1 MB1 MB1
Feed Type C-DAO C-DAO C-DAO C-DAO
Pressure [psig] 300 300 300 300
Reaction Temperature [°C] 416 416 423 423
LHSV [h^-1] 2 1.5 3 2.5
Residence Time [min] 30 40 20 24
Mass Balance Time [min] 60 60 60 60
Hydrocarbon Feed Flow [cc/min] 3.4487 2.5865 5.1730 4.3108
Water Feed [%wt] 5 5 5 5
Water Feed Flow [cc/min] 0.1778 0.1333 0.2667 0.2222
Total Liquid Hydrocarbon Feed [g] 198.64 148.98 297.96 248.30
Total Water Feed [g] 10.65 7.98 15.97 13.31
Total Hydrocarbon Feed [g] 198.64 148.98 297.96 248.30
Total Feed [g] 209.29 156.97 313.94 261.61
LIQUID PRODUCTS
Light Product - Light Collector [g] 18.50 16.18 28.98 25.21
Heavy Product [g] 179.82 132.33 268.60 222.75
Water in Light Product [g] 10.59 8.20 16.64 13.66
Water in Heavy Product [%] 0.00 0.00 0.00 0.00
Water in Heavy Product [g] 0.00 0.00 0.00 0.00
Total Water Collected [g] 10.59 8.20 16.64 13.66
Light Hydrocarbon Product [g] 7.91 7.99 12.34 11.55
Heavy Hydrocarbon Product [g] 179.82 132.33 268.60 222.75
Total Liquid Hydrocarbon Product [g] 187.73 140.32 280.94 234.29
% Light HC Product 4.21 5.69 4.39 4.93
% Heavy HC Product 95.79 94.31 95.61 95.07
GAS PRODUCTS
Measured Gas Flow [sccm] 119.05 106.64 136.00 171.43
Gas Hydrocarbon Product [g] 8.15 7.89 9.15 12.85
Total Hydrocarbon Product [g] 195.89 148.21 290.09 247.14
USEFUL CALCULATIONS
Overall Mass Balance [%] 98.7 99.6 97.7 99.7
Hydrocarbon Mass Balance [%] 98.6 99.5 97.4 99.5
Water Mass Balance [%] 99.4 102.7 104.2 102.7
Liquid Yield [%] 94.8 94.6 94.8 94.8
Liquid HC Yield [%] 95.8 94.7 96.8 94.8
Naphtha (28 - 190 °C) [%] 3.9 5.3 4.4 4.6
Kerosene (190 - 260 °C) [%] 6.2 7.1 6.1 6.2
Diesel (260 - 343 °C) [%] 10.4 10.9 10.2 10.4
LVGO (343 - 453 °C) [%] 13.6 13.4 13.4 13.5
HVGO (453 - 560 °C) [%] 19.2 17.2 19.9 19.2
DAO (560°C+) [%] 30.8 27.6 31.7 29.3
Asphaltenes-C5 [%] 11.8 13.1 11.2 11.7
Vacuum Residue (560°C+) [%] 42.6 40.7 42.9 41.0
Gas [%] 4.2 5.3 3.2 5.2
Conversion DAO (560°C+) [%] 54.0 58.8 52.7 56.2
Conversion Residue (560 °C+) [%] 37.1 39.9 36.7 39.5

198
Table 4. Pilot plant data and operational conditions for kinetic modeling of thermal cracking at
different conditions for converted DAO
RUN DATA C6 C15 C17 C16
Mass Balance MB1 MB2 MB1 MB2
Feed Type C-DAO C-DAO C-DAO C-DAO
Pressure [psig] 300 300 300 300
Reaction Temperature [°C] 423 380 380 380
LHSV [h^-1] 2 0.75 0.5 0.25
Residence Time [min] 30 80 120 240
Mass Balance Time [min] 60 160 240 480
Hydrocarbon Feed Flow [cc/min] 3.4487 1.2933 0.8622 0.4311
Water Feed [%wt] 5 5 5 5
Water Feed Flow [cc/min] 0.1778 0.0667 0.0444 0.0222
Total Liquid Hydrocarbon Feed [g] 198.64 198.64 198.64 198.64
Total Water Feed [g] 10.65 10.65 10.64 10.64
Total Hydrocarbon Feed [g] 198.64 198.64 198.64 198.64
Total Feed [g] 209.29 209.30 209.28 209.28
LIQUID PRODUCTS
Light Product - Light Collector [g] 21.87 15.70 17.58 21.91
Heavy Product [g] 173.07 186.47 187.42 181.86
Water in Light Product [g] 10.75 10.60 12.05 12.73
Water in Heavy Product [%] 0.00 0.00 0.00 0.00
Water in Heavy Product [g] 0.00 0.00 0.00 0.00
Total Water Collected [g] 10.75 10.60 12.05 12.73
Light Hydrocarbon Product [g] 11.12 5.11 5.53 9.19
Heavy Hydrocarbon Product [g] 173.07 186.47 187.42 181.86
Total Liquid Hydrocarbon Product [g] 184.19 191.58 192.95 191.05
% Light HC Product 6.04 2.66 2.87 4.81
% Heavy HC Product 93.96 97.34 97.13 95.19
GAS PRODUCTS
Measured Gas Flow [sccm] 175.85 26.65 21.04 11.03
Gas Hydrocarbon Product [g] 12.26 4.83 5.72 6.00
Total Hydrocarbon Product [g] 196.45 196.41 198.67 197.04
USEFUL CALCULATIONS
Overall Mass Balance [%] 99.0 98.9 100.7 100.2
Hydrocarbon Mass Balance [%] 98.9 98.9 100.0 99.2
Water Mass Balance [%] 100.9 99.5 113.3 119.6
Liquid Yield [%] 93.1 96.6 98.0 97.4
Liquid HC Yield [%] 93.8 97.5 97.1 97.0
Naphtha (28 - 190 °C) [%] 5.6 2.0 2.8 3.9
Kerosene (190 - 260 °C) [%] 7.2 3.9 4.5 5.6
Diesel (260 - 343 °C) [%] 10.6 7.2 7.5 8.9
LVGO (343 - 453 °C) [%] 13.7 9.6 10.2 11.8
HVGO (453 - 560 °C) [%] 18.3 18.8 18.7 18.9
DAO (560°C+) [%] 25.1 52.8 46.9 39.3
Asphaltenes-C5 [%] 13.2 3.3 6.6 8.6
Vacuum Residue (560°C+) [%] 38.4 56.1 53.5 47.9
Gas [%] 6.2 2.5 2.9 3.0
Conversion DAO (560°C+) [%] 62.5 21.1 30.0 41.3
Conversion Residue (560 °C+) [%] 43.3 17.1 21.0 29.2

199
Table 5. Pilot plant data and operational conditions for kinetic modeling of thermal cracking at
different conditions for virgin DAO
RUN DATA C4 C1 C3
Mass Balance MB2 MB3 MB2
Feed Type V-DAO V-DAO V-DAO
Pressure [psig] 300 300 300
Reaction Temperature [°C] 409 409 416
LHSV [h^-1] 2 1 2.5
Residence Time [min] 30 60 24
Mass Balance Time [min] 60 120 48
Hydrocarbon Feed Flow [cc/min] 3.4487 1.7243 4.3108
Water Feed [%wt] 5 5 5
Water Feed Flow [cc/min] 0.1778 0.0889 0.2222
Total Liquid Hydrocarbon Feed [g] 198.64 198.64 198.64
Total Water Feed [g] 10.65 10.65 10.65
Total Hydrocarbon Feed [g] 198.64 198.64 198.64
Total Feed [g] 209.29 209.29 209.29
LIQUID PRODUCTS
Light Product - Light Collector [g] 12.92 17.70 16.23
Heavy Product [g] 185.29 178.44 185.03
Water in Light Product [g] 7.70 8.28 9.00
Water in Heavy Product [%] 0.00 0.00 0.00
Water in Heavy Product [g] 0.00 0.00 0.00
Total Water Collected [g] 7.70 8.28 9.00
Light Hydrocarbon Product [g] 5.22 9.41 7.23
Heavy Hydrocarbon Product [g] 185.29 178.44 185.03
Total Liquid Hydrocarbon Product [g] 190.51 187.85 192.27
% Light HC Product 2.74 5.01 3.76
% Heavy HC Product 97.26 94.99 96.24
GAS PRODUCTS
Measured Gas Flow [sccm] 54.56 40.86 72.80
Gas Hydrocarbon Product [g] 5.52 8.76 5.81
Total Hydrocarbon Product [g] 196.02 196.61 198.08
USEFUL CALCULATIONS
Overall Mass Balance [%] 97.3 97.9 98.9
Hydrocarbon Mass Balance [%] 98.7 99.0 99.7
Water Mass Balance [%] 72.3 77.8 84.5
Liquid Yield [%] 94.7 93.7 96.2
Liquid HC Yield [%] 97.2 95.5 97.1
Naphtha (28 - 190 °C) [%] 3.4 6.1 4.2
Kerosene (190 - 260 °C) [%] 3.3 5.5 3.9
Diesel (260 - 343 °C) [%] 5.1 7.6 5.7
LVGO (343 - 453 °C) [%] 8.9 11.5 9.7
HVGO (453 - 560 °C) [%] 18.0 17.3 18.2
DAO (560°C+) [%] 53.7 37.7 49.8
Asphaltenes-C5 [%] 4.7 9.8 5.5
Vacuum Residue (560°C+) [%] 58.4 47.5 55.4
Gas [%] 2.8 4.5 2.9
Conversion DAO (560°C+) [%] 37.6 56.2 42.1
Conversion Residue (560 °C+) [%] 32.7 45.2 36.2

200
Table 6. Pilot plant data and operational conditions for kinetic modeling of thermal cracking at
different conditions for virgin DAO
RUN DATA C2 C6 C5
Mass Balance MB2 MB1 MB2
Feed Type V-DAO V-DAO V-DAO
Pressure [psig] 300 300 300
Reaction Temperature [°C] 416 423 423
LHSV [h^-1] 1.5 3 2
Residence Time [min] 40 20 30
Mass Balance Time [min] 80 40 60
Hydrocarbon Feed Flow [cc/min] 2.5865 5.1730 3.4487
Water Feed [%wt] 5 5 5
Water Feed Flow [cc/min] 0.1333 0.2667 0.1778
Total Liquid Hydrocarbon Feed [g] 198.64 198.64 198.64
Total Water Feed [g] 10.64 10.65 10.65
Total Hydrocarbon Feed [g] 198.64 198.64 198.64
Total Feed [g] 209.29 209.29 209.29
LIQUID PRODUCTS
Light Product - Light Collector [g] 19.25 18.46 19.14
Heavy Product [g] 178.08 183.03 177.44
Water in Light Product [g] 8.60 9.48 7.86
Water in Heavy Product [%] 0.00 0.00 0.00
Water in Heavy Product [g] 0.00 0.00 0.00
Total Water Collected [g] 8.60 9.48 7.86
Light Hydrocarbon Product [g] 10.65 8.99 11.28
Heavy Hydrocarbon Product [g] 178.08 183.03 177.44
Total Liquid Hydrocarbon Product [g] 188.73 192.01 188.71
% Light HC Product 5.64 4.68 5.98
% Heavy HC Product 94.36 95.32 94.02
GAS PRODUCTS
Measured Gas Flow [sccm] 56.21 139.54 93.38
Gas Hydrocarbon Product [g] 7.50 9.26 9.24
Total Hydrocarbon Product [g] 196.23 201.28 197.95
USEFUL CALCULATIONS
Overall Mass Balance [%] 97.9 100.7 98.3
Hydrocarbon Mass Balance [%] 98.8 101.3 99.7
Water Mass Balance [%] 80.8 89.0 73.8
Liquid Yield [%] 94.3 96.3 93.9
Liquid HC Yield [%] 96.2 95.4 95.3
Naphtha (28 - 190 °C) [%] 6.4 4.9 6.8
Kerosene (190 - 260 °C) [%] 5.6 4.4 5.9
Diesel (260 - 343 °C) [%] 7.7 6.4 8.3
LVGO (343 - 453 °C) [%] 11.6 10.4 12.1
HVGO (453 - 560 °C) [%] 17.6 17.9 17.3
DAO (560°C+) [%] 37.5 44.6 34.7
Asphaltenes-C5 [%] 9.8 6.9 10.2
Vacuum Residue (560°C+) [%] 47.2 51.4 44.9
Gas [%] 3.8 4.6 4.7
Conversion DAO (560°C+) [%] 56.5 48.2 59.6
Conversion Residue (560 °C+) [%] 45.6 40.7 48.2

201
Table 7. Products characterization results - effect of total pressure and steam partial pressure on
converted DAO thermal Cracking

ANALYSIS C1 C2 C3 C4 C5
Saturates [%wt] 13.3 11.6 10.0 12.7 12.4
Aromatics [%wt] 67.3 69.0 67.3 64.9 65.6
SARA
Resins [%wt] 6.8 7.2 8.7 7.8 7.1
Asphaltenes- C5 [%wt] 12.6 12.3 14.0 14.6 14.9
P-Value 1.15 1.15 1.15 1.10 1.10
Specific Gravity [API] 8.9 8.6 7.7 8.6 8.4
MCR [%wt] 16.57 16.24 16.89 16.82 16.62
Viscosity @25°C [cP] 2080 2863 2899 3011 2802
Sediments [%wt] 0.27 0.24 0.24 0.25 0.23
GAS ANALYSIS (% v/v)
H2 10.0 6.6 5.3 4.7 4.1
H2S 2.1 1.0 2.3 0.3 0.5
SO2 0.0 0.0 0.0 0.0 0.0
CO2 1.3 1.4 0.7 1.8 1.6
CO2 0.0 0.0 0.0 0.0 0.0
Methane 38.2 45.8 42.2 39.5 46.2
Ethylene 5.3 9.5 4.5 11.8 11.8
Ethane 15.1 17.2 16.6 16.9 17.3
Propylene 2.9 2.6 2.5 4.0 2.8
Propane 10.1 9.0 12.9 11.0 9.3
Isobutane 1.6 0.8 1.8 1.1 0.9
1-Butene 2.4 1.2 2.4 2.9 1.6
n-Butane 5.6 3.2 5.2 3.7 2.5
i-Pentane 1.6 0.6 1.1 0.6 0.4
Pentane 3.9 1.2 2.5 1.7 1.0
SimDist (% wt. off)
0 35.1 71.8 86.6 40.2 47.1
5 164.5 172.4 186.8 173.3 174.3
10 223.6 229.8 240.1 229.3 232.3
15 264.6 270.4 279.7 269.3 272.7
20 302.8 308.6 317.2 306.1 310.9
25 340.1 344.8 354.4 340.8 347.8
30 377.9 380.8 390.3 376.4 385.7
35 417.8 419.8 431.3 415.3 428.6
40 461.4 461.2 471.0 457.4 471.3
45 497.9 495.8 503.3 494.0 505.8
50 525.8 522.6 528.8 521.8 532.6
55 549.7 546.0 551.0 545.5 555.4
60 571.2 567.1 571.1 566.9 576.0
65 593.0 587.4 591.7 587.4 597.9
70 617.8 610.0 614.7 610.5 623.2
75 647.4 636.4 641.7 637.4 652.8
80 685.7 669.3 675.7 671.0 691.9
85 709.8 717.8 712.1
90
95
100

202
Table 8. Products characterization results for kinetic modeling thermal cracking runs at different
conditions for converted DAO

ANALYSIS FEED C10 C11 C12 C9

Saturates [%wt] 4.1 8.2 8.8 10.3 8.2


Aromatics [%wt] 75.0 74.2 74.3 70.3 72.0
SARA
Resins [%wt] 20.2 10.3 7.5 7.5 9.4
Asphaltenes -C5 [%wt] 0.7 7.3 9.4 11.8 10.4
P-Value N/A 1.65 1.5 1.25 1.5
API Gravity 6.0 7.7 7.6 7.6 7.9
MCR [%wt] 12.95 14.96 15.95 17.54 15.50
Viscosity @25°C [cP] 5338.5 (60 °C) 14382 10348 5868 9303
Sediments [%wt] 0.00 0.09 0.09 0.18 0.09
GAS ANALYSIS (% v/v)
H2 21.2 15.5 23.7 13.5
H2S 1.4 1.4 1.1 0.9
SO2 0.00 1.4 0.0 0.0
CO2 3.0 1.5 2.2 2.3
CO 0.9 0.4 1.3 0.5
Methane 28.4 27.1 29.5 33.2
Ethylene 1.8 0.8 1.6 1.4
Ethane N/A 13.3 12.4 12.9 13.2
Propylene 4.5 5.8 4.0 3.9
Propane 11.8 16.6 11.4 11.6
Isobutane 1.3 2.0 1.3 2.1
1-Butene 1.9 3.2 1.4 2.4
n-Butane 5.4 7.7 6.1 8.5
i-Pentane 2.5 1.7 1.0 3.7
Pentane 2.6 2.5 2.5 2.9
SimDist (% wt. off)
0 185.5 85.5 100.8 99.9 31.2
5 297.7 220.6 215.8 198.2 203.5
10 367.1 272.8 268.1 248.7 257.9
15 450.7 316.4 310.8 288.4 300.6
20 510.2 358.1 351.5 323.6 341
25 535.6 401.8 393.5 360.0 381.2
30 553.0 446.2 433.0 398.6 423.8
35 567.6 488.5 479.7 434.5 468.3
40 580.6 518.2 511.9 477.3 503.8
45 594.0 540.1 534.8 508.9 529.6
50 607.7 558.5 553.6 532.8 549.7
55 622.3 574.6 570.5 552.7 567.7
60 637.6 591.4 587.0 570.9 585
65 654.1 608.8 604.1 588.7 603.2
70 673.0 627.7 623.2 607.6 623.1
75 693.0 648.8 643.4 628.9 644.9
80 712.8 675.0 666.1 651.5 673
85 705.6 691.3 677.3 705.9
90 706.9
95
100

203
Table 9. Products characterization results for kinetic modeling of thermal cracking at different
conditions for converted DAO

ANALYSIS C7 C8 C13 C14


Saturates [%wt] 9.8 10.3 8.8 9.2
Aromatics [%wt] 70.7 69.2 71.4 70.3
SARA
Resins [%wt] 7.2 6.7 8.2 8.1
Asphaltenes- C5 [%wt] 12.3 13.9 11.5 12.3
P-Value 1.35 1.2 1.4 1.3
API Gravity 7.7 7.9 8.0 8.1
MCR [%wt] 15.41 16.64 16.20 16.61
Viscosity @25°C [cP] 6690 3666 5717 4782
Sediments [%wt] 0.11 0.19 0.11 0.20
GAS ANALYSIS (% v/v)
H2 16.0 12.0 13.0 7.8
H2S 2.1 1.2 0.8 0.6
SO2 1.6 0.8 0.00 0.0
CO2 3.1 2.0 2.3 1.5
CO 1.1 0.2 0.7 0.2
Methane 33.6 35.5 39.0 39.3
Ethylene 1.8 1.2 1.9 1.2
Ethane 13.6 13.9 14.6 15.2
Propylene 4.7 4.1 4.3 4.4
Propane 11.3 13.4 11.7 13.7
Isobutane 1.3 1.8 1.1 1.8
1-Butene 2.1 3.1 3.0 2.3
n-Butane 5.1 6.7 5.3 7.7
i-Pentane 0.7 1.5 0.6 1.3
Pentane 1.8 2.8 1.7 2.9
SimDist (% wt. off)
0 73.7 71.2 36.2 29.6
5 204.6 181.7 198.2 192.7
10 255.2 234.9 252.9 248.8
15 295.4 273.7 294.9 289.9
20 332.4 309.9 332.4 325.8
25 369.0 345.6 370.2 362.5
30 409.1 381.8 411.3 401.7
35 448.2 421.0 451.4 438.4
40 485.8 463.2 489.8 479.9
45 514.9 500.8 517.7 510.6
50 538.1 528.9 538.9 533.9
55 557.9 551.6 557.5 553.5
60 575.8 571.9 574.3 571.4
65 594.7 592.7 591.3 589.1
70 614.7 615.1 609.2 607.8
75 636.7 640.3 629.3 628.9
80 663.8 672.9 650.6 651.3
85 696.4 674.7 676.7
90 702.1 705.8
95
100

204
Table 10. Products characterization results for kinetic modeling of thermal cracking at different
conditions for converted DAO

ANALYSIS C6 C15 C17 C16


Saturates [%wt] 11.9 6.2 7.4 10.6
Aromatics [%wt] 62.6 75.6 75.1 73.1
SARA
Resins [%wt] 11.4 14.9 10.7 7.5
Asphaltenes-C5 [%wt] 14.1 3.4 6.8 8.9
P-Value 1.15 >2.0 1.9 1.35
API Gravity 8.1 7.0 7.1 8.2
MCR [%wt] 17.18 13.43 14.07 15.30
Viscosity @25°C [cP] 2802 45264 22697 4695
Sediments [%wt] 0.23 0.00 0.00 0.00
GAS ANALYSIS (% v/v)
H2 4.5 21.2 21.2 21.2
H2S 0.2 1.4 1.4 1.4
SO2 0.0 0.0 0.0 0.0
CO2 2.1 3.0 3.0 3.0
CO 0.6 0.9 0.9 0.9
Methane 46.2 28.4 28.4 28.4
Ethylene 1.8 1.8 1.8 1.8
Ethane 17.9 13.3 13.3 13.3
Propylene 4.4 4.5 4.5 4.5
Propane 12.6 11.8 11.8 11.8
Isobutane 1.2 1.3 1.3 1.3
1-Butene 1.7 1.9 1.9 1.9
n-Butane 4.8 5.4 5.4 5.4
i-Pentane 0.7 2.5 2.5 2.5
Pentane 1.5 2.6 2.6 2.6
SimDist (% wt. off)
0 54.2 125.2 109.1 93.0
5 175.9 246.6 227.5 206.4
10 230.8 307.5 289.6 262.1
15 270.2 360.1 340.9 309.0
20 307.8 415.5 390.6 352.4
25 343.3 471.9 445.7 395.9
30 377.5 509.6 491.6 443.1
35 415.0 533.8 521.3 485.2
40 456.2 551.9 542.4 516.0
45 491.6 567.7 560.2 538.8
50 519.3 582.2 575.5 558.0
55 542.7 597.3 591.6 574.8
60 563.7 613.7 608.6 592.4
65 583.4 631.6 627.5 611.3
70 605.0 650.9 647.9 632.4
75 630.3 672.7 671.7 656.0
80 659.4 698.5 699.6 686.0
85 698.8 718.9
90
95
100

205
Table 11. Products characterization results for kinetic modeling of thermal cracking at different
conditions for virgin DAO
ANALYSIS FEED C4 C1 C3
Saturates [%wt] 2.4 7.3 10.0 8.5
Aromatics [%wt] 76.0 70.5 63.8 75.0
SARA
Resins [%wt] 20.8 17.4 15.9 10.9
Asphaltenes- C5 [%wt] 0.8 4.8 10.3 5.7
P-Value N/A 1.60 1.10 1.45
API Gravity 6.30 7.7 10.1 9.4
MCR [%wt] 12.96 13.92 15.53 13.85
Viscosity @25°C [cP] 1328.00@100°C 16777 3329 9416
Sediments [%wt] 0.00 0.00 0.00 0.00
GAS ANALYSIS (% v/v)
H2 1.8 2.6 3.2
H2S 0.0 0.2 0.0
SO2 1.1 19.9 1.3
CO2 40.9 32.7 40.5
CO 8.8 7.4 8.6
Methane 0.9 0.7 0.9
Ethylene 4.8 2.2 4.2
Ethane N/A 16.0 18.0 16.3
Propylene 3.9 3.2 3.8
Propane 10.6 10.8 10.4
Isobutane 1.3 0.4 1.4
1-Butene 3.0 0.4 2.9
n-Butane 5.0 1.3 4.9
i-Pentane 0.6 0.1 0.6
Pentane 1.3 0.2 1.2
SimDist (% wt. off)
0 429.1 143.3 129.1 141.2
5 525.3 265.9 227.7 256.8
10 548.7 343.9 284.6 328.1
15 565.7 407.1 333.6 386.8
20 579.7 459.6 379.8 438.1
25 592.7 499.3 423.0 480.9
30 605.1 526.0 463.7 512.3
35 617.4 547.3 497.9 536.2
40 629.7 566.0 524.6 556.4
45 641.9 582.6 548.1 574.0
50 654.3 599.0 569.5 591.5
55 668.3 616.0 589.9 608.8
60 683.3 633.5 610.9 627.2
65 696.9 652.3 633.4 646.5
70 711.0 674.1 658.1 668.7
75 698.5 690.5 693.9
80
85
90
95
100

206
Table 12. Products characterization results for kinetic modeling of thermal cracking at different
conditions for virgin DAO
ANALYSIS C2 C6 C5
Saturates [%wt] 9.9 8.4 10.6
Aromatics [%wt] 62.6 69.0 69.9
SARA
Resins [%wt] 17.3 15.4 8.8
Asphaltenes -C5 [%wt] 10.2 7.2 10.7
P-Value 1.10 1.30 1.10
API Gravity 10.1 9.6 10.5
MCR [%wt] 15.08 14.87 15.55
Viscosity @25°C [cP] 2664 6939 1860
Sediments [%wt] 0.00 0.00 0.00
GAS ANALYSIS (% v/v)
H2 2.8 2.8 3.9
H2S 0.0 0.0 0.0
SO2 1.3 1.1 1.0
CO2 40.9 39.6 46.4
CO 7.8 6.1 4.3
Methane 0.8 1.1 0.8
Ethylene 2.3 6.2 3.7
Ethane 19.2 17.9 18.7
Propylene 3.4 4.0 3.3
Propane 11.5 10.7 10.4
Isobutane 1.5 1.3 1.3
1-Butene 2.1 2.7 2.3
n-Butane 4.8 4.8 2.2
i-Pentane 0.6 0.6 0.7
Pentane 1.1 1.2 0.9
SimDist (% wt. off)
0 127.1 140.7 126.9
5 227.1 248.9 221.9
10 284.6 313.6 274.4
15 333.1 368.3 320.6
20 379.0 417.7 364.0
25 422.0 461.4 405.7
30 462.2 497.3 445.3
35 496.4 523.5 481.2
40 522.9 545.5 511.2
45 546.2 565.2 536.5
50 567.4 583.1 559.2
55 587.3 601.0 580.0
60 607.6 619.9 601.1
65 629.4 639.8 623.9
70 653.0 661.8 648.6
75 682.8 689.6 678.9
80 714.7 717.4 713.4
85
90
95
100

207
Table 13. Microscope images from P-value analysis of converted DAO thermal cracking samples
– effect of total pressure and steam partial pressure
P-value = 1.0 P-value = 1.10 P-value = 1.15
Condition 1
423 °C
2 h-1
7.5 % wt H2O
300 psig
Condition 2
423 °C
h-1
5.0 % wt H2O
300 psig
(N2 added)
Condition 3
423 °C
2 h-1
2.5 % wt H2O
300 psig
(N2 added)
Condition 4
423 °C
2 h-1
5 % wt H2O
150 psig
Condition 5
423 °C
2 h-1
5 % wt H2O
225 psig
Condition 6
423 °C
2 h-1
5 % wt H2O
300 psig
(No N2 added)

208
Table 14. Microscope images from P-value analysis of converted DAO thermal cracking samples
at 380 and 409 °C
P-value = 1.95 P-value = 2.00 P-value = 2.05

Condition 15
380 °C
0.75 h-1

P-value = 1.80 P-value = 1.85 P-value = 1.90

Condition 17
380 °C
0.5 h-1

P-value = 1.30 P-value = 1.35 P-value = 1.40

Condition 16
380 °C
0.25 h-1

P-value = 1.60 P-value = 1.65 P-value = 1.70

Condition 10
409 °C
2 h-1

P-value = 1.4 P-value = 1.45 P-value = 1.50

Condition 11
409 °C
1.5 h-1

P-value = 1.20 P-value = 1.25 P-value = 1.30

Condition 12
409 °C
1 h-1

209
Table 15. Microscope images from P-value analysis of converted DAO thermal cracking samples
at 416 and 423 °C
P-value = 1.40 P-value = 1.45 P-value = 1.50
Condition 9
416 °C
2.5 h-1

P-value = 1.25 P-value = 1.30 P-value = 1.35

Condition 7
416 °C
2 h-1

P-value = 1.15 P-value = 1.20 P-value = 1.25

Condition 8
416 °C
1.5 h-1

P-value = 1.30 P-value = 1.35 P-value = 1.40

Condition 13
423 °C
3 h-1

P-value = 1.20 P-value = 1.25 P-value = 1.30

Condition 14
423 °C
2.5 h-1

P-value = 1.1 P-value = 1.15

Condition 6
423 °C
2 h-1

210
Table 16. Microscope images from P-value analysis of virgin DAO thermal cracking samples
P-value = 1.0 P-value = 1.10 P-value = 1.15

Condition 1
409 °C
1 h-1

P-value = 1.0 P-value = 1.1

Condition 2
416 °C
1.5 h-1

P-value = 1.40 P-value = 1.45 P-value = 1.55

Condition 3
416 °C
2.5 h-1

P-value = 1.55 P-value = 1.60 P-value = 1.65

Condition 4
409 °C
2 h-1

P-value = 1.0 P-value = 1.1

Condition 5
423 °C
2 h-1

P-value = 1.20 P-value = 1.25 P-value = 1.30

Condition 6
423 °C
3 h-1

211
Appendix D – Chapter 5 Operational Data and Experimental Results
Table 1. Gas chromatography results for water splitting test using Ni/K UDC formulation
THERMAL CATALYTIC RESULTS
T=423 T=423 T=430 T=437 T=440
Conditions:
LHSV =2h-1 LHSV =2h-1 LHSV =2h-1 LHSV =2h-1 LHSV =3 h-1
Component Thermal C1 C2 C3 C4
H2 4.4 10.0 10.0 10.8 11.7
H2S 0.2 0.2 0.1 0.1 0.1
SO2 0.0 0.0 0.0 0.0 0.0
CO2 2.1 2.0 2.8 3.4 3.8
CO 0.6 0.6 0.6 0.7 0.6
Methane 46.2 47.1 47.1 46.7 43.8
Ethylene 1.8 1.3 1.3 1.5 1.6
Ethane 17.9 16.4 17.5 18.3 20.7
Propylene 4.4 3.4 3.2 2.9 2.8
Propane 12.6 10.0 9.8 9.0 8.8
Isobutane 1.2 1.4 1.2 1.0 1.0
1-Butene 1.7 1.5 1.2 1.2 1.1
n-Butane 4.8 4.1 3.7 3.1 2.8
i-Pentane 0.7 0.8 0.6 0.5 0.4
Pentane 1.5 1.3 1.0 0.8 0.8

212
Table 2. Pilot plant data and operational conditions for kinetic modeling of catalytic steam
cracking of converted DAO at different conditions using Ni/K UDC formulation
RUN DATA C14 C8 C7 C11 C10
Mass Balance MB1 MB1 MB2 MB2 MB1
Feed Type C-DAO C-DAO C-DAO C-DAO C-DAO
Pressure [psig] 300 300 300 300 300
Reaction Temperature [°C] 435 435 435 440 440
LHSV [h^-1] 4 3.5 3 4.5 4
Residence Time [min] 15 17.14 20 13.33 15
Mass Balance Time [min] 30 34.29 40 26.67 30
Hydrocarbon Feed Flow [cc/min] 6.8973 6.0352 5.1730 7.7595 6.8973
Water Feed [%wt] 5 5 5 5 5
Water Feed Flow [cc/min] 0.3555 0.3111 0.2667 0.4000 0.3555
Total Liquid Hydrocarbon Feed [g] 196.57 196.57 196.57 196.57 196.57
Total Water Feed [g] 10.65 10.65 10.65 10.65 10.65
Total Hydrocarbon Feed [g] 196.57 196.57 196.57 196.57 196.57
Total Feed [g] 207.22 207.22 207.22 207.22 207.22
Catalyst Matrix K/Ni K/Ni K/Ni K/Ni K/Ni
Catalyst Concentration [wt.ppm] 400/300 400/300 400/300 400/300 400/300
LIQUID PRODUCTS
Light Product - Light Collector [g] 24.52 25.22 25.72 24.00 25.99
Heavy Product [g] 171.42 168.09 165.92 167.29 168.57
Water in Light Product [g] 10.75 10.39 10.59 10.59 10.29
Water in Heavy Product [%] 0.00 0.00 0.00 0.00 0.00
Water in Heavy Product [g] 0.00 0.00 0.00 0.00 0.00
Total Water Collected [g] 10.75 10.39 10.59 10.59 10.29
Light Hydrocarbon Product [g] 13.77 14.82 15.13 13.41 15.70
Heavy Hydrocarbon Product [g] 171.42 168.09 165.92 167.29 168.57
Total Liquid Hydrocarbon Product [g] 185.19 182.92 181.05 180.69 184.27
% Light HC Product 7.44 8.10 8.36 7.42 8.52
% Heavy HC Product 92.56 91.90 91.64 92.58 91.48
GAS PRODUCTS
Measured Gas Flow [sccm] 254.00 273.70 222.50 370.00 300.00
Gas Hydrocarbon Product [g] 7.38 9.92 9.40 8.09 8.83
Total Hydrocarbon Product [g] 192.57 192.84 190.44 188.78 193.10
USEFUL CALCULATIONS
Overall Mass Balance [%] 98.1 98.1 97.0 96.2 98.2
Hydrocarbon Mass Balance [%] 98.0 98.1 96.9 96.0 98.2
Water Mass Balance [%] 101.0 97.6 99.4 99.5 96.7
Liquid Yield [%] 94.6 93.3 92.5 92.3 93.9
Liquid HC Yield [%] 96.2 94.9 95.1 95.7 95.4
Naphtha (28 - 190 °C) [%] 7.8 8.6 9.0 7.4 8.2
Kerosene (190 - 260 °C) [%] 7.1 8.1 8.2 7.7 8.4
Diesel (260 - 343 °C) [%] 9.9 11.4 11.9 11.2 11.6
LVGO (343 - 453 °C) [%] 12.6 14.1 14.4 13.8 14.2
HVGO (453 - 560 °C) [%] 17.1 17.6 17.4 17.8 17.6
DAO (560°C+) [%] 29.5 21.9 19.0 24.1 20.8
Asphaltenes-C5 [%] 12.0 13.2 15.1 13.7 14.6
Vacuum Residue (560°C+) [%] 41.5 35.2 34.1 37.8 35.5
Gas [%] 3.8 5.1 4.9 4.3 4.6
Conversion DAO (560°C+) [%] 54.4 66.1 70.5 62.7 67.8
Conversion Residue (560 °C+) [%] 37.1 46.7 48.3 42.7 46.2

213
Table 3. Pilot plant data and operational conditions for kinetic modeling of catalytic steam
cracking of converted DAO at different conditions using Ni/K UDC formulation
RUN DATA C9 C15 C13 C12
Mass Balance MB1 MB1 MB1 MB1
Feed Type C-DAO C-DAO C-DAO C-DAO
Pressure [psig] 300 300 300 300
Reaction Temperature [°C] 440 445 445 445
LHSV [h^-1] 3.5 5.5 5 4.5
Residence Time [min] 17.14 10.91 12 13.33
Mass Balance Time [min] 34.29 21.82 24 26.67
Hydrocarbon Feed Flow [cc/min] 6.0352 9.4838 8.6217 7.7595
Water Feed [%wt] 5 5 5 5
Water Feed Flow [cc/min] 0.3111 0.4889 0.4444 0.4000
Total Liquid Hydrocarbon Feed [g] 196.57 195.20 196.57 196.57
Total Water Feed [g] 10.65 12.02 10.65 10.65
Total Hydrocarbon Feed [g] 196.57 195.20 196.57 196.57
Total Feed [g] 207.22 207.22 207.22 207.22
Catalyst Matrix K/Ni K/Ni K/Ni K/Ni
Catalyst Concentration [wt.ppm] 400/300 400/300 400/300 400/300
LIQUID PRODUCTS
Light Product - Light Collector [g] 28.12 25.54 26.68 27.22
Heavy Product [g] 161.35 162.32 163.41 167.79
Water in Light Product [g] 10.30 11.78 10.40 10.04
Water in Heavy Product [%] 0.00 0.00 0.00 0.00
Water in Heavy Product [g] 0.00 0.00 0.00 0.00
Total Water Collected [g] 10.30 11.78 10.40 10.04
Light Hydrocarbon Product [g] 17.82 13.76 16.28 17.18
Heavy Hydrocarbon Product [g] 161.35 162.32 163.41 167.79
Total Liquid Hydrocarbon Product [g] 179.17 176.08 179.69 184.96
% Light HC Product 9.94 7.81 9.06 9.29
% Heavy HC Product 90.06 92.19 90.94 90.71
GAS PRODUCTS
Measured Gas Flow [sccm] 395.00 380.00 428.00 442.00
Gas Hydrocarbon Product [g] 12.69 6.78 9.49 11.36
Total Hydrocarbon Product [g] 191.86 182.86 189.18 196.33
USEFUL CALCULATIONS
Overall Mass Balance [%] 97.6 93.9 96.3 99.6
Hydrocarbon Mass Balance [%] 97.6 93.7 96.2 99.9
Water Mass Balance [%] 96.7 98.0 97.7 94.3
Liquid Yield [%] 91.4 90.7 91.7 94.1
Liquid HC Yield [%] 93.4 96.3 95.0 94.2
Naphtha (28 - 190 °C) [%] 9.7 7.2 8.7 8.6
Kerosene (190 - 260 °C) [%] 8.9 7.8 8.3 8.4
Diesel (260 - 343 °C) [%] 12.2 12.0 11.4 11.6
LVGO (343 - 453 °C) [%] 14.2 14.6 13.5 13.8
HVGO (453 - 560 °C) [%] 16.4 17.9 16.6 16.2
DAO (560°C+) [%] 16.2 23.7 21.9 18.8
Asphaltenes-C5 [%] 15.8 13.0 14.7 16.8
Vacuum Residue (560°C+) [%] 32.1 36.8 36.6 35.6
Gas [%] 6.6 3.7 5.0 5.8
Conversion DAO (560°C+) [%] 74.9 63.2 66.1 70.8
Conversion Residue (560 °C+) [%] 51.4 44.2 44.6 46.1

214
Table 4. Products characterization results for kinetic modeling of catalytic steam cracking of
converted DAO at different conditions using Ni/K UDC formulation
RUN DATA FEED C14 C8 C7 C11
Saturates [%wt] 3.58 11.8 13.2 13.4 13.2
Aromatics [%wt] 76.32 67.4 64.7 63.6 65.3
SARA
Resins [%wt] 18.73 8.3 7.2 7.2 7.2
Asphaltenes-C5 [%wt] 1.37 12.5 14.0 15.9 14.3
P-Value N/A Pv>1.15 1<Pv<1.15 1<Pv<1.15 1.15
API Gravity 6.7 9.1 9.4 9.6 8.6
MCR [%wt] 14.41 17.40 18.02 18.37 18.38
Viscosity @25°C [cP] 3436 (60 °C) 1981 1499 1383 1499
Sediments [%wt] 0.00 0.17 0.35 0.43 0.56
GAS ANALYSIS (% v/v)
H2 19.6 19.4 20.3 36.9
H2S 0.1 0.3 0.3 0.3
SO2 0.0 0.0 0.0 0.0
CO2 5.2 6.1 7.4 10.2
CO 2.6 2.6 0.6 1.1
Methane 38.6 34.7 36.5 28.5
Ethylene 1.5 1.1 1.2 1.0
Ethane N/A 14.6 13.1 11.5 8.9
Propylene 3.2 3.3 3.0 2.4
Propane 8.2 9.1 7.0 5.8
Isobutane 0.8 1.2 1.2 0.6
1-Butene 1.2 1.5 1.6 1.1
n-Butane 3.3 5.2 6.1 2.4
i-Pentane 0.4 0.9 2.0 0.3
Pentane 0.8 1.6 1.2 0.6
SimDist (% wt. off)
0 123.1 72.9 71.7 64.7 71.4
5 260.8 157.9 148.9 141.7 160.3
10 342.6 208.8 199.1 195.3 212.2
15 422.1 255.2 241.3 236.3 254.3
20 497.4 297.3 277.7 272.1 292.0
25 528.4 336.0 312.3 306.2 326.0
30 547.4 375.3 345.9 338.5 360.2
35 562.8 417.3 380.6 371.6 397.4
40 575.8 461.1 417.5 407.1 437.4
45 588.8 498.6 457.8 446.4 476.7
50 601.8 527.6 493.9 483.4 509.3
55 615.5 551.7 522.8 514.7 535.4
60 629.8 573.4 546.8 540.7 557.8
65 644.4 596.4 568.6 564.0 578.5
70 660.2 623.6 590.1 586.2 601.1
75 679.9 656.9 614.3 611.4 627.2
80 701.3 708.1 641.9 640.5 656.4
85 675.7 676.9 694.9
90
95
100

215
Table 5. Products characterization results for kinetic modeling of catalytic steam cracking of
converted DAO at different conditions using Ni/K UDC formulation
RUN DATA C10 C9 C15 C13 C12
Saturates [%wt] 12.8 14.4 14.1 14.3 13.9
Aromatics [%wt] 64.5 61.5 59.3 63.1 62.7
SARA
Resins [%wt] 7.3 7.1 13.1 7.2 5.6
Asphaltenes-C5 [%wt] 15.4 17.0 13.6 15.4 17.8
P-Value 1<Pv<1.15 1<Pv<1.15 1.15 1<Pv<1.15 1<Pv<1.15
API Gravity 10.0 10.1 9.0 9.5 10.3
MCR [%wt] 18.15 18.57 17.20 17.91 18.75
Viscosity @25°C [cP] 1010 832 1034 1036 1096
Sediments [%wt] 0.35 0.61 0.35 0.48 0.51
GAS ANALYSIS (% v/v)
H2 28.7 32.0 28.8 19.9 24.6
H2S 0.4 0.5 0.7 0.1 0.2
SO2 0.0 0.0 0.0 0.0 0.0
CO2 11.0 9.0 3.3 5.9 7.3
CO 1.4 0.9 0.0 0.8 0.8
Methane 28.7 28.1 38.9 41.6 34.7
Ethylene 0.9 0.9 2.0 1.7 1.3
Ethane 10.1 9.4 11.9 15.1 12.2
Propylene 2.7 2.8 3.1 2.8 3.0
Propane 7.3 7.4 6.6 7.2 7.9
Isobutane 1.0 1.0 0.5 0.7 1.0
1-Butene 1.3 1.4 1.1 1.0 1.4
n-Butane 4.4 4.3 2.3 2.4 3.9
i-Pentane 0.7 0.7 0.3 0.3 0.5
Pentane 1.4 1.5 0.5 0.5 1.1
SimDist (% wt. off)
0 76.3 51.9 69.6 33.5 70.0
5 153.3 138.5 159.2 148.7 151.4
10 202.0 185.7 216.0 198.1 198.1
15 242.6 225.4 255.9 238.4 237.8
20 278.2 260.6 291.2 275.4 273.2
25 312.4 293.1 323.2 310.1 307.5
30 345.8 324.1 355.7 344.3 340.6
35 380.5 356.3 390.2 379.7 374.6
40 417.8 389.7 427.6 419.2 412.3
45 457.2 427.8 466.6 460.6 453.1
50 492.1 467.7 500.9 497.8 491.0
55 521.3 502.6 528.4 527.5 522.3
60 546.4 531.5 551.9 552.8 548.7
65 569.7 556.9 573.7 575.9 573.3
70 593.2 580.6 596.9 601.3 599.4
75 620.1 607.3 623.6 631.1 630.4
80 651.5 638.5 653.9 666.5 667.5
85 693.0
90
95
100

216
Table 6. Microscope images from P-value analysis of converted DAO catalytic steam cracking
samples using Ni/K UDC formulation at 435 and 440 °C
P-value = 1.0 P-value = 1.15
P-value = 1.0
(Filtered) (Filtered)

Condition 14
435 ºC
4.0 h-1

Condition 8
435 ºC
3.5 h-1

Condition 7
435 ºC
3 h-1

Condition 11
440 ºC
4.5 h-1

Condition 10
440 ºC
4.0 h-1

Condition 9
440 ºC
3.5 h-1

217
Table 7. Microscope images from P-value analysis of converted DAO catalytic steam cracking
samples using Ni/K UDC formulation at 445 °C
P-value = 1.0 P-value = 1.15
P-value = 1.0
(Filtered) (Filtered)

Condition 15
445 ºC
5.5 h-1

Condition 13
445 ºC
5.0 h-1

Condition 12
445 ºC
4.5 h-1

218
Appendix E – Chapter 6 Operational Data and Experimental Results
Table 1. Pilot plant data and operational conditions for converted DAO catalytic steam cracking
at different conditions using Ni/Ce UDC formulation

RUN DATA C1B C1 C2 C3 C4


Mass Balance MB1 MB1 MB1 MB1 MB3
Feed Type C-DAO C-DAO C-DAO C-DAO C-DAO
Pressure [psig] 300 300 300 300 300
Reaction Temperature [°C] 380 380 380 380 423
LHSV [h^-1] 0.25 0.25 0.25 0.25 2
Residence Time [min] 240 240 240 240 30
Mass Balance Time [min] 480 480 480 480 60
Hydrocarbon Feed Flow [cc/min] 0.4311 0.4311 0.4311 0.4311 3.4487
Water Feed [%wt] 5 5 5 5 5
Water Feed Flow [cc/min] 0.0222 0.0222 0.0222 0.0222 0.1778
Total Liquid Hydrocarbon Feed [g] 198.64 198.64 198.64 198.64 198.64
Total Water Feed [g] 10.64 10.64 10.64 10.64 10.65
Total Hydrocarbon Feed [g] 198.64 198.64 198.64 198.64 198.64
Total Feed [g] 209.28 209.28 209.28 209.28 209.29
Catalyst Matrix None Ni/Ce (700) ppm)
Ni/Ce Atomic Ratio N/A 2 1.5 1 1.5
LIQUID PRODUCTS
Ni/Ce
Light Product - Light Collector [g] 17.31 17.48 19.75 19.63 20.28
Heavy Product [g] 183.76 184.41 184.45 184.30 178.37
Water in Light Product [g] 9.03 9.31 10.26 9.99
(700) ppm) 9.13
Water in Heavy Product [%] 0.41 0.41 0.28 1.14 0.00
Water in Heavy Product [g] 0.75 0.76 0.52 2.10 0.00
Total Water Collected [g] 9.78 10.07 10.78Ni/Ce12.09 9.13
Light Hydrocarbon Product [g] 8.28 8.17 9.49 9.64 11.15
Heavy Hydrocarbon Product [g] 183.00 183.65 183.93 182.20
(700) ppm) 178.37
Total Liquid Hydrocarbon Product [g] 191.28 191.82 193.42 191.84 189.52
% Light HC Product 4.33 4.26 4.90 5.02 5.89
% Heavy HC Product 95.67 95.74 95.10Ni/Ce94.98 94.11
GAS PRODUCTS
Measured Gas Flow [sccm] 11.03 13.56 13.84 14.76
(700) ppm) 98.90
Gas Hydrocarbon Product [g] 5.98 7.38 7.59 8.03 6.72
Total Hydrocarbon Product [g] 197.27 199.20 201.00 199.87 196.24
USEFUL CALCULATIONS
Overall Mass Balance [%] 98.9 100.0 101.2 101.3 98.1
Hydrocarbon Mass Balance [%] 99.3 100.3 101.2 100.6 98.8
Water Mass Balance [%] 91.9 94.7 101.3 113.7 85.7
Liquid Yield [%] 96.1 96.5 97.6 97.4 94.9
Liquid HC Yield [%] 97.0 96.3 96.2 96.0 96.6
Naphtha (28 - 190 °C) [%] 4.6 4.1 4.6 4.9 6.5
Kerosene (190 - 260 °C) [%] 5.7 5.1 5.4 5.3 6.9
Diesel (260 - 343 °C) [%] 9.0 8.7 8.8 8.8 10.9
LVGO (343 - 453 °C) [%] 11.9 11.5 11.7 11.6 13.5
HVGO (453 - 560 °C) [%] 18.3 17.8 18.0 17.8 18.2
DAO (560°C+) [%] 38.9 39.8 38.5 39.7 26.5
Asphaltenes-C5 [%] 8.6 9.3 9.2 7.8 14.2
Vacuum Residue (560°C+) [%] 47.5 49.1 47.7 47.6 40.7
Gas [%] 3.0 3.7 3.8 4.0 3.4
Conversion DAO (560°C+) [%] 41.9 40.6 42.5 40.7 60.5
Conversion Residue (560 °C+) [%] 29.9 27.5 29.5 29.8 40.0

219
Table 3. Pilot plant data and operational conditions for converted DAO catalytic steam cracking
at different conditions using Ni/Ce-based FB catalyst

RUN DATA C1-T C2-T C1-C C2-C C3-C C4-C


Mass Balance MB2 MB2 MB2 MB2 MB2 MB3
Feed Type C-DAO C-DAO C-DAO C-DAO C-DAO C-DAO
Pressure [psig] 300 300 300 300 300 300
Reaction Temperature [°C] 360 380 360 380 360 380
LHSV [h^-1] 0.25 0.25 0.25 0.25 0.25 0.25
Residence Time [min] 240 240 240 240 240 240
Mass Balance Time [min] 480 480 480 480 480 480
Hydrocarbon Feed Flow [cc/min] 0.2208 0.2208 0.2208 0.2208 0.2208 0.2208
Water Feed [%wt] 5 5 5 5 5 5
Water Feed Flow [cc/min] 0.0119 0.0119 0.0119 0.0119 0.0119 0.0119
Total Liquid Hydrocarbon Feed [g] 106.00 106.00 106.00 106.00 106.00 106.00
Total Water Feed [g] 5.70 5.70 5.70 5.70 5.70 5.70
Total Hydrocarbon Feed [g] 106.00 106.00 106.00 106.00 106.00 106.00
Total Feed [g] 111.70 111.70 111.70 111.70 111.70 111.70
Catalyst Matrix N/A N/A Ni/Ce Ni/Ce Ni/Ce Ni/Ce
Ni/Ce Atomic Ratio N/A N/A 1.7 1.7 1.7 1.7
LIQUID PRODUCTS
Light Product - Light Collector [g] 6.79 10.10 9.23 12.98 8.20 12.57
Heavy Product [g] 98.54 98.54 101.87 92.97 103.01 92.59
Water in Light Product [g] 4.77 5.18 5.67 5.15 5.38 4.59
Water in Heavy Product [%] 0.00 0.00 0.00 0.00 0.00 0.00
Water in Heavy Product [g] 0.00 0.00 0.00 0.00 0.00 0.00
Total Water Collected [g] 4.77 5.18 5.67 5.15 5.38 4.59
Light Hydrocarbon Product [g] 2.02 4.92 3.55 7.83 2.82 7.98
Heavy Hydrocarbon Product [g] 98.54 98.54 101.87 92.97 103.01 92.59
Total Liquid Hydrocarbon Product [g] 100.56 103.46 105.43 100.80 105.84 100.57
% Light HC Product 2.01 4.76 3.37 7.77 2.67 7.94
% Heavy HC Product 97.99 95.24 96.63 92.23 97.33 92.06
GAS PRODUCTS
Measured Gas Flow [sccm] 4.37 8.09 7.93 11.42 4.97 14.47
Gas Hydrocarbon Product [g] 2.56 4.59 3.93 5.96 2.56 6.54
Total Hydrocarbon Product [g] 103.12 108.05 109.36 106.77 108.40 107.11
USEFUL CALCULATIONS
Overall Mass Balance [%] 96.6 101.4 103.0 100.2 101.9 100.0
Hydrocarbon Mass Balance [%] 97.3 101.9 103.2 100.7 102.3 101.0
Water Mass Balance [%] 83.6 90.9 99.5 90.3 94.3 80.4
Liquid Yield [%] 94.3 97.3 99.5 94.9 99.6 94.1
Liquid HC Yield [%] 97.5 95.8 96.4 94.4 97.6 93.9
Naphtha (28 - 190 °C) [%] 1.8 4.4 3.0 7.0 2.5 6.3
Kerosene (190 - 260 °C) [%] 3.4 5.4 4.0 6.9 4.0 6.5
Diesel (260 - 343 °C) [%] 6.5 8.9 7.5 10.4 7.4 9.8
LVGO (343 - 453 °C) [%] 8.6 11.8 10.4 13.5 9.8 12.5
HVGO (453 - 560 °C) [%] 17.9 18.5 18.9 18.0 18.6 17.2
DAO (560°C+) [%] 54.8 37.0 46.3 26.9 49.6 29.4
Asphaltenes-C5 [%] 4.6 9.7 6.4 11.8 5.8 12.2
Vacuum Residue (560°C+) [%] 59.4 46.7 52.6 38.6 55.3 41.6
Gas [%] 2.5 4.2 3.6 5.6 2.4 6.1
Conversion DAO (560°C+) [%] 18.2 44.7 30.9 59.9 26.0 56.1
Conversion Residue (560 °C+) [%] 12.3 31.0 22.3 42.9 18.3 38.6

220
Table 3. Products characterization results for converted DAO catalytic steam cracking at different
conditions using Ni/Ce UDC formulation

Analysis FEED C1B C1 C2 C3 C4


Saturates [%wt] 3.78 10.6 10.4 10.2 10.7 10.2
SARA Aromatics [%wt] 76.59 73.1 72.5 76.3 77.2 63.2
Resins [%wt] 18.89 7.5 7.5 3.9 3.9 11.9
Asphaltenes-C5[%wt] 0.74 8.9 9.6 9.6 8.1 14.7
P-Value N/A 1.35 1.35 1.35 1.35 1.15
API Gravity 6.00 8.19 7.53 8.04 8.17 8.87
MCR [%wt] 12.86 14.6 15.86 15.49 14.66 16.45
Viscosity @25°C [cP] 5338.5(60°C) 4695 6286 4747 4609 1389
Sediments [%wt] 0.00 0.11 0.08 0.11 0.12 0.20
GAS ANALYSIS (% v/v)
H2 10.1 12.0 12.0 12.3 12.3
H2S 0.3 0.4 0.5 0.4 0.4
SO2 0.4 0.5 0.6 0.5 0.5
CO2 4.1 3.9 4.4 4.7 4.7
CO 0.5 0.5 0.2 0.2 0.2
Methane 38.9 35.6 35.1 35.5 35.5
Ethylene 1.1 0.6 0.8 0.9 0.9
Ethane N/A 19.4 20.9 21.0 19.8 19.8
Propylene 2.7 2.7 1.8 2.6 2.6
Propane 14.9 15.0 15.5 15.5 15.5
Isobutane 1.4 1.4 1.3 1.3 1.3
1-Butene 0.6 0.2 0.1 0.2 0.2
n-Butane 4.2 4.8 5.0 4.5 4.5
i-Pentane 0.5 0.6 0.6 0.6 0.6
Pentane 0.8 0.9 0.9 0.9 0.9
SimDist (% wt. off)
0 185.5 141.9 147.8 146.5 147.2 128.3
5 297.7 238.9 247.5 245.3 246.6 221.8
10 367.1 289.8 298 295 296.4 266.6
15 450.7 332.3 341.6 337.7 339.2 303.6
20 510.2 373.6 384 379 380.7 338
25 535.6 417.4 427.6 422.4 423.9 372.7
30 553.0 461.3 474.2 468.3 469.8 410.1
35 567.6 498.2 508.6 504.1 505.2 449.5
40 580.6 525 533.6 529.7 530.6 485.1
45 594.0 546 553.2 549.5 550.4 514
50 607.7 564.6 570.8 567.3 568 537.6
55 622.3 581.6 588 584 584.9 558.5
60 637.6 599.5 606.3 601.6 602.5 577.8
65 654.1 619.2 626.4 621.1 622.1 598.5
70 673.0 639.8 647.5 641.3 642.4 621.9
75 693.0 663.4 671.5 664.8 665.9 648.3
80 712.8 690.9 700.9 691.7 692.8 684.8
85
90
95
100

221
Table 4. Products characterization results for converted DAO catalytic steam cracking at different
conditions using Ni/Ce-based FB catalyst

RUN DATA C1-T C2-T C1-C C2-C C3-C C4-C


Saturates [%wt] 5.7 9.5 7.5 12.9 6.9 14.0
SARA Aromatics [%wt] 74.9 68.3 70.4 68.2 76.6 66.5
Resins [%wt] 14.7 12.1 15.5 6.4 10.5 6.5
Asphaltenes-C5 [%wt] 4.7 10.1 6.6 12.5 5.9 13.0
P-Value >2.0 1.65 >2.0 1.3 >2.0 1.35
API Gravity 7.2 9.0 7.5 9.5 7.4 0.0
MCR [%wt] 13.75 14.90 14.28 16.52 14.43 16.88
Viscosity @25°C [cP] 59806 4626 13748 1304 25890 1199
Sediments [%wt] N/A N/A N/A N/A N/A N/A0
GAS ANALYSIS (% v/v)
H2 6.3 3.9 20.3 18.4 19.4 18.3
H2S 15.5 5.0 6.0 5.1 6.4 1.7
SO2 0.0 0.0 0.0 0.0 0.0 0.0
CO2 4.6 3.8 9.9 9.6 10.2 6.1
CO 0.0 0.0 0.0 0.0 0.0 0.0
Methane 36.0 43.0 33.5 32.6 32.0 43.3
Ethylene 1.2 1.3 0.4 0.4 0.5 0.7
Ethane 13.0 19.5 13.1 13.2 12.8 18.5
Propylene 3.1 1.9 0.5 1.6 0.9 1.8
Propane 9.4 11.1 9.0 9.8 8.7 1.5
Isobutane 1.4 2.0 2.0 1.8 1.8 2.0
1-Butene 3.5 2.3 0.7 1.3 1.3 0.8
n-Butane 4.9 5.1 3.9 5.0 4.9 4.7
i-Pentane 1.1 1.0 0.9 1.2 1.1 0.9
Pentane 0.0 0.0 0.0 0.0 0.0 0.0
SimDist (% wt. off)
0 168.2 140 160.3 132.5 158.4 131.1
5 276.7 242.2 265.3 226.4 264.2 231.9
10 339 293.1 321.3 272 321.1 278
15 397.9 335.5 369.8 310 372.2 316.3
20 461.4 377.1 420.1 344.7 426.9 353
25 504.9 420.2 468.5 379.3 478.2 390.3
30 530.9 462.8 504.2 415.6 512.5 429.1
35 549.9 498 528.7 452.6 535.7 467.6
40 566.5 523.6 547.8 485.8 553.8 499.9
45 581.6 544.6 564.7 513.2 569.7 525.9
50 597.1 563.6 580.2 536.5 585 547.5
55 613.6 581.3 596.2 557.2 600.7 567.8
60 631.3 600.1 613.4 576.7 617.8 587.4
65 650.3 621.5 632 597.6 635.8 608.9
70 674.9 644.9 652.3 621.3 657.1 633.1
75 699.8 676.5 680.1 648.1 682.5 660.2
80 711 705.8 685.1 706.4 696.1
85
90
95
100

222
Table 5. Microscope images from P-value analysis of converted DAO catalytic steam cracking
samples using Ni/Ce UDC formulation

P-value = 1.30 P-value = 1.35 P-value = 1.4

C1-Blank
380 °C
0.25 h-1

P-value = 1.30 P-value = 1.35 P-value = 1.4

C1(AR1.0)
380 °C
0.25 h-1

P-value = 1.30 P-value = 1.35 P-value = 1.4

C2 (AR 1.5)
380 °C
0.25 h-1

P-value = 1.30 P-value = 1.35 P-value = 1.4

C3 (AR 2.0)
380 °C
0.25 h-1

P-value = 1.0 P-value = 1.10 P-value = 1.15

C4 (AR 1.5)
423 °C
2 h-1

223
Table 6. Microscope images from P-value analysis of converted DAO catalytic steam cracking
samples using Ni/Ce-based FB catalyst

P-value = 2.10 P-value = 2.20 P-value = 2.30

C1-Blank
360 °C
0.25 h-1

P-value = 1.55 P-value = 1.60 P-value = 1.65

C2-Blank
380 °C
0.25 h-1

C1 P-value = 2.00 P-value = 2.10 P-value = 2.30


360 °C
0.25 h-1
(Reduced
Catalyst)

C2 P-value = 1.20 P-value = 1.25 P-value =1.30


380 °C
0.25 h-1
(Reduced
Catalyst)

C3 P-value =2.00
360 °C
0.25 h-1
(Reduced
Catalyst)

P-value = 1.25 P-value = 1.30 P-value = 1.35


C4
380 °C
0.25 h-1
(Oxidized
Catalyst)

224
Appendix F – Copyrights Authorization Letter
Dr. Pedro Pereira Almao
Professor Chemical and Petroleum Engineering
Catalysis for Fuels and Energy research group
Schulich School of Engineering, University of Calgary

December 9, 2016

To Whom It May Concern:

The purpose of this this letter is to grant permission to Fredy Antonio Cabrales Navarro to
publish in his doctoral thesis entitled Reactivity Study and Kinetic Modeling of Deasphalted Oil
Upgrading via Thermal and Catalytic Steam Cracking, the two manuscripts described below,
which are currently undergoing review for publication in Energy and Fuels (American Chemical
Society – ACS) Journal. As required by the Copyrights Office at the University of Calgary,
manuscripts included in the thesis that are undergoing review for publication, need permission
from co-authors.

Included manuscripts:

 Cabrales-Navarro, F. and Pereira-Almao, P. “Reactivity and Comprehensive Kinetic


Modeling of Deasphalted Vacuum Residue Thermal Cracking”.
Submitted to: Energy and Fuels journal on 30-Sep-2016.
Manuscript ID: ef-2016-02544r.
 Cabrales-Navarro, F. and Pereira-Almao, P. “Catalytic Steam Cracking of Deasphalted
Vacuum Residue using a Ni/K Ultradispersed Catalyst”.
Submitted to: Energy and Fuels journal on 13-Nov-2016.
Manuscript ID: ef-2016-030047

Sincerely,

________________________

Dr. Pedro Pereira Almao

225

You might also like