+ All Categories
Home > Documents > Conversion of Petroleum Coke to Porous Materials

Conversion of Petroleum Coke to Porous Materials

Date post: 22-Dec-2021
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
185
University of Calgary PRISM: University of Calgary's Digital Repository Graduate Studies The Vault: Electronic Theses and Dissertations 2019-12-13 Conversion of Petroleum Coke to Porous Materials Wu, Jingfeng Wu, J. (2019). Conversion of Petroleum Coke to Porous Materials (Unpublished doctoral thesis). University of Calgary, Calgary, AB. http://hdl.handle.net/1880/111352 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
Transcript
Page 1: Conversion of Petroleum Coke to Porous Materials

University of Calgary

PRISM: University of Calgary's Digital Repository

Graduate Studies The Vault: Electronic Theses and Dissertations

2019-12-13

Conversion of Petroleum Coke to Porous Materials

Wu, Jingfeng

Wu, J. (2019). Conversion of Petroleum Coke to Porous Materials (Unpublished doctoral thesis).

University of Calgary, Calgary, AB.

http://hdl.handle.net/1880/111352

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

Page 2: Conversion of Petroleum Coke to Porous Materials

UNIVERSITY OF CALGARY

Conversion of Petroleum Coke to Porous Materials

by

Jingfeng Wu

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 AND PETROLEUM ENGINEERING

CALGARY, ALBERTA

DECEMBER, 2019

© Jingfeng Wu 2019

Page 3: Conversion of Petroleum Coke to Porous Materials

ii

Abstract

Petroleum coke (petcoke) is a low value by-product from oil and gas refinery. The production of

oil sand petcoke has been continually increasing over the last 20 years. However, only 11-20%

of the petcoke produced has been utilized as site fuel. The remainder has been largely stockpiled

in northern Alberta. As oil sand petcoke contains higher carbon but a lower ash content

compared to conventional crude oil petcoke, this project was designed to prepare porous carbon

materials from petcoke.

The aim of this thesis was to develop methods to convert by-products from oil refinery (eg.

petcoke and asphaltenes) to value-added porous carbon materials. In order to combine nanoscale

pores and macroscale pores into one monolithic structure, activation was proposed to develop

micro and mesopores on oil sand petcoke as a first step. Both chemical activation (using

KOH/NaOH) and chemical steam co-activation were studied to prepare activated carbon (AC)

from petcoke. A salt template was then utilized to form macroscale pores between AC particles

for hierarchical porous carbon (HPC) preparation.

The co-activation of KOH and steam reduced the chemical agent amount without compromising

pore volume. Before steam was introduced into the system, a molten phase around petcoke

particles is presumed to be formed. A greater amount of chemical agent corresponded to a

thicker molten chemical layer, which restricted the rate of steam gasification. By lowering the

activation temperature to 500 ˚C, a 0.34 cm3/g pore volume and 800 m2/g surface area were

obtained with an AC yield of 94%. Since there was almost no carbon consumption, the pores

developed at 500 ˚C were most likely due to the opening of initial closed pores of petcoke.

Page 4: Conversion of Petroleum Coke to Porous Materials

iii

Finally, by using asphaltenes as natural binders to connect non-washed AC particles, HPC was

fabricated with multiple scale pores after washing away the salts.

The experimental results in this thesis provide feasible approaches to prepare porous materials

from petcoke and asphaltenes. A better understanding of pore development during the activation

process will help to optimize the process and control the properties of the final product.

Page 5: Conversion of Petroleum Coke to Porous Materials

iv

Acknowledgements

I would like to acknowledge my supervisor, Dr. Josephine M. Hill, for her guidance and support

throughout my graduate program. Thank you for your patience and hard work to train me to

become an independent researcher. I have gradually developed strategic experiment design and

analytical skills. I’m also grateful to the help of my supervisory committee. Dr. Maen Husein.

and Dr. Qingyue (Gemma) Lu, for their assistance and mentorship in experiment design. Also, I

would like to thank Dr. Milana Trifkovic, Dr. Kevin B Thurbide to be part of my candidacy

exam committee, and thanks to Dr. Ron Chik-Kwong Wong, Dr. Nicolas Abatzoglou for being

part of my final exam committee.

This project was funded by the Natural Sciences and Engineering Research Council of Canada

(NSERC: STPGP 447411-13) through a Strategic Grant, and Shell Canada. I am grateful for the

financial support to attend 24th Canadian Symposium on Catalysis from Faculty of Graduate

Studies, the Canadian Catalysis Foundation and the North American Catalysis Society.

I would like to thank all of the members in Laboratory for Environmental Catalytic Applications

(LECA) for their valuable advices and accompany throughout all the years as a graduate student.

My sincere thank to Dr. Stephanie Keptep, Dr. Luis Virla, Dr. Vicente Montes, Dr. Ye Xiao, Dr.

Ross Arnold, Dr. Melanie Hazlett, Sip Chen Liew and Qing Huang for their guidance and

support to my learning and research journey.

Page 6: Conversion of Petroleum Coke to Porous Materials

v

Special thanks to the advisors and mentors to motivate me to explore myself and discover my

curiosity in learning. Thanks for all your help Dr. Karen Quinn, Dr. Paul Papin, Dr. Anna-Lisa

Ciccocioppo, Dr. Ann Laverty, Dr. Stephanie Warner, Liliana Gonzalez and Renata Gordon.

I am thankful to my family and friends around me. You are my pillars to succeed in graduate

study and daily life in Canada. I’ve appreciated the support from every members in my big

family back home. I can feel your caring love though far away from hometown. An extreme

thank for my parents and boyfriend, Jundong, for your patience with me and accompany on the

phone. I’m very grateful to have all of you share my moment to grow up and experience

Canadian work and life together. Though I may not be the talent and excellent one as you think I

would be, you are always believe in me and encourage me to keep moving forward.

Page 7: Conversion of Petroleum Coke to Porous Materials

vi

Table of Contents

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

Acknowledgements ............................................................................................................ iv Table of Contents ............................................................................................................... vi List of Tables ................................................................................................................... viii List of Figures and Illustrations ......................................................................................... ix

CHAPTER ONE: INTRODUCTION ..................................................................................1

1.1 Overview ....................................................................................................................1 1.2 Objective ....................................................................................................................2

1.3 Organization of the thesis ..........................................................................................3

CHAPTER TWO: LITERATURE REVIEW ......................................................................5 2.1 Waste carbon source ..................................................................................................5 2.2 Activation .................................................................................................................10

2.3 Mixing impact ..........................................................................................................24 2.4 Hierarchical porous carbon ......................................................................................26

CHAPTER THREE: EXPERIMENTAL METHODS ......................................................35

3.1 Materials ..................................................................................................................35 3.2 Preparation of activated carbon ...............................................................................36

3.3 Mixing impact ..........................................................................................................40 3.4 Preparation of carbon foam ......................................................................................41

3.5 Characterization .......................................................................................................42 3.6 Error source ..............................................................................................................45

3.7 Repeatability of the experiments .............................................................................46

CHAPTER FOUR: THE IMPACT OF THE AMOUNT OF CHEMICAL AGENT ........49 4.1 Reduction of chemical ratios by adding steam as another source of oxygen ..........50

4.2 Impact of steam exposure time ................................................................................62 4.3 Discussion of pore development ..............................................................................69

4.4 Economic estimation of petcoke activation .............................................................74

CHAPTER FIVE: PORE DEVELOPMENT DURING CHEMICAL ACTIVATION ....76

5.1 Experimental methods – Cross-sectioning ..............................................................77 5.2 Results ......................................................................................................................77 5.3 Pore development with chemical activation of high sulfur petcoke ......................115 5.4 Summary ................................................................................................................117

CHAPTER SIX: HIERACHICAL POROUS CARBON FROM ASPHALTENES AND

PETCOKE...............................................................................................................119 6.1 Initial Experiments .................................................................................................120 6.2 Parameters in salt template method .......................................................................122

6.3 HPC preparation ....................................................................................................128

CHAPTER SEVEN: CONCLUSIONS AND RECOMMENDATIONS ........................137

Page 8: Conversion of Petroleum Coke to Porous Materials

vii

7.1 Conclusions ............................................................................................................137 7.2 Recommendations ..................................................................................................138

REFERENCE ...................................................................................................................142

APPENDIX A: MIXING EFFICIENCY ON ACTIVATED CARBON FROM

PETROLEUM COKE .............................................................................................153

APPENDIX B: PREPARATION OF ACTIVATED CARBON FROM PETCOKE .....155

APPENDIX C: LOW TEMPERATURE ACTIVATION ...............................................157

APPENDIX D: CARBON FOAM BY EMULSION TEMPLATE ................................158

APPENDIX E. REPRINT PERMISSION LETTERS .....................................................159

Page 9: Conversion of Petroleum Coke to Porous Materials

viii

List of Tables

Table 2-1. Yields and surface areas of AC prepared from various carbon precursor by KOH

activation and steam activation. Data revised from [25]. ....................................................... 7

Table 2-2. Properties of KOH activated carbon from natural materials. Revised from [41] ........ 13

Table 2-3. Standard Gibbs Energy calculations (∆G0) (for Equations 2-1 and 2-2) during

KOH and NaOH activation. .................................................................................................. 17

Table 2-4. Diffusion mechanism in porous media. [85-87] .......................................................... 19

Table 2-5 Summary of AC properties prepared from high sulfur petcoke (4-8 wt% sulfur). ...... 23

Table 2-6 Summary of AC properties prepared from low sulfur petcoke (< 1 wt% sulfur)......... 23

Table 3-1. Proximate, ultimate, and ash analysis of petcoke. ....................................................... 36

Table 3-2. Equivalent molar ratio for mass ratio of NaOH and KOH. ......................................... 37

Table 3-3. The physical properties selected AC. .......................................................................... 48

Table 4-1. Physical properties of activated carbon produced from the activation of petcoke at

800 ˚C with KOH or NaOH. ................................................................................................. 55

Table 5-1 KOH concentration after separating zirconia balls measured by titration. .................. 81

Table 5-2. Physical properties of the pure K and Na species in activation process [157, 158]. ... 85

Table 5-3. Estimated and actual pore volumes for petcoke activated with KOH. ........................ 97

Table 5-4. Estimated and actual pore volumes for petcoke activated with NaOH. ...................... 98

Table 5-5. Total pore volume and yield versus temperature for same activation time. .............. 106

Table 5-6. Surface area from N2 adsorption and CO2 adsorption. .............................................. 112

Page 10: Conversion of Petroleum Coke to Porous Materials

ix

List of Figures and Illustrations

Figure 2-1. 3D cubic model of Qingdao petcoke. ........................................................................... 9

Figure 2-2. The proposed structure of asphaltene molecule. ........................................................ 10

Figure 2-3. IR-active functional groups on an activated carbon surface. ..................................... 11

Figure 2-4. Schematic graph to illustrate diffusion in hierarchical zeolite (a porous material

with both micropores and mesoporous). ............................................................................... 19

Figure 2-5. The schematic representation of hierarchical graphitic carbon. ................................. 26

Figure 2-6. A scheme of the HPC prepared from pitch-based carbon. ......................................... 28

Figure 2-7. A scheme for ice-segregation-induced self-assembly (ISISA). ................................. 29

Figure 2-8. Synthetic pathway for the HPC monolith by using salt templating and self-

binding. ................................................................................................................................. 30

Figure 2-9. (a) Illustration of the synthesis process of the HPC; (b) SEM image of HPC with

salt template; (c) SEM image of HPC without salt template. ............................................... 32

Figure 3-1. Activation setup for converting petcoke to porous materials..................................... 39

Figure 4-1 Impact of amount of KOH on (a) total pore volume per gram of AC, (b) yield and

(c) pore volume per gram of petcoke during the activation of petcoke (800 ˚C, 30 min)

with KOH () and KOH/St () activation .......................................................................... 51

Figure 4-2 Impact of amount of NaOH on (a) total pore volume per gram of AC, (b) yield

and (c) pore volume per gram of petcoke during the activation of petcoke (800 ˚C, 30

min) with NaOH () and NaOH/St () activation ............................................................. 51

Figure 4-3. Total pore volume of KOH activated carbon at 800˚C for 0 min () and 30 min

() with KOH to carbon ratios of 0.11, 0.21, 0.43, 0.64 without steam addition. .............. 52

Figure 4-4. Cross-section of (a) raw petcoke and (b) petcoke after heated up to 800 ˚C in N2

atmosphere for 30 min. ......................................................................................................... 53

Figure 4-5. SEM analysis of 0.43KOH/St (a) before washing (b) with elemental mapping. ....... 57

Figure 4-6. Scanning electron micrographs and particle size distribution for raw petcoke and

AC activated with KOH with or without steam. ................................................................... 59

Figure 4-7. Scanning electron micrographs and particle size distribution for petcoke and AC

activated with NaOH with or without steam. ........................................................................ 60

Page 11: Conversion of Petroleum Coke to Porous Materials

x

Figure 4-8. (a) N2 isotherms and (b) pore size distributions from 2D-NLDFT-HS for petcoke

activated with KOH with/without steam at 800 ˚C for 30 min. ............................................ 62

Figure 4-9. (a) N2 isotherms and (b) pore size distributions from 2D-NLDFT-HS for petcoke

activated with NaOH and/or steam at 800 ˚C for 30 min. .................................................... 62

Figure 4-10. Impact of steam activation time on total pore volume per gram of AC() and

yield (×) at 800 ˚C with different amounts of KOH. ............................................................ 64

Figure 4-11. The influence of steam activation time on pore size distributions of petcoke

activated at 800 ˚C with different KOH to petcoke ratios. ................................................... 66

Figure 4-12. Impact of steam activation time on molar ratio H2O:Carbon consumed when 2.5

g petcoke activated in the presence of steam at 800 ˚C with different amounts of KOH. .... 67

Figure 4-13. Impact of steam activation time pore volume created by 2.5 g petcoke activated

in the presence of steam at 800 ˚C with different amounts of KOH. .................................... 67

Figure 4-14. Impact of steam activation time at 800 ˚C on the change of pore volume from

pores smaller than 1.4 nm() versus pores between 1.4-5 nm(). ..................................... 68

Figure 4-15. Impact of steam time on mesopore volume change during KOH/St activation at

800 ˚C with various chemical ratios ..................................................................................... 69

Figure 4-16. Yields obtained for the activation of petcoke at 800 ˚C for 30 min with different

amounts of KOH () or NaOH(). .................................................................................... 71

Figure 4-17. The relationship between steam time and the pore volume change in the

presence of steam at 800 ˚C with different amounts of KOH. .............................................. 72

Figure 4-18. Schematic graph for KOH activation and co-activation of petcoke. ....................... 75

Figure 5-1 Particle size change before and after ball-milling petcoke. ........................................ 78

Figure 5-2. The impact of mixing method for KOH and petcoke on pore volume (micropore

volume indicated by black bar, mesopore volume by grey bar) and yield(solid red circle)

of AC. .................................................................................................................................... 79

Figure 5-3. The impact of petcoke particle size (0-500 µm) on (a) pore volume(grey bars are

micropores, black bars are mesopores) and yield(solid red circle). ...................................... 83

Figure 5-4. The impact of mixing method for NaOH and petcoke on pore volume (micropore

volume indicated by black bar, mesopore volume by grey bar) and yield(solid red circle)

of AC prepared from NaOH to petcoke molar ratio of 0.71at 800 ˚C for 30 min. ............... 86

Figure 5-5. Mixing impact on porosity and yield with the same particle size range (0-150

µm). ....................................................................................................................................... 88

Page 12: Conversion of Petroleum Coke to Porous Materials

xi

Figure 5-6. The impact of mixing methods and amount of NaOH on pore volumes

(represented by bars) and yields (represented by solid red circles) during NaOH

activation. .............................................................................................................................. 92

Figure 5-7. Particle size distribution of raw petcoke and ball-milled samples mixed with

NaOH to petcoke molar ratios of 1.06/0.71/0.35 after washing but before activation. ........ 93

Figure 5-8. SEM images of (a) physically mixed petcoke, (c, e, g)ball-milled petcoke with

NaOH to petcoke molar ratios at 0.35/0.71/1.06, and(d, f, h) AC after NaOH activation.

(b) AC prepared by physically mixed with NaOH to petcoke at molar ratio of 0.71. .......... 94

Figure 5-9. Large particles of petcoke mixed with (a) KOH, (c) NaOH, and (e) K2CO3, and

corresponding images (b, d and f) after activation at 800 ˚C for 30 min with a chemical

to petcoke mass ratio of 2. .................................................................................................... 96

Figure 5-10. SEM images and potassium, sodium and sulfur mapping of non-washed AC

from petcoke with (a, b) 0.43KOH, and (c, d) 0.71NaOH.................................................... 99

Figure 5-11. SEM analysis of petcoke activated with KOH before wash.a-c) SEM images of

0.43KOH with different magnifications, (d) potassium mapping and (e) sulfur mapping. 100

Figure 5-12. SEM images of (a, b) raw petcoke and (c-h) AC prepared from KOH activation

at 800 ˚C for 30 min with various KOH to petcoke molar ratios of 0.21/0.43/0.64. .......... 102

Figure 5-13. The SEM images of (a, b) KOH activated petcoke at 800 ˚C for 30 min, and (c,

d) KOH steam co-activated petcoke at 800 ˚C for 60 min.................................................. 103

Figure 5-14. Pore volume (●) and yield (×) of AC prepared from petcoke with increasing

temperatures. ....................................................................................................................... 104

Figure 5-15. Pore size distributions of raw petcoke and AC (0.43KOH) prepared from

petcoke activated with increasing temperatures.................................................................. 105

Figure 5-16. XRD patterns of (a) raw petcoke and (b) AC activated with KOH to petcoke

molar ratio of 0.43 at 800 ˚C without any holding time. .................................................... 107

Figure 5-17. Evolution of the Raman spectra of (a) raw petcoke and AC (0.43KOH) activated

at (b) 500 ˚C and (c) 800 ˚C. ............................................................................................... 108

Figure 5-18. FTIR spectra (ATR detector) of raw petcoke and AC (0.43KOH) prepared with

increasing activation temperature. ...................................................................................... 110

Figure 5-19. Pore volume (●) and yield (×) of AC (0.43KOH) prepared with extended

holding time at various activation temperatures from 500 to 800 ˚C ................................. 112

Figure 5-20. The pore size distribution of AC (0.43KOH) activated at (a) 500 ˚C, (b) 600 ˚C,

(c) 700 ˚C and (d) 800 ˚C with holding time from 0-240 min. ........................................... 115

Page 13: Conversion of Petroleum Coke to Porous Materials

xii

Figure 6-1. (a) A pellet of ball-milled petcoke, asphaltene and NaCl (mass ratio is 1:1:4), and

(b) mixed powder of ball-milled petcoke and NaCl (mass ratio of 1:4) after applying a

hydraulic force of 5 MPa for 20 s. ...................................................................................... 121

Figure 6-2. Carbon materials after carbonizing (a) ball-milled petcoke, asphaltenes, NaCl ...... 122

Figure 6-3. SEM images of carbon foam prepared by salt template with ball-milled petcoke,

asphaltenes and NaCl mixing at a mass ratio of 1:1:4 in (a, c) wet ball-mill (b, d) dry

ball-mill. .............................................................................................................................. 124

Figure 6-4. SEM image of carbon foam using (a, c, e) carbon black as a carbon source, or (b,

d, f) ball-milled petcoke as a carbon source........................................................................ 126

Figure 6-5. SEM image of carbon foam by carbonizing carbon black, asphaltenes and NaCl

(a, c) at 600 ˚C for 2 h, or (b, d) 400 ˚C for 2 h .................................................................. 127

Figure 6-6. SEM image of HPC created by carbonizing 400 ˚C for 2 h (a, c, e) ball-milled

petcoke, asphaltenes and NaCl (mass ratio = 1:1:4) or (b, d, f) petcoke-derived AC

(0.43KOH/St) with asphaltenes and NaCl (mass ratio = 1:1:4). ......................................... 129

Figure 6-7. Pore size distribution of AC (●) prepared from petcoke through co-activation of

KOH and steam (0.43KOH/St), and the corresponding HPC (○) by using the same AC

as carbon precursor. ............................................................................................................ 130

Figure 6-8. Digital photos (a-c) and SEM images (d-f) with EDX mapping (g-i) of HPC

prepared from non-washed AC (0.43KOH) with 1:1 mass ratioof AC and asphaltenes. ... 132

Figure 6-9 Digital photos and SEM images of HPC prepared from non-washed AC

(0.43KOH) and asphaltene at mass ratios of 1:1 (a, d) and 1:0.1 (b, e). ............................. 133

Figure 6-10. Pore size distribution of HPC prepared from non-washed AC (0.43KOH) and

asphaltenes at mass ratios of 1:1 (● for water wash, ○ for water and acid wash) and

1:0.1( for water wash). .................................................................................................... 135

Figure 6-11. The schematic graph of pore development for HPC structure. .............................. 136

Page 14: Conversion of Petroleum Coke to Porous Materials

xiii

List of Symbols, Abbreviations and Nomenclature

Abbreviations Definition

AC Activated carbon

BM Ball-milling

DFT Density functional theory

EDX Energy-dispersive X-ray spectroscopy

FTIR Fourier-transform infrared spectroscopy

HPC Hierarchical porous carbon

Petcoke Petroleum coke

PM Physically mixing

PSD Particle size distribution

SEM Scanning Electron microscopy

XRD X-ray Diffraction

Symbols Definition

n sample size

s sample standard deviation

�̅� sample mean

𝛼 significance level

La crystallite size

ID Integrated intensity of D Raman band

IG Integrated intensity of G Raman band

Page 15: Conversion of Petroleum Coke to Porous Materials

1

Chapter One: INTRODUCTION

1.1 Overview

Over recent decades, the search for innovative methodologies to convert waste by-products from

oil refineries into value-added materials has been a significant challenge. The production of

bitumen crude oil increased to 4.6 million barrels per day in 2018 from Western Canadian

reserves [1]. Among the various types of waste by-products from oil refineries, petcoke and

asphaltenes are appealing raw materials for further exploration. More than 50% of the petcoke

was stockpiled in oil sand fields in Northern Alberta because of the relative high cost of

transportation but low utilization of petcoke either as itself or other petcoke-derived materials

[2]. First asphaltenes are generally defined as the fraction insoluble in n-heptane. More

specifically, asphaltenes have a polar and high molecular weight (250 g/mol to 2000 g/mol)

fraction in bitumen [3]. Asphaltene deposition resulting in plugging causes severe problems in

oil reservoirs. Therefore, the majority of researchers have focused on the diagnosis, prevention

and mitigation of the asphaltene issues [4], but not the reuse of asphaltene after removal from the

reservoirs.

In contrast to waste materials, such as biomass or coal, used as common carbon precursors in the

preparation of activated carbon, petcoke produce AC with high yields (50-80 % by KOH

activation) which is benefited from its high amount of fixed carbon (> 80 wt%). However, non-

porous carbon surface and releasing of sulfur during petcoke activation hindered the

investigation of pore development mechanism and further application. Among the limited

research studies, petcoke has been activated either physically with CO2/N2/steam or chemically

with NaOH/KOH to prepare porous carbon materials since the early 1990s [5-8]. A chemical to

Page 16: Conversion of Petroleum Coke to Porous Materials

2

petcoke mass ratio of 3 was normally used to develop significant porosity during petcoke

activation. However, the extensive use of chemical agents resulted in a higher production cost

and chemical waste. The co-activation of KOH and steam was attempted for petcoke activation

for the first time by Wu et al. [9] in 2005. The resulting activated carbon (AC) had a high surface

area of 2500-3000 m2/g by only applying KOH: petcoke mass ratio of two. However, the greatest

challenge during the co-activation was the low yield of 25-30%. A better understanding of pore

development during petcoke activation is required for designing AC with high porosity without

compromising yield.

Porous carbon material consisting of multiple scales of pores has been widely applied in the

fields of catalysis, adsorption, separation, energy conversion and storage [10-12]. These types of

materials are called hierarchical porous carbons (HPCs), which has become increasingly

attractive in recent decades because of their electrical conductivity, chemical stability and low

cost, in combination with various functional hierarchical levels of pore sizes spanning from

nanoscales to macroscales [10]. An investigation of HPC synthesis with waste by-products from

oil refineries is a worthwhile in the attempt to design an environmental friendly and cost-

effective approach.

1.2 Objective

The objective of this thesis is to propose methodologies to convert waste by-products from

Athabasca oil sands to porous carbon materials. Petcoke was used as a carbon precursor to

prepare AC with high microporosity, and activation experiments were performed either with

only chemical agents or co-activated with chemical and steam. The advantages of high sulfur

Page 17: Conversion of Petroleum Coke to Porous Materials

3

petcoke for developing pores at low temperature emphasized. The correlation between petcoke

structure (partially ordered) and pore development on petcoke was studied for the first time.

Asphaltenes, being another waste material from refineries, were used for the first time as binders

for the formation of carbon foam and HPC structure. There are three main topics related to

prepare porous material from these by-products in this thesis:

The impact of the amount of chemical agents (KOH or NaOH) for the activation of

petcoke

Pore development of high sulfur petcoke during chemical activation

Carbon foam and HPC structure formation by salt template with petcoke and asphaltenes

1.3 Organization of the thesis

This thesis contained seven chapters. Chapter Two included the literature review of petcoke and

asphaltenes, and the approaches of preparing AC and HPC were also summarized with

corresponding physical and chemical properties. Some parts of the literature review were

previously published by Wu el al. in Fuel Processing Technology [13]. Chapter Three introduced

the experimental procedures and specific characteristic techniques. Chapter Four discussed the

co-activation of chemical and steam for the reduction of chemical amount. Most results in

Chapter Four were previously published by Wu et al. in Fuel Processing Technology [13]. My

role in this paper was designing and performing the experiments, analyzed data, and wrote the

manuscript. Dr. Vicente Montes Jiménez performed SEM, organized the data to discuss the

mechanism of co-activation. Dr. Luis Daniel Virla Alvarado contributed with the discussion of

data analysis and the explanation of the mechanism of co-activation of petcoke. Chapter Five

Page 18: Conversion of Petroleum Coke to Porous Materials

4

reported on the low temperature activation of high sulfur petcoke. Chapter Six covered carbon

foam and HPC synthesis by using petcoke and asphaltene with a salt templating approach.

Chapter Seven outlined a summary and recommendations for further work from the project.

Page 19: Conversion of Petroleum Coke to Porous Materials

5

Chapter Two: LITERATURE REVIEW

This chapter introduces two types of waste materials from oil refineries, petroleum coke and

asphaltenes, and their potential applications in various fields. The activation of petcoke is one of

the applications described in detail in this chapter, especially focused on various mixing

approaches and activation methods. The synthesis of HPC is also introduced as another

application.

2.1 Waste carbon source

2.1.1 Petroleum coke

Petroleum coke (petcoke) is a carbon-rich solid by-product from crude oil refineries. Crude

bitumen production was 4.6 million barrels per day in 2018 from Western Canadian reserves,

and is estimated to grow up to 5.9 million barrels per day in 2035 [1]. A coker unit in oil

refineries is designed for the thermal cracking process to upgrade hydrocarbons with high value

overhead and leave solid petcoke at the bottom. Depending on the reactor selected for the coking

process, there are two different types of petcoke in the market: delayed coke produced in a semi-

batch reactor and a fluid petcoke from fluid reactor. Since the volatiles can not escape from

semi-batch reactor, delayed petcoke usually contains a higher number of volatiles (8.1 wt%) than

fluid petcoke [14, 15]. In this thesis, delayed coke was used as our activating and characterizing

material. Syncrude produces delayed coke and is based in Fort McMurray.

High grade petcoke is widely used in the metallurgical industry and the production of electrodes.

Over 80% of the petcoke produced worldwide, however, is low grade (i.e., 5-7 wt% sulfur and

heavy metals including nickel and vanadium), which prevents it from being used as a traditional

Page 20: Conversion of Petroleum Coke to Porous Materials

6

fuel due to environmental concerns [16]. Currently, the surplus petcoke is stockpiled near

refinery and upgrading facilities [17]. These stockpiles represent a potential hazard for public

health since petcoke particulates can be transported in the form of airborne dust [18, 19], and the

stockpiles occupy considerable space that could be otherwise utilized. Given these issues, there

is interest in valorizing this by-product. Previous studies have shown the potential of using

petcoke as a starting material to produce adsorbents and catalyst supports [5, 15, 20, 21]

Petcoke is a non-porous solid, which generally must be activated before being used for the

above-mentioned applications [15, 20, 22-24]. An economic evaluation of minimizing

production costs of activated carbon (AC) found that petcoke is the most promising raw material

among several carbonaceous materials including wood, used tires, carbon black, charcoal, and

lignite, its biggest challenge is its non-porous nature. The major reason is that petcoke has a

relatively high yield and surface area compared to the other carbonaceous materials (Table 2-1).

Therefore, the production cost of petcoke is the lowest (1.08 US$/kg) among the other carbon

materials (1.22-2.49 US$/kg). In addition, a higher surface area through KOH activation of

petcoke (Table 2-1) also reduces the production cost compared to steam activation.

Page 21: Conversion of Petroleum Coke to Porous Materials

7

Table 2-1. Yields and surface areas of AC prepared from various carbon precursor by KOH

activation and steam activation. Data revised from [25].

Raw material

KOH activation Steam activation

AC yield (wt%) N2 BET surface

area (m2/g)

AC yield

(wt%)

N2 BET

surface area

(m2/g)

Wood 22 800 13 800

Used tires 20 700 15 500

Petcoke 45 3000 63 1000

Carbon black 60 500 48 500

Charcoal 44 2000 45 900

Lignite 25 2200 16 800

Another challenge when utilizing petcoke is its impurities such as heavy metals and sulfur

content. The sulfur content of petcoke ranges from 0.5 wt% to 10 wt%, which strongly depends

on the specific feedstock [26]. High sulfur content is a major problem for low grade petcoke. A

recent study investigated the sulfur species of Zibo petcoke and American petcoke through X-ray

photoelectron spectroscopy. The results showed that almost half of the sulfur was thiophene, and

the smallest amount of sulfur was inorganic sulfur (< 20%). The remaining sulfur was in the

form of sulfoxide sulfur [27]. However, it should be mentioned that XPS is a surface sensitive

technique, and the bulk composition may be different. Many researchers have established both

physical and chemical approaches to removing sulfur from petcoke [26, 28, 29]. Among all the

desulfurization approaches, chemical activation with alkali metals has proved to remove over

90% of sulfur at activation temperatures lower than 827 ˚C [26, 27, 30]. Nonetheless, an abrupt

yield drop (about 10% - 40%) and the excessive amounts of chemicals involved are barriers to

considering chemical activation as a desulfurization approach in scale-up industries. However,

Page 22: Conversion of Petroleum Coke to Porous Materials

8

the porosity developed in carbonaceous materials during activation has inspired researchers to

prepare AC from petcoke.

The structure of petcoke is still a mystery in scientific research fields since petcoke is a

complicated mixture of carbon, hydrogen, nitrogen, oxygen, sulfur and other impurities [15].

Moreover, petcoke composition varies from region to region [26, 28]. As well, the challenges

with the development of analysis technology hinder progress in the structural illustration of

petcoke [31]. In the late 1990s, a group of researchers studied the microstructure of petcoke from

five different regions by using X-Ray diffraction (XRD) and electron microscopy [32].

Differences appeared in the line broadening of XRD patterns for the five different petcoke

samples from 5 different regions, which implied that the microstructure of petcoke is heavily

dependent on its region. The researchers also found the presence of turbostratic solids for certain

types of petcoke. This indicated the co-existence of crystalline graphitic carbon layers, whose

basal planes are in a random orientation. In 2010, another research group studied the

microstructure of petcoke using XRD, and similar microcrystalline and interlayer spacing were

observed [33]. With the advanced characteristic techniques of high-resolution transmission

electron micrographs (HRTEM) and selected area diffraction (SAED), the distorted crystals in

amorphous layers of carbon were observed to have a graphitic lattice fringe of 2.85 nm.

Recently, Zhong et al. investigated the structural features of Qingdao petcoke through high-

resolution transmission electron micrographs (HRTEM), X-ray powder diffraction (XRD) and

nuclear magnetic resonance (NMR) [34]. An atomistic structural representation (Figure 2-1) was

proposed by integrating experimental data with an image-guided construction strategy (Fringe3D

Page 23: Conversion of Petroleum Coke to Porous Materials

9

and Vol3D). Image analysis was first used to gather information about fringe length and

orientation degree. A ‘stack’ analysis was then performed to obtain the statistical stacking data

for all the lattice fringe micrographs. The structural constraints were determined by fringe

curvature statistical analysis through identifying the location, length, and angles of each fringe.

Figure 2-1. 3D cubic model of Qingdao petcoke. Green for carbon, red for oxygen, yellow for

sulfur, blue for nitrogen, and grey for hydrogen atoms. Reprinted from Carbon, 129, Q. Zhong,

Q. Mao, L. Zhang, J. Xiang, J. Xiao, J.P. Mathews, Structural features of Qingdao petroleum

coke from HRTEM lattice fringes: Distributions of length, orientation, stacking, curvature, and a

large-scale image-guided 3D atomistic representation, 790-802 [34], Copyright (2018), with

permission from Elsevier.

2.1.2 Asphaltenes

Asphaltenes are the heaviest fraction of crude oil. Asphaltenes are a mixture defined as toluene-

soluble, n-heptane-insoluble fraction from crude oil. The polyaromatic rings surrounded by

aliphatic tails contain heteroatoms (O, N, S), which comprise asphaltene molecules. Both the

‘continental’ and ‘archipelago’ model (Figure 2-2) have been widely reported in the literature

Page 24: Conversion of Petroleum Coke to Porous Materials

10

[35]. There is still a debate on the molecular structure of asphaltenes because of their chemical

complexity as a group of chemical materials.

Figure 2-2. The proposed structure of asphaltene molecule. A. Continental; B Archipelago.

Reprinted from Journal of Petroleum Science and Engineering, 158, K. Gharbi, K. Benyounes,

M. Khodja, Removal and prevention of asphaltene deposition during oil production: a literature

review, 351-360 [35], Copyright (2017), with permission from Elsevier.

2.2 Activation

2.2.1 Activated carbon (AC)

Activated carbon (AC), also called activated charcoal, refers to the highly porous carbonaceous

materials which are utilized as adsorbents [5, 36, 37] , catalyst supports [38-40] and electrodes

[41-43]. Depending on the pore width, AC can be classified into micropores (< 2 nm), mesopores

(2-50 nm) and macropores (>50 nm). Since activated carbon is a black solid without a distinctive

chemical formula, activated carbon could also be classified based on its shape and industrial

applications: powered activated carbon [44], granular activated carbon [45], extruded activated

carbon [46], activated cloths [47], and others. Several heteroatoms, such as oxygen, nitrogen,

hydrogen, sulfur and other elements, are attached to carbon atoms as functional groups [48, 49].

The amount of heteroatoms depends on the types of carbon precursor. For instance, carbon black

has over 99 wt% carbon content [50], however, biomass has only 40-50 wt% carbon [51]. It has

Page 25: Conversion of Petroleum Coke to Porous Materials

11

been reported that the surface oxygen groups on the carbon precursors before activation, acted as

‘active sites’ in the continuous activation process [52, 53]. A representative of IR-active

functional groups after chemical modification is shown in Figure 2-3. The functional groups

include carboxylic, lactonic, carbonyl, quinone and others.

Figure 2-3. IR-active functional groups on an activated carbon surface. (a) aromatic C=C

stretching; (b, c) carboxyl-carbonates; (d) carboxylic acid; (e, f) lactone; (g) ether bridge; (h)

cyclic ether; (i, j) cyclic anhydride; (k) quinine; (l) phenol; (m) alcohol; and (n) ketene.

2.2.2 Carbon precursor for activation

Although commercial AC has been used to purify water for 150 years, it is still challenging to

design a cost-effective and environmentally-friendly AC. Therefore, waste materials with high

carbon content have become one of a growing number of carbon precursors for AC preparation

[54]; for instance, wood [55], vegetables [36, 56], coal [57, 58] and petroleum residues [30, 59].

The type of carbon precursor determines the properties of the resulting AC, because the strength

Page 26: Conversion of Petroleum Coke to Porous Materials

12

and structure of the raw materials has been shown correlation to the physical and chemical

properties of AC product [60].

As materials from nature, such as biomass, vegetables or animal bones, are abundant, many

researchers have preferred to utilize the inherently porous structure of these materials to develop

further porosity through activation. These materials are also regarded as renewable and economic

carbon precursors for activation, because the porosity can either be developed with or without

chemical agents [61]. Table 2-2 is a summary of KOH activated carbon prepared from natural

materials. Most produced AC have surface area between 1000 and 2500 m2/g and pore volumes

above 1.2 cm3/g. However, the low yield (< 50%) is mostly neglected in the literature, even

though it is a negative factor when considering natural materials for AC preparation. Among

natural materials, the porosity and yield of AC reflects the specific raw material used. Minkova

prepared AC from five different biomass materials under the same activation conditions, and

found relative high yields with larger adsorption capacities obtained with olive wastes, birch and

bagasse compared to straw and miscanthus [62].

Page 27: Conversion of Petroleum Coke to Porous Materials

13

Table 2-2. Properties of KOH activated carbon from natural materials. Revised from [41]

Carbon precursor BET surface area (m2/g) Pore volume (cm3/g) Reference

Cherry stones 1273 __ [63]

Fish scale 2273 2.74 [64]

Pig bone 2157 2.26 [65]

Sunflower seed shell 2585 1.41 [66]

Waste paper 526 __ [67]

Wood sawdust 2967 1.35 [68]

Wheat straw 2316 1.50 [69]

Potato starch 2342 1.24 [70]

In the literature, coal is another abundant carbon precursor for AC production. Since Turkish

coal has fewer volatiles (~50 wt%) than biomass materials (60-80 wt%), the AC yield from four

Turkish deposits ranged from 50-60%, which is slightly higher than the AC yield from biomass

(<50%) under similar experimental conditions [7]. Extensive research on chemical activation of

Spanish coal showed that the chemical reaction between carbon and hydroxides for lowest rank

coal (less fixed carbon, more fractions of moisture and volatiles) started at a lower activation

temperature than the highest rank coal [8]. This indicated that the material purity also influences

the carbon reactivity for the oxidation reactions.

As high temperatures (> 1000 ˚C) were applied when producing petcoke, fewer volatiles (< 10

wt%) remain on petcoke after thermal treatment. Therefore, high yields (50-80 %) of petcoke

derived AC were reported in the literature [13, 39]. Except for more fixed carbon on petcoke,

other significant differences to distinguish petcoke from other carbon precursors are its non-

porous surface and the partially ordered structure, respectively: one, petcoke is a non-porous

Page 28: Conversion of Petroleum Coke to Porous Materials

14

material (surface area < 2 m2/g) [15] while coal and biomass are porous materials (surface are ~

12 m2/g for coal [71], 100-250 m2/g for biomass [72] ); two, coal has amorphous amorphous

structure [73], while petcoke has crystalline graphitic carbon layers with a random orientation

within the structure [33].

2.2.3 Pore development approach

Activation generally involves oxidation and/or intercalation of species within the structure [74,

75]. In the former process, solid carbon is transformed into CO and CO2, creating pores, while in

the latter process, a species, generally a metal, acts as a spacer and increases the pore volume by

widening the distance between the carbon layers [76]. In either method, the product AC is

generally washed to remove excess reactants and to ensure that the pores are accessible.

The advantage of intercalation during activation is the increase in porosity without carbon

consumption, resulting in higher yields and lower emissions than oxidation. Metallic potassium

(K) and metallic sodium (Na) were the most accepted intercalation species in the literature. The

metallic potassium was widely reported as a more active species for intercalation when

compared to metallic sodium. The active intercalation could be explained by the reactions

between distilled metal and carbon precursors of various structural orders [75]. Distilled K (K

vapor) was capable of reacting with highly ordered carbon precursors, as indicated by the

increase of interplanar distance from 0.35 to 0.4 nm (calculated from XRD). For distilled Na,

similar reactions could only happen with a less ordered carbon precursor. The study also found

that the interplanar distance from K vapor reactions was similar to the one produced from KOH

Page 29: Conversion of Petroleum Coke to Porous Materials

15

activation of the same carbon precursor. Therefore, it is clear that the increase in interplanar

distance during KOH activation is related to the formation of metallic potassium.

However, it was difficult to know the exact temperature to form these active species since these

metals are sensitive to temperature and are highly reactive with oxygen when exposed to air.

Since at 25 ˚C and 1 atm the standard reduction potential for K and Na are -2.93 and -2.71,

respectively [77]. It indicated a high tendency for metallic K or Na to be oxidized even at room

temperature. Maintaining the metallic forms of K or Na is even harder for chemical activation at

800 ˚C.

Additionally, the presence of graphitic structure in petcoke has been reported as a benefit for

lithium intercalation, leading to a larger capacity when using petcoke as anode [32, 33].

However, the correlation between petcoke structure and intercalation during petcoke activation

has not been understood as yet.

In order to understand the oxidation reactions involved in the chemical activation process, many

researchers have studied the product of activation by determining the species in the solid and gas

phase after AC samples are cooled down (e.g. [78] [8]). Table 2-3 shows the experiment results

in comparison with thermodynamics data calculated for each proposed reaction. Similar

oxidation reactions (Equation 2-1 and 2-2) during NaOH activation and KOH activation were

proposed by Lillo-Ródenas et al.[8] and Yuan et al. [78], respectively. During KOH activation,

the major product for both reactions (Equation 2-1 and 2-2) is K2CO3. The XRD showed that

Page 30: Conversion of Petroleum Coke to Porous Materials

16

over 80% of potassium species were in the form of K2CO3 after petcoke was activated at 800 ˚C

before washing with HCl and water [78].

A group of researchers studied the chemical bonding of carbon (C 1s), oxygen (O 1s) and

potassium (K 2p3/2) using the XPS on the surface of lignocellulose AC (3-4 nm) before the

washing process [79]. The deconvolution of O 1s implied the presence of K-O bond on non-

washed AC. However, if the K species was washed away after activation, the O/C ratio also

reduced to a very low value, which indicated that all the oxygen on the surface is attached to

potassium rather than carbon. K/O ratio on the AC surface was lower than the theoretical ratio as

when forming pure K2CO3 or a mixture of K2CO3 and K2O, other K species, such as metallic K

or KO2, may also exist on the surface.

The possible oxidation reactions during the chemical activation with metal hydroxides were

reported as following [5, 78, 80]:

6𝑀𝑂𝐻 + 2𝐶 → 2𝑀 + 3𝐻2 + 2𝑀2𝐶𝑂3 Equation 2-1

2𝑀𝑂𝐻 + 𝐶𝑂2 → 𝑀2𝐶𝑂3 + 𝐻2𝑂 Equation 2-2

𝑀𝑂𝐻 + 𝐶 → 𝑀 + 𝐶𝑂 +1

2𝐻2 Equation 2-3

where M refers to either potassium (K) or sodium (Na).

Page 31: Conversion of Petroleum Coke to Porous Materials

17

Table 2-3. Standard Gibbs Energy calculations (∆G0) (for Equations 2-1 and 2-2) during KOH

and NaOH activation.

Temperature

(˚C)

KOH activation [78] Temperature

(˚C)

NaOH activation [8]

∆G0 (kJ/mol)

Equation 2-1

∆G0 (kJ/mol)

Equation 2-2

∆G0 (kJ/mol)

Equation 2-1

∆G0 (kJ/mol)

Equation 2-2

25 182.4 -122.8

530 27.3 -101.2

600 -2.6 -103.6 630 7.14 -93.3

700 -17.5 -93.7 730 -12.3 -85.4

800 -36.9 -83.8 830 -31.3 -77.3

900 -66.6 -74.1

Lu et al. further investigated the role of K2CO3 during KOH activation of petcoke [81]. The

formation of K2CO3 in the initial pores through the reactions proved to improve the porosity of

AC. Further reactions on converting K2CO3 to K2O or K (see Equation 2-4 to Equation 2-7) were

likely because they were thermodynamically favored at temperatures above 430 ˚C. FTIR spectra

also indicated the formation of -CH-, -CH2- as ‘active carbons’ at 500 ˚C, the active carbon

species disappeared as the temperature increased to 730 ˚C. Meanwhile, the XRD pattern of AC

before the washing process clearly showed the formation of metallic K and K2O at 500 ˚C and

730 ˚C, respectively. Other research groups also found the decomposition of K2CO3 (Equation 2-

4 and 2-5) and the reduction of potassium compounds (Equation 5-6 and 2-7) at temperatures

above 700 ˚C. [5, 75, 82]. However, all the sample analysis was performed after cooled down to

room temperature and exposed to air. Thus, the conclusions should be considered cautiously.

𝐾2𝐶𝑂3 + −𝐶𝐻2− → 𝐾2𝑂 + 2𝐶𝑂 + 𝐻2 Equation 2-4

𝐾2𝐶𝑂3 + 2 − 𝐶𝐻− → 𝐾 + 3𝐶𝑂 + 𝐻2 Equation 2-5

𝐾2𝑂 + −𝐶𝐻2− → 2𝐾 + 𝐶𝑂 + 𝐻2 Equation 2-6

2𝐾2𝑂 + 2 − 𝐶𝐻− → 4𝐾 + 2𝐶𝑂 + 𝐻2 Equation 2-7

Page 32: Conversion of Petroleum Coke to Porous Materials

18

2.2.4 Diffusion into porous media

As activation produced porous carbon materials, a large number of chemical species are

produced and then co-existed within the porous structure during the activation. These chemical

species (listed in Equation 2-1 to Equation 2-7), either in the gas, solid or liquid phases,

spontaneously transport into the pores of carbonaceous materials. The movement is called

diffusion, which refers to a net flow of molecules from places of the high concentrations to

places of the low concentrations, resulting in an equilibrium system [83]. Diffusion is normally

driven by the concertation gradient, chemical potential, temperature and pressure [84].

The pore diffusion is governed by the diameter of pores (shown in Table 2-4). The molecular

diffusion is the ordinary diffusion happened when two components are mixed with a

concentration gradient [85]. The pore diameter (> 10 nm) is larger than the mean free path of

molecules (the average distance for a molecule to travel from one place to another). Knudsen

diffusion, sometimes refer to surface diffusion, occurs when the mean free path of molecules is

larger than the pore diameter (2-100 nm) [86]. The molecules collide with the pore wall instead

of with other molecules. As the interaction is between molecules and pore walls, both surface

force and capillary force made contribution to the diffusion process [87]. An example of

Knudsen diffusion is that gas molecules diffuse into capillaries or through porous solids [85].

There is a transition region between molecular diffusion and Knudsen diffusion, which occurs

when the mean free path is similar to the pore diameter. Configurational diffusion is the

dominant transport mechanism when pore sizes are smaller than 1.5 nm [86]. The overlapping

surface force from the opposite pore walls is the driving force in this stage.

Page 33: Conversion of Petroleum Coke to Porous Materials

19

Table 2-4. Diffusion mechanism in porous media. [85-87]

Diffusion mechanism Pore size (d) Driving force

Molecular diffusion d > 10 nm Concentration gradient

Knudsen diffusion (or

Surface diffusion)

2 nm < d < 100 nm Surface force and Capillary

force

Configurational diffusion d < 1.5 nm Overlapping surface force

Vattipalli et al. studied the mass transport on a hierarchical zeolite (a porous material with both

micropores and mesoporous). A longer diffusion path was observed in comparison to the

physical length of an adsorbent. The result showed that combined surface and configurational

diffusions created longer path for the adsorbent to travel in and out between the external surface

and the internal micropores (simple scheme illustrated in Figure 2-4), which reduced the mass

transfer limitation within the interconnected pore structure.

Figure 2-4. Schematic graph to illustrate diffusion in hierarchical zeolite (a porous material with

both micropores and mesoporous).[87]. Adapted with permission from V. Vattipalli, X. Qi, P.J.

Dauenhauer, W. Fan, Long Walks in Hierarchical Porous Materials due to Combined Surface

and Configurational Diffusion, Chemistry of Materials, 28 (2016) 7852-7863. Copyright (2016)

American Chemical Society.

Page 34: Conversion of Petroleum Coke to Porous Materials

20

2.2.5 Activation method

Chemical Activation

Lillo-Ródenas et al. compared the activation of non-porous anthracite with KOH and NaOH at

730 ˚C, and they found KOH produced more pore volume than NaOH with a higher than one

mass ratio of KOH and anthracite [58]. Thus, the nature of chemical agents has been proved to

be an essential factor during chemical activation. KOH was more effective to develop pores for

ordered carbon precursors like anthracite while NaOH produced higher pore volume and surface

area for less ordered lignocellulose materials [88].

In other studies involving the NaOH activation with different feeds, including non-porous

anthracite [89], plum kernels [90], and corncobs [91], higher pore volumes and pore sizes were

obtained as the chemical agent to feed mass ratio increased to above one, but none of these

studies discussed the process of activation in depth.

Co-activation of chemical and steam

While there are many studies on gasification, in which the goal is to consume all of the feed, co-

activation (e.g., simultaneous chemical - H3PO4, KOH, ZnCl2 - activation and physical activation

- CO2, steam, air) has only been studied by a few groups. The synergistic effect of chemical and

physical agents (2-5 h co-activation from 600 – 900 ˚C) enhanced the pore volume and surface

area on various natural materials including coconut shells [92], palm stones [92], Zizania

latifolia leaves [93], and woody biomass [55] but resulted in yields less than 30%. The

interaction between the chemical and physical agents is not well understood, especially as most

studies chose a fixed amount of the agents and varied the other parameters (temperature, time,

Page 35: Conversion of Petroleum Coke to Porous Materials

21

etc.). In addition, the porosity of the feedstock influences pore development [81, 94], so the

results that have been obtained with porous lignocellulosic materials may not be applicable to

other feedstocks.

2.2.6 Activation of petcoke

In 1993, the Kansai Coke and Chemicals Company developed MAXSORB, a high-surface-area

(3100 m2/g) AC derived from petcoke[5] by KOH activation with a chemical (KOH) to petcoke

weight ratio of 5. Even though a pore volume of 1.76 cm3/g was obtained using a ratio of 5, the

higher chemical usage led to low yields (< 30%), higher production costs, and more waste.

Generally a ratio of at least 3 is required to develop significant porosity (pore volume above 0.8

cm3/g) at 800 ˚C on AC from petcoke [22, 38, 52, 81, 95, 96]. Various approaches have been

tried to reduce the amount of chemical agent required. Deng et al. used intercalation with HClO4

to produce an AC from petcoke with a surface area (~3000 m2/g) that was similar to the AC

prepared using chemical activation with a KOH to petcoke weight ratio of 5 [97]. Wu et al. used

the weight ratio of 2 by co-activation with steam at 800 ˚C for only 25 min, and the surface area

of AC reached 2500-3000 m2/g [9]. The porosity created during the catalytic steam gasification

promoted the diffusion of steam. Other gasification catalysts such as alkali metal carbonate (e.g.

K2CO3) have also promoted porosity development [81]. In many other studies, the yields

obtained and the wastes created by the methods used have been ignored. On an industrial scale,

washing away 75% of the reactants (i.e., for a 3:1 mass ratio of chemical agent to carbon feed)

may not be economically feasible.

Page 36: Conversion of Petroleum Coke to Porous Materials

22

Most raw petcoke used for activation is green petcoke without desulfurization treatment. Thus, a

variety of research groups in the past 10 years have been working on converting high sulfur

petcoke to AC, and applied it as a catalyst base or an adsorbent [21, 39, 98-101]. There are only

limited publications on using low sulfur petcoke as carbon precursor for activation [22, 102,

103]. Considering the summarized data in Table 2-5 (AC from high sulfur petcoke) and Table 2-

6 (AC from low sulfur petcoke), it was difficult to compare AC properties prepared from high

sulfur with low sulfur petcoke, since the petcoke composition and activation conditions,

including chemical to petcoke mass ratio (R), activation temperature, and holding time varied

significantly.

However, no matter what activation condition was applied, pore volume of 0.25 cm3/g started to

be developed at 500 ˚C for high sulfur petcoke (6.5 wt%) [21]; similar pore volume (0.28 cm3/g)

could be only achieved at 600 ˚C for low sulfur petcoke (0.48 wt%) [102]. The different

temperatures for initial pore formation indicated that the breakage of the C-S bond in high sulfur

petcoke promoted pore development at low activation temperatures. Only a few papers recorded

corresponding yields of AC prepared from petcoke. Considering similar activation temperature

above 800 ˚C and the same chemical to carbon ratio of three, AC prepared from higher sulfur

petcoke (4.8 wt%, [101]) produced a higher yield (27% higher) than AC derived from low sulfur

petcoke (0.3 wt%, [22]). The higher yields and lower activation temperatures are two key points

to understand the pore development mechanism for high sulfur petcoke.

Page 37: Conversion of Petroleum Coke to Porous Materials

23

Table 2-5 Summary of AC properties prepared from high sulfur petcoke (4-8 wt% sulfur). The

activation temperature from 500 to 850 ˚C.

Sulfur

content

(wt%)

Activation

condition*

Activation temperature (˚C)

500 600 700 750 800

Pore volume (cm3/g) / Surface area (m2/g)

7.7 [98] R=2, Tc=300-

600 ˚C, tc = 1

h, ta =1 h

0.38-0.67

/936-1673**

7.5 [99] R=2, ta =1 h 0.22/538 0.6/1433 0.7/1600

7.7 [39] R=2, Tc = 400

˚C, tc = 1 h, ta

= 1.5 h

0.65/1174

5.3

[100] R=2 0.73/1204 0.83/1409 0.86/1471

4.8

[101] R=3, Tc = 400

˚C, ta = 1 h

0.72/1281

/69.4***

6.5 [21] R=3, Tc = 300

˚C, tc = 1 h

0.254/

692

* R refer to mass ratio of chemical agent to petcoke.

Tc is carbonization temperature.

tc and ta are the holding times at either carbonization or activation temperatures.

If there is no carbonization temperature or holding time shown in the table, no specific

condition was reported in the literature. ** Maximum pore volume and surface area were carbonized at 450 ˚C ***Yield of AC

Table 2-6 Summary of AC properties prepared from low sulfur petcoke (< 1 wt% sulfur). The

activation temperature was from 500 to 850 ˚C.

Sulfur

content

(wt%)

Activation

condition

Activation temperature (˚C)

600 700 800 850

Pore volume (cm3/g) / Surface area (m2/g) / Yield (%)

0.5 [102]** R=2, ta* 0.283/867 0.437/1265 0.477/1070

0.8 [22] R=2, Tc = 500 ˚C, tc

= 1h, ta =2h 0.68/1841/64

0.3 [103] R=3, Tc = 500 ˚C, tc

= 1h, ta = 2h 0.05/864/65 0.11/1063/59 0.31/750/51 0.44/577/42

Page 38: Conversion of Petroleum Coke to Porous Materials

24

* ta varied according to different temperature. Holding time for 600 and 800 ˚C were 1 h, but for

700 ˚C was 2 h ** Yield was not reported

2.3 Mixing impact

2.3.1 Physically mixing and Impregnation

There are a limited number of publications discussing the impact of different mixing methods on

AC properties. Researchers in this area focused on the different impact of physically mixing and

impregnation. Physically mixing of chemical agents and carbon precursors were accomplished in

a mortar at room temperature, and the entire process was in a dry environment without water

being involved [104, 105]. The impregnation was in wet environment of concentrated activating

agent solution [106]. The chemical agents were impregnated into the carbon precursor by stirring

at given temperature (mostly around 60 ˚C) for several hours, and then samples were dried in the

oven overnight. As described above, impregnation took a longer time and greater energy than

physically mixing.

From the literature, the impact of these two mixing methods was highly related to the nature of

hydroxide and the structure of the carbon precursor. Lillo-Rodenas et al was the first research

group who suggested physically mixing as an easy and effective way to develop porosity [58].

Additionally, their results showed that physically mixing NaOH and Spanish anthracite with a

mass ratio of 2 produced double the amount of micropore volume compared to impregnated AC.

Similar to Spanish anthracite, the physical mixture of NaOH and Mongolian anthracite produced

35% higher pore volume than the impregnated mixture [89]. However, KOH impregnation on

Spanish anthracite created similar micropore volume, and a slightly higher surface area (60 m2/g)

Page 39: Conversion of Petroleum Coke to Porous Materials

25

than physically mixed AC [58]. However, the researchers did not provide further explanation for

the reason NaOH and KOH had different effect on pore development when using different

mixing methods.

2.3.2 Ball-mill mixing

The ball-milling technique is a mechanical grinding or blending process. The grinding material

and grinding balls are mixed in a container for a certain rotation speed and a certain period of

time. For the size reduction in a large scale production, ball-milling is an easy operating method

[106-108]. Homogenous mixture is another advantage of ball-milling technique. By adding

liquid (such as ethylene glycol, isopropanol alcohol, water, acetone, etc.) into the container, wet

ball-mill is used for a better dispersibility in liquid phase [108, 109]. The poor dispersion of

grinding materials in dry ball-milling led to the agglomerates with a much rougher surface of the

particles in SEM images when compared to wet ball-milled boron particles [106]. The thermal

problem caused by operating the rotary tumbler is also reduced with extra liquid in contrast to

the dry ball-mill process. The surface area was increased after wet and dry ball-milling biochar

[110], graphite (dry ball-mill) [107, 111], titanium monoxide (wet ball-mill) [109] and composite

nanoparticles like ZnFe2O4-C (dry ball-mill) [108]. New chemical bonds are formed during high-

energy ball-milling [106, 112]. However, no previous study has investigated the impact of ball

milling mixing between chemical agents and carbon precursor before activation for the final AC

property. With consideration of less energy consumption, low-energy dry ball-milling was used

for the first time to obtain a homogenous mixture before activation. In order to get more

dispersed mixture with smaller particles, low-energy wet ball-milling was also applied in the

synthesis of carbon foam and hierarchical porous carbon (HPC) through a salt template process.

Page 40: Conversion of Petroleum Coke to Porous Materials

26

2.4 Hierarchical porous carbon

Many biological materials are built in hierarchical components over large length scales. The

structure is related to diverse functionalities, such as mechanical protection, chemical barriers,

energy transfer and storage. [113]. The functional adaptability from natural structures has

inspired scientists to solve engineering problems by man-made hierarchical materials. The

advanced bio-inspired materials have been widely exposed to research fields including catalysis,

adsorption, separation, drug delivery, energy conversion and storage [114]. Among these

materials, HPC has become increasingly attractive in recent decades. A representation of 3D

hierarchical graphitic carbon is shown in Figure 2-5. The nanoscale pores with interconnected

macroscale pores within the carbon wall is crucial for the synthesis of HPC materials [115].

Because the electrical conductivity, chemical stability and production cost are combined with

various functional hierarchical levels of pore size, spanning from nanoscales to macroscales.

Figure 2-5. The schematic representation of hierarchical graphitic carbon. Reprinted from S.

Dutta, A. Bhaumik, K.C.W. Wu, Energy & Environmental Science, 2014, 7, 3574-3592 [115]

with the permission from The Royal Society of Chemistry.

2.4.1 HPC synthesis by template

Page 41: Conversion of Petroleum Coke to Porous Materials

27

Template synthesis is an easy approach to operating and controlling the pore structure of HPCs.

It is usually classified as hard template or soft template depending on the structure of template.

The hard templates need to be removed at the end because of either chemical or physical

differences between the template and the final product [116]. However, the soft templates are

thermally decomposable polymers, which lowers the risk of environmental pollution caused by

the corresponding step of removing the templates [115, 116]. Moreover, a sustainable method by

which the template could be recycled is more competitive for large scale production.

Dual-templating has been explored to create pores on two length scales – small mesopores and

large mesopores or macropores. The templates can be both hard templates, both soft templates or

combined hard/soft templates. However, two hard templates make the procedure tedious and

time-consuming [115]. In-situ one-step templating has become an advanced technique to prepare

HPCs. Li et al. reused petroleum pitch as a carbon precursor [117]. The detailed experiment

procedure was shown in Figure 2-6. Silica gel and silica nano-aggregates were chosen to be dual-

templates in one-step templating. 3D-HPCs were fabricated with wormhole-like small mesopores

(3-4 nm) and large particle-like mesopores (about 17 nm).

Page 42: Conversion of Petroleum Coke to Porous Materials

28

Figure 2-6. A scheme of the HPC prepared from pitch-based carbon. Reprinted from Carbon,

48, S. Li, Y. Liang, D. Wu, R. Fu, Fabrication of bimodal mesoporous carbons from petroleum

pitch by a one-step nanocasting method, 839-843 [117], Copyright (2010), with permission from

Elsevier.

Three well-used polymers in soft templating are trilock copolymer surfactant template (CEO-

PPO-PEC), colloidal crystal and polyurethane foams. The former is accepted as a template for

ordered mesoporous structure. The latter two are used for macropores. An example of a dual-

template of polyurethane foam and triblock copolymer was produced by Xue et al. [118]. An

evaporation induced self-assembly (EISA) procedure facilitated the organic-organic self-

assembly between mesoporous coating layer and polyurethane foam. The large internal surface

area of polyurethane facilitated evaporation and assembly.

Limited soft templates are available for mesopores. A tunable pore size of HPCs is obtained by

the combination of hard and soft templates. Liang et al. proposed an approach for 3D

interconnected HPCs with three different levels of pore sizes. Macropores were developed from

PMMA colloidal crystal. After removing the templates, the wormhole-like mesoproes (2.7 nm)

and large spherical mesopores (10 nm) were created from SiO2 nanoparticles.

Ice-segregation-induced self-assembly (ISISA in Figure 2-7) is an environmentally friendly

method because ice rather than polymers is selected as hard template [115, 119]. The aqueous

phase is segregated by ice for macroporous channel formation. After the sublimation of ice, the

porous structure develops.

Page 43: Conversion of Petroleum Coke to Porous Materials

29

Figure 2-7. A scheme for ice-segregation-induced self-assembly (ISISA). Ice channels were

formed with particle segregation during the freezing procedure. Reprinted with permission from

M.C. Gutiérrez, M.L. Ferrer, F. del Monte, Ice-Templated Materials: Sophisticated Structures

Exhibiting Enhanced Functionalities Obtained after Unidirectional Freezing and Ice-Segregation-

Induced Self-Assembly, Chemistry of Materials, 20 (2008) 634-648 [119]. Copyright (2008)

American Chemical Society."

The salt template method is another environmentally friendly approach because the salt (eg.

NaCl) could be easily removed by water [120]. The recycling and reuse of the salt are feasible as

well [121]. This lowers the production cost as well as reducing the risk of environmental hazards

if the salts require pre-treatment before disposal. Furthermore, the size of salts can be controlled;

therefore, the HPC is likely to be synthesized with a expected structure [122]. Lu et al.

established a simple approach to synthesize the HPC monolith using a salt template and a self-

binder [120]. NaCl was the salt template for the macropores structure within the monolith. By

impregnating furfuryl alcohol into SBA-15 (mesoporous silica sieves with uniform hexagonal

Page 44: Conversion of Petroleum Coke to Porous Materials

30

pores ranging from 5 to 15 nm), polymerization reactions occurred to bind the templates (both

NaCl and SBA-15) with furfuryl alcohol together for the formation of the HPC monolith.

Furfuryl alcohol in this study acted as both the carbon source and the binder. The NaCl template

and SBA-15 template were removed by washing with water and HF solution (~10%),

respectively at the end .The pathway to synthesize the HPC monolith is illustrated in Figure 2-8.

Figure 2-8. Synthetic pathway for the HPC monolith by using salt templating and self-binding.

Reprinted from Microporous and Mesoporous Materials, 95, A.-H. Lu, W.-C. Li, W. Schmidt, F.

Schüth Fabrication of hierarchically structured carbon monoliths via self-binding and salt

templating, 187-192. [120], Copyright (2006), with permission from Elsevier.

A recent study by Qiu et al. also used salt templating to precisely control carbon structure of

pitch [123]. NaCl was selected to be a salt template for macropores formation due to its stability

and low cost. Both phenolic resin (hard carbon precursor) and pitch (soft carbon precursor) are

carbon precursors in the study. Three components mixed in a wet ball milling container (Figure

2-9a) to obtain a homogenous mixture before carbonization. After pre-treatment at 750 ˚C and

then carbonization at 1500 ˚C, the HPC with nanoscale and macroscale pores was successfully

produced (See Figure 2-9b). Without the NaCl template, the produced HPC was not able to

Page 45: Conversion of Petroleum Coke to Porous Materials

31

develop macropores as shown in Figure 2-9c. This implied the presence of NaCl has an essential

role in the formation of macrpores during HPC synthesis.

Page 46: Conversion of Petroleum Coke to Porous Materials

32

Figure 2-9. (a) Illustration of the synthesis process of the HPC; (b) SEM image of HPC with salt

template; (c) SEM image of HPC without salt template.. Reprinted from Journal of Energy

Chemistry, 31, D. Qiu, T. Cao, J. Zhang, S.-W. Zhang, D. Zheng, H. Wu, W. Lv, F. Kang, Q.-H.

Yang, Precise carbon structure control by salt template for high performance sodium-ion storage,

Journal of Energy Chemistry, 101-106 [123]. Copyright (2019), with permission from Elsevier.

The binder during monolith preparation should: 1) Increase the amount of carbon to binder ratio

to improve the mechanical strength of the monolith; 2) Ensure nanoscale pores on the particles

b c

a

Page 47: Conversion of Petroleum Coke to Porous Materials

33

are not blocked. Though binders work to form a monolith, most of binders reported in the

literature have an adverse impact on AC property, resulting in the blockage of micropores during

the carbonization process [124]. Pitches have been proved to be a suitable carbon precursor as

well as a binder due to their thermoplastic properties when they are heated [125, 126].

2.4.2 HPC synthesis by activation

Activation is a traditional way for pores to develop without any template. Targets for HPCs,

natural materials with hierarchical structure are considered as economical carbon precursors,

such as luffa sponge [127], animal bone [65, 128], silk cocoon [129], and rice straw [130]. The

pore size distribution is restricted by the pore structure of the raw material. Since the silk cocoon

is made up of natural microfibers, Liang et al. transformed silk cocoon into microporous

carbonaceous material via a simple carbonization process [129]. The activation via chemical

agents (KOH, NaOH, H2SO4) enhanced the porosity and shortened the activation time. A luffa

sponge has four-level hierarchical cellular structure. Fan et al. have suggested KOH activated

carbon derived from luffa sponge combines micrometer channels with micro/mesoporosity

[127]. For the convenience of large-scale application, the useful biopolymers are also extracted

from the industries. A lab scale experiment has been completed by Zhao et al. They prepared

HPCs using phenolic resin synthesized by replacement phenol derived from bark [131].

Commercial polymer is another carbon precursor. Through one-step chemical activation, well-

structured HPCs were fabricated from commercial polymers including phenol-formaldehyde

resin [132] and poly(vinylidene fluoride) [133, 134].

Page 48: Conversion of Petroleum Coke to Porous Materials

34

As a low-cost operation procedure, the 3D hierarchical porous net was constructed via bottom-up

technique. Macro/mesopores were formed by aggregated carbon spheres. Gong et al. developed

two ranges of pore size (macro-mesopores and micro-mesopores) in respective hydrothermal

carbonization and air activation [135]. The carbon sphere of 60-80 nm aggregated slightly with

the help of a structure-directing agent (commercial kayexalate) via hydrothermal carbonization,

and the interparticle void contributed to the macro-mesopore. The consequent air activation

improved the micropore volume. In addition, loose agglomeration of carbon nanoparticles for

macropore formation was reported by KOH activation process [136] or only carbonization [137].

From the reviewed literature, simple chemical activation or carbonization is an approachable

method to fabricate HPCs.

Most HPCs only focus on the micro and mesopores range applied in the adsorption, catalysis and

electrochemical fields. However, there are few research papers on HPCs integrated micropores

and macropores together.

Page 49: Conversion of Petroleum Coke to Porous Materials

35

Chapter Three: EXPERIMENTAL METHODS

This chapter describes the production of AC from petcoke, and the fabrication of hierarchical

porous carbon by using asphaltenes as a binder in between the prepared AC particles. The

characterization techniques include N2 adsorption, CO2 adsorption, scanning electron

microscopy, and Fourier-transform infrared spectroscopy. In addition, the error analysis of AC

preparation is discussed in this chapter.

3.1 Materials

Delayed petcoke provided by Suncor Energy Inc. was ground with a mortar and pestle manually.

By using nesting 8 inches stainless steel standard sieves of No. 100 and No. 50 (W.S. Tyler, ON,

Canada), the petcoke particles were shaken and sieved for 2 min to obtain the sizes between 150-

300 µm. The composition of petcoke is listed in Table 3-1. NaOH (97%, Sigma-Aldrich Inc.,

USA) and KOH (85%, Alfa Aesar, MA, USA) were used for the chemical activation. NaCl

(99%, EMC Chemicals Inc., Germany), isopropanol (>99.5%, Sigma-Aldrich Inc, USA) and

Athabasca Asphaltenes precipitated in n-pentane (C5) were prepared for HPC synthesis. N2

(99.999%, Praxair Canada Inc, Mississauga, ON, Canada) gas was utilized to provide an inert

atmosphere during activation and carbonization.

Page 50: Conversion of Petroleum Coke to Porous Materials

36

Table 3-1. Proximate, ultimate, and ash analysis of petcoke.

Delayed petcoke [15] Athabasca

asphatlenes[138]

Proximate analysis (wt%)

Moisture 0.3 0.1

Ash (dry basis) 3.7 0.6

Volatile (dry basis) 15 63

Fixed carbon (dry basis) 82 36.3

Ultimate analysis (wt%), dry and ash free

Carbon, C 84 83

Hydrogen, H 3.8 8.3

Nitrogen, N 1.8 1.2

Sulfur, S 6.5 7.5

Oxygen, O (by difference) 3.8 0

Ash Analysis (wt%)

Si 17

Al 14

Ti 2.5

Fe 5.4

Ca 3.4

Mg 1.2

Na 1.4

K 2.9

P 0.1

S 0.8

Balance 51

3.2 Preparation of activated carbon

The mass ratios are commonly reported in the literature. However, the molar ratios are more

instructive when considering the chemical reactions. In this set of experiment, AC was prepared

Page 51: Conversion of Petroleum Coke to Porous Materials

37

from petcoke by chemical activation or chemical steam co-activation. In a mortar and pestle, 2.5

g of petcoke was hand mixed with NaOH or KOH (the chemical agents) with mass ratios of

chemical agent to petcoke from 0.5-3. In the following chapters, the discussion was based on the

molar ratios of NaOH or KOH to carbon calculated according to the purities of petcoke (84%

carbon), KOH (85%) and NaOH (97%). As shown in Table 3-2, the equivalent molar ratios of

NaOH to carbon were from 0.18-1.06, and for KOH to carbon were 0.11-0.64. The samples are

named according to the molar ratio and activation method (eg. KOH stands for KOH activation;

KOH/St indicates KOH steam co-activation) used, then followed by activation temperature and

holding time at that temperature. (eg. 0.43KOH_600_40 refers to a sample activated with a

molar ratio of KOH to carbon of 0.43 at 600 ˚C with 40 min holding time). The default activation

temperature and holding time were 800 ˚C and 30 min, respectively.

Table 3-2. Equivalent molar ratio for mass ratio of NaOH and KOH.

Mass Ratio Molar Ratio

NaOH KOH

0.5 0.18 0.11

1 0.35 0.21

2 0.71 0.43

3 1.06 0.64

3.2.1. Chemical activation

Chemical activation is a traditional approach to obtain highly microporous and high surface area

(> 950 m2/g) AC from petcoke [5, 8, 30, 99]. The mixture of petcoke and chemical agent was

placed in a ceramic boat, which was then placed inside a horizontal furnace and heated under

flowing nitrogen (100 cm3/min) at a rate of 5 ˚C/min from room temperature to the activation

Page 52: Conversion of Petroleum Coke to Porous Materials

38

temperature of 800 ˚C. The mixture was then held at this temperature for 30 min. The produced

AC was cooled down to room temperature, then washed with 250 cm3 of 1 molar HCl solution

and de-ionized (DI) water until the pH of the washing solution was neutral. The sample was

dried in oven at 120 ˚C overnight before further analysis. Thermogravimetric analysis (TGA)

was completed by another PhD student in our group Ross Arnold. 10 mg AC was placed in an

alumina crucible in TGA unit (Q600, TA Instruments, New Castle, DE). Oxygen was flowed

into the system at 100 mL/min to burn the sample, and the changes in sample mass were

measured as the temperature increased from room temperature to 800 ˚C at the rate of 50 ˚C/min.

The result showed that samples after washing contained less than 1 wt% ash.

The yield obtained after activation was calculated according to Equation 3-1 in which mi and mf

are the initial and final masses, respectively. The pore volume per gram of petcoke was obtained

by multiplying the total pore volume by the yield.

𝑌𝑖𝑒𝑙𝑑 (%) =𝑚𝑓

𝑚𝑖× 100 Equation 3-1

Different temperatures from 400-800 ˚C were also studied for KOH activation. KOH and carbon

molar ratio of 0.43 (equivalent to a 2:1 mass ratio of KOH:petcoke) was selected for all the

experiment in this section. A similar procedure was followed as the previous chemical activation

of petcoke, however, the temperature profile was different. The furnace was heated up from

room temperature to certain activation temperatures (400 ˚C, 500 ˚C, 600 ˚C, 700 ˚C and 800

˚C), and there is no holding time (0 min) applied at these activation temperatures.

Page 53: Conversion of Petroleum Coke to Porous Materials

39

In order to understand the impact of total activation time, the activation holding time was

extended from 0 to 40 and 240 min at various activation temperatures.

The activation setup is displayed in Figure 3-1. Petcoke was activated in a horizontal ceramic

tube reactor (L: 70 cm, OD: 4 cm). The nitrogen was introduced into the reactor, and the nitrogen

flow was modulated with a mass flow controller (Brooks Instrument, Hatfield, PA, USA). The

steam was generated before activation started at 800˚C by heating up the temperature above 110

˚C in steam generator. A two-way valve was used to control the activation atmosphere, either

pure N2 or N2 mixed with steam. The actual temperature of the furnace was measured with a

thermocouple in the center of the furnace, and it was compared with the programmed

temperature and then controlled by a temperature controller. Water and KOH solution were used

to remove toxic emissions such as CO and H2S. The fine charcoal particles in the gas flow were

filtered before the exhaust gases were vented to a fume hood.

Figure 3-1. Activation setup for converting petcoke to porous materials. (Revised from [139])

Page 54: Conversion of Petroleum Coke to Porous Materials

40

3.2.2. Co-activation (Chemical and steam activation)

In order to reduce the chemical amount utilized during activation, steam as another source of

oxygen was introduced to the chemical activation system. Steam was injected into the furnace, at

a rate of 5 cm3/h, when the sample temperature reached 800 ˚C. In the presence of steam,

activation times of 15 min, 30 min, 45 min, and 60 min were tested. Samples produced from co-

activation have “St” added to the name. For example, 0.43KOH/St refers to a sample co-

activated with KOH and steam with a molar ratio of KOH to carbon of 0.43. After cooling down

to room temperature, the products were washed and dried in the same way described in 3.2.1.

3.3 Mixing impact

In order to understand the mixing impact on AC properties, petcoke and chemical agent

(NaOH/KOH) were combined in three different ways before activation.

3.3.1 Physically mixing

Physically mixing is a common mixing approach for AC preparation due to its simple process as

well as effective porosity development. The physically mixing of petcoke and chemical agent

was achieved manually using a mortar and a pestle. In this study, chemical agent pellets

(KOH/NaOH) were ground into powder in the mortar for 2 minutes. Then, 2.5 g petcoke (150-

300 µm) was mixed with the chemical powders in the mortar for another minute to obtain a

uniform mixture.

3.3.2 Ball-milling

Page 55: Conversion of Petroleum Coke to Porous Materials

41

Ball-milling was performed with a Rotary Tumble (LORTONE Model 33B, WA, USA) at a

rotational speed of 400 rpm for 14 h. Ball-milling mixed 2.5 g petcoke and chemical agent

(NaOH/KOH) using zirconia grinding balls (with diameter of 5 mm) in a dry plastic milling jar

(200 mL). The mass of grinding ball was ten times of the total mass of petcoke and chemical

agent.

3.3.3 No mixing

In these experiments, chemical pellet was placed below or above 2.5 g petcoke in the ceramic

boat. This method was used to observe the bulk mobility of the chemical agents.

In the mixing methods mentioned above, the mass ratio of chemical agent to petcoke was 2,

which corresponds to molar ratio of 0.71 for KOH or 0.43 for KOH (See Table 3-2). The

chemical activation was performed with either KOH or NaOH at 800 ˚C for 30 min.

3.4 Preparation of carbon foam

Carbon foam was synthesized by a salt template approach. Asphaltenes were attempted to be

binder in this study to form carbon foam structure. The carbon source in this set of experiments

is either carbon black (CB, Monarch 120 from Cabot Cop, Boston, MA) or ball-milled petcoke.

The ball-milling process in 3.3.3 was followed to further reduce the size of petcoke (150-300

µm). In salt template process, wet ball-milling was utilized to mix 1 g carbon precursor (carbon

black or ball-milled petcoke), 1 g binder (asphaltenes), 4 g NaCl and 60 g zirconia grinding balls

in 60 mL isopropanol for 9 h. After removing grinding balls, the mixture was dried in a muffle

furnace at 70 ˚C for 12 h. The powder that resulted was pressed with a hydraulic mounting press

Page 56: Conversion of Petroleum Coke to Porous Materials

42

(Buehler LTD, IL, USA). A hydraulic pressure of 5 MPa was applied to the powder for 20 s,

which led to the formation of a pellet. In order to fix the carbon structure, the pellet was then

carbonized in N2 atmosphere at certain temperature (400 ˚C or 600 ˚C) for 2 h. After the pellet

was cooled down, the salt was removed by washing alternatively with hot water and in an

ultrasonic bath for 1 h consecutively five times. Finally, the sample was dried in the oven at 40

˚C overnight.

3.5 Characterization

3.4.1. N2 adsorption

Surface area, pore volume and pore size distribution were determined by N2 adsorption on the

TriStar II Plus apparatus (Micromeritics Instrument Corp., Norcross, GA, USA) at -196°C.

Before the physisorption experiments, all samples were degassed in a separate unit under

vacuum at 150 °C for 5 h to remove the remaining moisture and volatiles on the carbon surface.

The typical sample mass used was 0.1 g [140]. The dry sample mass was measured immediately

after completing the TriStar analysis. The relative pressure (P/P0) for N2 adsorption was from

0.01-0.995.

The total pore volume was obtained at a relative pressure of 0.97 on the isotherms, and the pore

size distributions were determined with the SAIEUS program (Micromeritics). The best fits to

the isotherms were obtained with the 2-Dimensional Non-Local Density Functional Theory with

heterogeneous surface (2D-NLDFT-HS) model for samples activated with KOH, and the 2D-

NLDFT model for samples activated with NaOH. More details on these methods and their

applications to carbon-based materials are available elsewhere [140]. The micropore (< 2 nm)

Page 57: Conversion of Petroleum Coke to Porous Materials

43

and mesopore volume (2-50 nm) were the cumulative pore volume read from the pore size

distributions.

3.4.2 CO2 adsorption

CO2 adsorption was proposed as a standard adsorbate for microporous material, especially for

pore size smaller than 1 nm, since N2 molecule has diffusion problems into micropores smaller

than 0.45 nm at low temperature (-196 ˚C). CO2 adsorption conducted at higher temperature (0

˚C) and higher saturation pressure (34.5 atm), which resolve the diffusion issues into narrow

micropores. Therefore, the use of CO2 adsorbate was suggested as a complement for N2

adsorption in several publications [141-143].

In this study, CO2 adsorption was performed when N2 adsorption did not complete after 24 h,

which is an indication of the existence of ultra-micropores (pore size < 0.7 nm). The pore volume

and surface area determined through CO2 adsorption were collected by TriStar II Plus apparatus.

The same degassing procedures (in 3.4.1) were followed for all CO2 adsorption experiments. Ice

water was required to keep the adsorption temperature at 0 ˚C for the entire CO2 adsorption

process (around 9 h depending on the sample properties). The ice water was prepared by adding

ice cubes into in the dewar with DI water. An adequate amount of ice was needed, but the liquid

phase must remain in the dewar. The mobile mixture lowered the risk to break the glass sample

holders. CO2 adsorption was suggested to collected data up to P/P0 at 0.3 with NLDFT model

[140]. The saturation pressure at 0 ˚C for CO2 is 26400 torr (equivalent to ~35 atm) [144]. It is

noted that the relative pressure (p/p0) should be carefully chosen when analyzing with CO2

adsorption (See Appendix A).

Page 58: Conversion of Petroleum Coke to Porous Materials

44

3.4.3. Scanning Electron Microscopy (SEM)

For scanning electron microscopy (SEM) analysis, samples were placed on a holder using carbon

tape and then analyzed with an SEM (FEI Quanta 250 field emission SEM, Hillsboro, Oregon,

USA). In order to obtain sample cross-sections, the samples were embedded in an epoxy resin

(Devcon). A glass Pasteur pipette was filled with the mixture and then allowed to solidify

overnight. The pipette was dipped into liquid N2 for 10 s, and then broken to obtain the sample

cross-section after removing the glass pipette. The sample was mounted on carbon tape in a

sample holder for SEM analysis. The Energy Dispersive X-ray (EDX) mapping in SEM analysis

is a surface technique to measure the elemental composition of samples. However, the sulfur

composition measured by the benchtop SEM may not accurate since the samples were not under

vacuum (See Appendix A). An elemental analyzer is suggested to be utilized for sulfur

composition in the future research.

3.4.3 Fourier-transform infrared spectroscopy (FTIR)

The surface groups were determined by Fourier transform infrared spectroscopy (FTIR, Nicolet

iS 50 Spectrometer, Thermo Fisher Scientific, Madison, US). A DRIFTS accessory was selected

for the measurement. The sample was mixed with KBr (0.3-0.5 wt%), and then placed in a

cylindrical holder for the analysis. Spectra were collected between 400 and 4000 cm-1 with a

resolution of 4 cm-1, and 120 scans.

3.4.4 Raman spectroscopy

Page 59: Conversion of Petroleum Coke to Porous Materials

45

Raman spectroscopy measures the rotational and vibrational modes of molecules. Once an ultra

violet-visible beam is excited and interact with the tested molecule, a shift in energy provides

information on chemical structure, crystallinity and molecular interactions. In this study, Raman

analysis was used to gather structural information for petcoke and AC samples. Raman

spectroscopy was conducted by a Research Associate Nael Yasri in Dr. Robert’s group at

University of Calgary, using WITec Raman Microscope (alpha300 RA, WITec, Germany)

3.4.5 X-ray Diffraction (XRD)

XRD was also used to gather information of the crystallographic structure of petcoke and AC

samples. A Multiflex X-ray diffractomer (Figaku, Woodland, Texas, US) in Department of

Geoscience at the University of Calgary was operated by previous PhD student Luis Virla. The

experiment was performed by using Cu/Kα radiation, with 24kV, 20 mA and a scan rate of 0.5

˚/min from 5-90˚ of 2θ.

3.6 Error source

Variability in the data could be from variations in the carbon precursor, sample preparation and

sample characterization. Since the particle size of petcoke has a significant influence on pore

volume, surface area and AC yield [10], the petcoke was ground with a mortar and pestle

manually. By using standard sieves of No. 100 and No. 50 (W.S. Tyler, ON, Canada), the

petcoke particles were shaken and sieved 2 min to obtain the sizes between 150-300 µm. At least

three times of alternative sieving particles between 150 µm and 300 µm were applied to obtain

final petcoke particles for the experiments. In order to get an accurate N2 adsorption result, the

sample was degassed for 5 h at 150 ˚C after analysis was completed. Since the sample mass is

Page 60: Conversion of Petroleum Coke to Porous Materials

46

very sensitive to the N2 adsorption, fingerprints should be avoided by using cotton gloves while

transferring the sample to the balance. At least two times of mass measurements were recorded

until the variation was within + 0.2 mg. Since precision of scale for the analytical balance

(MS304S, Mettler Toledo, QC, Canada) in our lab is + 0.2 mg.

3.7 Repeatability of the experiments

To determine the repeatability of AC preparation, four AC samples were prepared with petcoke

at activation temperature of 800 ˚C for 30 min with a molar ratio of KOH to carbon of 0.43.

Since the AC properties in Table 3-3 (total pore volume, DFT surface area and yield) did not

have any outliers, it is appropriate to use t-test to find out if the difference between the means for

each AC property is significant. Most researchers select 95% confidence interval [145], t-

distribution critical value at α/2 of 0.025 and degree of freedom of 3 was 3.182 [145]. The

calculated confidence interval (based on Eq 3-4) for total pore volume, DFT surface area and

yield are 1.04 + 0.08 cm3/g, 1590 + 80 m2/g, 47 + 1.5%. Thus, there isn’t significance difference

between the sample means within 95% confidence interval.

The average was calculated by Eq. 3-2

�̅� =∑ 𝑥𝑖

𝑛𝑖=1

𝑛 Equation 3-2

The standard deviation was calculated by Eq. 3-3

s = √1

𝑛−1∑ (𝑥𝑖 − �̅�)2𝑛

𝑖=1 Equation 3-3

The confidence interval was calculated by Eq 3-4

�̅� ± 𝑡𝑛−1,𝛼/2𝑠

√𝑛 Equation 3-4

Page 61: Conversion of Petroleum Coke to Porous Materials

47

In the equations above, n is sample size, s is sample standard deviation, �̅� is sample mean, 𝛼 is

the significance level (calculated by 95% confidence interval)

Raw petcoke is not a homogenous material. AC produced from petcoke may not be homogenous

as well. In order to check if AC properties analyzed by N2 adsorption is stable and uniform, two

samples of a highly porous AC (0.43KOH_800_240) were selected and then analyzed in two

different ports of TriStar Plus II in one experiment. Two sample properties were listed in Table

3-3. The results were the same within instrumental error. If compared to the properties of same

AC sample analyzed by N2 adsorption two months earlier, the similar total pore volumes were

obtained but surface area decreased ~330 m2/g. The decrease of surface area may be related to

the adsorption of small molecules in the air, such as H2O and CO2 when AC samples were stored

not under vacuum. Therefore, if AC was not newly-prepared samples, they were suggested to be

degassed at 250 ˚C for at least 3 h under vacuum before the physisorption tests and the future

applications for these materials [10]. The same sample was degassed at 250 ˚C overnight after 5

months preparation, both pore volumes (0.94 cm3/g) and surface areas (1800 m2/g) were the

similar as newly prepared AC (See Table 3-3).

Page 62: Conversion of Petroleum Coke to Porous Materials

48

Table 3-3. The physical properties selected AC.

Date Sample Pore volume

(cm3/g)

DFT surface area

(m2/g)

Yield

(%)

2016.08.03 0.43KOH/St_800_30 1.07 1550 48

2016.08.24 0.43KOH/St_800_30 1.09 1600 46

2017.03.27 0.43KOH/St_800_30 1.04 1650 47

2018.09.11 0.43KOH/St_800_30 0.98 1540 46

Average 1.04 1590 47

Standard derivation 0.048 50 0.957

2019.06.08 0.43KOH_800_240 0.932 1440 71

2019.06.08 0.43KOH_800_240 0.929 1520 71

Average 0.931 1480 71

Standard derivation 0.002 54 0

2019.04.02 0.43KOH_800_240 0.944 1810 71

2019.09.27 0.43KOH_800_240 0.938 1800 71

Page 63: Conversion of Petroleum Coke to Porous Materials

49

Chapter Four: THE IMPACT OF THE AMOUNT OF CHEMICAL AGENT

An investigation of reducing chemical amount without compromising pore volume will be

focused in the next section (4.1). Steam was added as another source of oxygen to a chemical

activation system, which is called chemical steam co-activation. Pore development during

chemical activation, with and without steam, will be discussed (4.2). This chapter has been

adapted with the permission from Jingfeng Wu, Vicente Montes, Luis D. Virla, Josephine M.

Hill (2018) Impact of amount of chemical agent and addition of steam for activation of

petroleum coke with KOH and NaOH, Fuel Processing Technology [13]. Copyright (2018), with

permission from Elsevier (See Appendix E). My role in this paper was designing and performing

the experiments, analyzing data, and writing the manuscript.

Page 64: Conversion of Petroleum Coke to Porous Materials

50

4.1 Reduction of chemical ratios by adding steam as another source of oxygen

The total pore volumes and yields as a function of the amount of chemical agent are shown in

Figures. 4-1 and 4-2 for KOH and NaOH, respectively. Chemical activation generally produced

microporosity (>70%, the specific pore size distributions will be discussed later). In the absence

of steam, the pore volumes increased (Figures. 4-1a and 4-2a) while the yields decreased - from

83% to 67% (Figures. 4-1b and 4-2b) - as the amount of chemical agent increased. Specifically,

the pore volume of AC activated with KOH increased from 0.16 cm3/g to 1.1 cm3/g as the molar

ratio increased from 0.11 to 0.64 (Figure 4-1a). There was a linear increase in total pore volume

with the amount of KOH (e.g., doubling the amount of KOH, approximately doubled the total

pore volume) and this linear relationship appears to pass through the origin (Figure 4-1a).

Because the porosity developed was mainly microporous, the trends for surface area are

essentially the same as those for total pore volume. The surface areas of these samples varied

from 380-2000 m2/g (Table 4-1 provides detailed information on the samples). Many researchers

have used the BET model to determine surface area but it is known that the BET model

overestimates the surface area for microporous materials [146]. For comparison to the literature,

however, the surface areas in terms of the BET model varied from 380-2800 m2/g. The

properties (pore volumes and yields) were the same whether the activation time was 0 or 30 min

(Figure 4-3), suggesting that all of the oxygen, hydrogen and intercalating species in the

chemical agents reacted with the carbon before reaching the final activation temperature.

Page 65: Conversion of Petroleum Coke to Porous Materials

51

.

Figure 4-1 Impact of amount of KOH on (a)

total pore volume per gram of AC, (b) yield

and (c) pore volume per gram of petcoke

during the activation of petcoke (800 ˚C, 30

min) with KOH () and KOH/St ()

activation

Figure 4-2 Impact of amount of NaOH on

(a) total pore volume per gram of AC, (b)

yield and (c) pore volume per gram of

petcoke during the activation of petcoke

(800 ˚C, 30 min) with NaOH () and

NaOH/St () activation

Page 66: Conversion of Petroleum Coke to Porous Materials

52

Figure 4-3. Total pore volume of KOH activated carbon at 800˚C for 0 min () and 30 min ()

with KOH to carbon ratios of 0.11, 0.21, 0.43, 0.64 without steam addition. The surface areas

and pore size distributions of the samples activated for 30 min are essentially the same as those

of the corresponding samples activated for 0 min as given in Table 4-1.

It is also noted that similar surface areas were obtained with raw petcoke and petcoke treated at

800 ˚C for 30 min in N2 atmosphere (Table 4-1). The result implied that no thermal

decomposition occurred on petcoke when petcoke was heated up to 800 ˚C under N2 atmosphere.

The SEM images of the petcoke cross-section verified that without chemical agents involved in

the activation, petcoke particles kept similar morphology (See Figure 4-4).

Page 67: Conversion of Petroleum Coke to Porous Materials

53

Figure 4-4. Cross-section of (a) raw petcoke and (b) petcoke after heated up to 800 ˚C in N2

atmosphere for 30 min.

When steam was added for 30 min at 800 ˚C, higher pore volumes were obtained for ratios of

0.21 and higher (Figure 4-1a). In fact with steam, essentially same total pore volumes could be

obtained with significantly less chemical agent. That is, a total pore volume of 1.1 cm3/g was

obtained with a molar ratio of 0.64 (3:1 mass ratio) without steam or 0.43 (2:1 mass ratio) with

steam. Similarly, a total pore volume of 0.67 cm3/g was obtained with a molar ratio of 0.43 (2:1

mass ratio) without steam or 0.21 (1:1 mass ratio) with steam. The yield was much lower with

steam and relatively constant (~45%) regardless of the amount of chemical agent used (Figure 4-

1b). For comparison, Wu et al. reported yields of 25 – 30% for co-activation of petcoke at 800

˚C for 25 min with a KOH to carbon molar ratio of 0.43 (no pore volume was reported) [9]. The

lower yields during co-activation are consistent with catalytic gasification of the petcoke. Note,

the washing step removed all potassium (and sodium) added and so the yield refers to the amount

of carbon (and less than 1 wt% ash) remaining in the product. Normalizing the pore volume by

the initial amount of petcoke, rather than by the amount of AC produced showed that the

Page 68: Conversion of Petroleum Coke to Porous Materials

54

increased pore volume obtained in co-activation was offset by the lower yields (Figure 4-1c).

Other analyses of the samples showed no differences in the surface functional groups (diffuse

reflectance infrared Fourier transform spectroscopy, DRIFTS) or the combustion behaviour

(thermogravimetric analysis, TGA) whether the AC was produced with or without steam.

The results for activation with NaOH are shown in Figure 4-2. Similar to activation with KOH,

the total pore volume achieved increased with increasing amount of chemical agent (Fig. 4-2a).

The total pore volumes, however, were lower than that achieved with KOH, ranging from 0.01

cm3/g to 0.73 cm3/g over the range of ratios studied, and the change in yield was greater – from

91% to 54% (Fig. 4-2b). The relationship between the total pore volume and amount of chemical

agent was linear but only between ratios of 0.35 and 1.06 (i.e., starting at a mass ratio of one),

similar to that reported for the activation of non-porous anthracite [58] and plum kernels [91]. In

contrast to co-activation with KOH, the addition of steam did not improve co-activation with

NaOH (Figure 4-2). Both lower pore volumes and yields were obtained, suggesting that the

particles were consumed rather than activated to develop porosity.

Page 69: Conversion of Petroleum Coke to Porous Materials

55

Table 4-1. Physical properties of activated carbon produced from the activation of petcoke at 800

˚C with KOH or NaOH.

Sample Steam Time

(min)

Total pore

volume(cm3/g)

DFT

SA(m2/g)

BET

SA(m2/g)

Yield

(%)

Petcoke* 0 < 0.01 < 0.1 < 0.1 98

0.11KOH 0 0.16 380 377 83

0.11KOH/St 15 0.15 259 321 50

0.11KOH/St 30 0.15 306 389 46

0.11KOH/St 45 0.1 149 180 7

0.21KOH 0 0.37 936 885 80

0.21KOH/St 15 0.64 1027 1332 51

0.21KOH/St 30 0.57 852 1115 44

0.21KOH/St 45 0.71 991 1296 19

0.21KOH/St 60 0.69 961 1212 7

0.43KOH 0 0.67 1867 1825 77

0.43KOH/St 15 0.99 1636 2221 61

0.43KOH/St 30 1.08 1576 2197 47

0.43KOH/St 45 1.17 1605 2289 38

0.43KOH/St 60 1.29 1739 2392 32

0.43KOH/St 90 1.23 1608 2183 14

0.43KOH/St** 30 0.6 951 1227 24

0.64KOH 0 1.1 1872 2411 67

0.64KOH/St 15 1.37 1998 2942 56

0.64KOH/St 30 1.4 1987 2880 48

0.64KOH/St 60 1.51 1945 2745 34

0.64KOH/St 120 1.53 1816 2464 6.5

0.18NaOH 0 <0.01 - - 91

0.18NaOH/St 30 0.07 33 37 58

0.35NaOH 0 0.27 396 390 74

0.35NaOH/St 30 0.05 28 27 56

0.71NaOH 0 0.6 1034 1064 59

0.71NaOH/St 30 0.45 560 614 52

1.06NaOH 0 0.73 1277 1351 54

1.06NaOH/St 30 0.69 907 995 36

1.06NaOH/St** 30 0.22 166 169 22

*Petcoke was activated in N2 at 800 ˚C for 30 min.

**For these samples, steam was introduced at 120 ˚C and then the sample was heated to 800 ˚C for extra

30 min in humidified N2.

Page 70: Conversion of Petroleum Coke to Porous Materials

56

These results may be explained by 1) the difference in the initial porosity (pore volume and

possibly pore shape) created by KOH and NaOH before steam was introduced; larger porosity

facilitates diffusion and gas/solid interactions, 2) the relative activities for gasification; K is a

more active catalyst than Na [147-149], and 3) the relative diffusivities of the active species.

Analysis of the cross-sections of particles after activation but before washing was attempted but

was difficult. In Figure. 4-5a, it is clear that there was not a clean break through the particles.

Based on the potassium and sulfur mapping shown in Figure. 4-5b, a layer rich in potassium was

evident around the particles. The smooth areas in Figure. 4-5a correspond to the resin while the

rough areas correspond to areas in which the carbon particles have separated from this layer. The

surface areas of all AC samples (with KOH and NaOH) before washing were less than 5 m2/g

compared to >380 m2/g after washing, which is consistent with the pore volume being occupied

by the chemical agent and/or species formed during activation. Sulfur is visible throughout the

particles.

Page 71: Conversion of Petroleum Coke to Porous Materials

57

Figure 4-5. SEM analysis of 0.43KOH/St (a) before washing (b) with elemental mapping. The

distribution of sulfur (orange) and potassium (green) showed in mapping with the grey areas

being the remaining elements (mainly carbon with oxygen).

Once chemical layers were washed away, cracks previously occupied by chemical agents

remained on AC particles. The SEM images in Figure 4-6 showed more cracks on AC activated

with higher KOH (Figure 4-6g and h). There was an increasing number of cracks formed by co-

activation of KOH and steam compared to KOH activation. Especially by co-activation with

highest chemical ratio of 0.64, intense reactions resulted in expanding of the particles and

burning them into several tiny particles, which corresponded to the peak at around 50 µm on

particle size distribution (Figure 4-6p). Except for 0.64KOH/St, other particles kept similar

particle size distribution as petcoke. The difference could be explained by 1) a higher number of

cracks filled with K species before steam addition provided less diffusion limitation for steam; 2)

K species changed to more active species when steam was added. For instance, the intercalated-

like compounds were reported as an essential intermediate to increase surface area during steam

gasification of bitumen coke [150] and AC [151].

Page 72: Conversion of Petroleum Coke to Porous Materials

58

A similar experiment was completed with NaOH activation and NaOH steam co-activation.

Cracks were observed on SEM images of AC with higher amount of NaOH (Figure 4-7 g, h). In

contrast to KOH steam activation, co-activated NaOH and steam did not increase the amount of

cracks when NaOH to petcoke molar ratio of 0.64 was applied. Less intense reactions may be

involved in the co-activation process with NaOH and steam than KOH and steam. Less activity

of Na compared to KOH may also be reflected by similar particle size distributions among AC

prepared from all experiment conditions shown in Figure 4-7.

Page 73: Conversion of Petroleum Coke to Porous Materials

59

Figure 4-6. Scanning electron micrographs and particle size distribution for raw petcoke and AC

activated with KOH with or without steam. KOH:petcoke molar ratios were 0.21/0.43/0.64.

Page 74: Conversion of Petroleum Coke to Porous Materials

60

Figure 4-7. Scanning electron micrographs and particle size distribution for petcoke and AC

activated with NaOH with or without steam. NaOH:petcoke molar ratios were 0.34/0.68/1.02.

To better understand the pore structure developed at different conditions, the isotherms and

corresponding pore size distributions were examined. The shapes of the isotherms depended on

the specific chemical agent used but not significantly on the amount of that agent. Representative

isotherms and pore size distributions are shown in Figures 4-8 and 4-9 for samples 0.43KOH(/St)

and 1.06NaOH(/St), respectively. With chemical activation, these samples had similar total pore

Page 75: Conversion of Petroleum Coke to Porous Materials

61

volumes - 0.67 cm3/g (0.43KOH) and 0.73 cm3/g (1.06NaOH). According to the IUPAC

classifications [152], the isotherm of sample 0.43KOH (Figure 4-8a) is Type I, typical of

microporous materials. The pores filled with N2 at a low relative pressure (~0.01) consistent with

a narrow micropore range (0.5-1.5 nm) as shown in Figure 4-8b. The addition of steam increased

the total pore volume by over 60% (0.67-1.08 cm3/g) (Figure 4-8a) through the creation of larger

pores (1.5-3 nm, Figure 4-8b). Steam would be able to diffuse into the micropores created by

potassium and enlarge the pores. Steam by itself did not activate petcoke under these conditions

(800 ˚C, 30 min)[96].

Although activation with NaOH produced less pore volume (Figure 4-9a), the pores produced

were larger than with KOH – 24% of the pore volume was related to the pores larger than 4 nm

(Figure 4-9b). Hysteresis loops are clearly visible in the isotherms, which resemble Type IV

isotherms typical of micro-mesoporous materials. The pore size distributions were similar for the

samples activated with and without steam, and contained peaks up to ~30 nm from 5 to 30 nm

(see inset in Figure 4-9b). Similar trend was followed with other NaOH to petcoke molar ratios

(See Appendix B).

Page 76: Conversion of Petroleum Coke to Porous Materials

62

Figure 4-8. (a) N2 isotherms and (b) pore size distributions from 2D-NLDFT-HS for petcoke

activated with KOH with/without steam at 800 ˚C for 30 min. ( 0.43KOH, 0.43KOH/St).

Figure 4-9. (a) N2 isotherms and (b) pore size distributions from 2D-NLDFT-HS for petcoke

activated with NaOH and/or steam at 800 ˚C for 30 min. ( 1.06NaOH, 1.06NaOH/St).

4.2 Impact of steam exposure time

As shown in Figure 4-1, the addition of steam increased the pore volume obtained with KOH,

but decreased the yield. Surprisingly, the yield was constant (~45%) regardless of the amount of

KOH. Further experiments were done varying the activation time (0-120 min) in the presence of

steam and the results are shown in Figure 4-10 (total pore volumes and yields) and Figure 4-11

Page 77: Conversion of Petroleum Coke to Porous Materials

63

(pore size distributions) below. The yield decreased with increasing steam exposure time,

essentially linearly, but with different slopes depending on the KOH:carbon ratio (slopes of -

0.039%/min, -0.030%/min, -0.017%/min, -0.012%/min for KOH:carbon ratios of 0.11, 0.21,

0.43 and 0.64, respectively). Thus, the amount of carbon consumed by the addition of steam

decreased with increasing amount of chemical agent, consistent with a thicker layer forming over

the petcoke particles that reduced the interaction of steam with the surface. Based on the widened

pores (Figures. 4-8 and 4-11 below), however, the presence of steam did result in a modification

of the pores.

Further experiments with varying amounts of sample (between 4.5 g and 7.5 g total) but the same

steam flow rate and same chemical to carbon mass ratio produced AC with the same properties

(yield, pore volume, and pore size). The SEM analysis (not shown) showed that the particles

were relatively uniform for specific activation conditions. These results indicate that the amount

of steam was not the limiting factor for most conditions used (with the lowest chemical amounts

where the most carbon was consumed, the amount of steam may have limited the reaction, see

Figure 4-12) and that the reaction atmosphere reached particles throughout the bed. Once the

steam reached the particles, however, there was likely diffusion resistance through the chemical

layer to the carbon. The least amount of porosity was presented on petcoke activated with the

lowest chemical ratio (0.11KOH). The solid carbon restricted the steam diffusion, therefore,

higher amount of steam was required to consume one mole of carbon when introducing steam at

800 ˚C (Figure 4-12).

Page 78: Conversion of Petroleum Coke to Porous Materials

64

Figure 4-10. Impact of steam activation time on total pore volume per gram of AC() and yield

(×) at 800 ˚C with different amounts of KOH. a. 0.11KOH/St, b. 0.21KOH/St, c. 0.43KOH/St, d.

0.64KOH/St.

The pore volume decreased slightly with steam exposure time for the KOH:carbon ratio of 0.11

(Figure 4-10a). For the ratios of 0.21 and 0.64, there was a significant increase within the first 15

min of steam exposure and then the pore volumes remained relatively constant (Figure 4-10b and

d). The results were different for the ratio of 0.43 with the pore volume continuing to increase up

to 60 min of steam exposure (Figure 4-10c). When the total pore volumes are normalized by the

initial amount of petcoke (rather than the amount of AC produced), the maximum total pore

volumes occurred within 15 min of steam exposure for the chemical ratios of 0.43 or 0.64 (see

Figure 4-13).

Page 79: Conversion of Petroleum Coke to Porous Materials

65

Although the total pore volumes plateaued, the pore sizes continued to increase with steam

exposure (Figure 4-11). Without steam (0 min exposure), there was one peak at ~0.7 nm for the

ratios of 0.11 and 0.21 (Figure 4-11a and b). For a ratio of 0.43 (Figure 4-11c), a shoulder has

developed just above a pore width of 1 nm, and for the highest ratio (Figure 4-11d), there is a

second peak at ~1.7 nm. When steam is added, the previous peak at 0.7 nm has shifted to ~1 nm

and there is a second peak at 1.7 nm in the pore size distributions for all samples (Figure 4-11a-

d). The highest intensities of the peaks at 1 nm were reached after 15 min of steam exposure,

after which the relative number of wider pores (i.e., pores > 2 nm) increased with a

corresponding decrease in the number of smaller pores. A quantitative analysis of the change in

pore sizes below 1.4 nm and those between 1.4-5 nm is given in Figure 4-14.

Page 80: Conversion of Petroleum Coke to Porous Materials

66

Figure 4-11. The influence of steam activation time on pore size distributions of petcoke

activated at 800 ˚C with different KOH to petcoke ratios. a. 0.11KOH/St, b. 0.21KOH/St, c.

0.43KOH/St, and d. 0.64KOH/St.

Page 81: Conversion of Petroleum Coke to Porous Materials

67

Figure 4-12. Impact of steam activation time on molar ratio H2O:Carbon consumed when 2.5 g

petcoke activated in the presence of steam at 800 ˚C with different amounts of KOH.

0.11KOH/St (), 0.21KOH/St (), 0.43KOH/St (), 0.64KOH/St ().

Figure 4-13. Impact of steam activation time pore volume created by 2.5 g petcoke activated in

the presence of steam at 800 ˚C with different amounts of KOH. 0.11KOH/St (), 0.21KOH/St

(), 0.43KOH/St (), 0.64KOH/St ().

Page 82: Conversion of Petroleum Coke to Porous Materials

68

Figure 4-14. Impact of steam activation time at 800 ˚C on the change of pore volume from pores

smaller than 1.4 nm() versus pores between 1.4-5 nm(). a. 0.11KOH/St, b. 0.21KOH/St, c.

0.43KOH/St, d. 0.64KOH/St.

Considering the mesopore instead of total pore volume, the mesopore volume in Figure 4-15 of

AC prepared with same chemical ratios increased linearly with steam time. The higher chemical

ratio was applied, the higher increment of mesopore volume was observed.

Page 83: Conversion of Petroleum Coke to Porous Materials

69

Figure 4-15. Impact of steam time on mesopore volume change during KOH/St activation at 800

˚C with various chemical ratios ( 0.64KOH/St, 0.43KOH/St, 0.21 KOH/St, 0.11

KOH/St).

4.3 Discussion of pore development

There are multiple reactions and phase changes occurring during the activation of carbon with a

chemical agent and steam. During heating, the chemical agent melts - the melting points of KOH

and NaOH are 360 ˚C and 318 ˚C, respectively - and/or decomposes, which releases the oxygen

to react with the carbon and/or the metals to intercalate in the structure, both of which could

create the porosity evident at the activation temperature (800 ˚C, Figures. 4-1, 4-2 and Table 4-

1). The carbon-sulfur bonds in the petcoke break first [27, 30], releasing the sulfur into the

molten phase around the particles (Figure. 4-5). Previous studies on the activation of low grade

petcoke without steam (N2 atmosphere) with chemical agent to carbon molar ratios between 0.43

and 0.64 examined the cooled samples by XPS, XRD and XANES. The species present after

Page 84: Conversion of Petroleum Coke to Porous Materials

70

activation with KOH included K2S, K2SO4, and K2O [27, 78]. Analysis of the tubing at the

exhaust after activation at 800 ˚C revealed potassium deposition. The species present after

activation with NaOH were Na2CO3 as detected by XRD and Na2S was assumed to be formed as

sulfur had been removed [78]. None of the analysis was done at the activation conditions. The

melting points of K2CO3 and Na2CO3 are 891 ˚C and 851 ˚C, respectively, but mixtures of

carbonates and hydroxides have melting points between those of the pure substances [153].

A previous study in our group [154] on gasification examined ash free coal with K2CO3

(potassium to carbon molar ratios < 0.14) using in-situ XRD at room temperature and 700 ˚C.

The peaks associated with K2CO3 disappeared at 700 ˚C and reappeared when the sample was

cooled. XRD only detects crystalline phases and so this analysis is also not direct evidence that

carbonate species are not present at activation conditions.

Metal hydroxides can react with carbon as shown in Equations 4-1 and 4-2 [5, 78, 80]:

6𝑀𝑂𝐻 + 2𝐶 → 2𝑀 + 3𝐻2 + 2𝑀2𝐶𝑂3 Equation 4-1

𝑀𝑂𝐻 + 𝐶 → 𝑀 + 𝐶𝑂 +1

2𝐻2 Equation 4-2

where M refers to either potassium (K) or sodium (Na). Based on these reactions, the theoretical

yields were calculated, and as shown in Figure 4-16, the actual yields were between the limits

provided by the two reactions. Carbonates react with carbon by first reducing to the metal [155].

This metal may react through a redox cycle to consume carbon and produce CO or intercalate

into the carbon matrix. Thus, the porosity developed during heating may be from Reaction (4-1),

Reaction (4-2), gasification, and/or intercalation. The literature is in agreement that potassium is

Page 85: Conversion of Petroleum Coke to Porous Materials

71

a more active gasification catalyst [147-149] as well as better able to intercalate than sodium [75,

88].

Figure 4-16. Yields obtained for the activation of petcoke at 800 ˚C for 30 min with different

amounts of KOH () or NaOH(). The lines are the theoretical yields based on Equation 4-1

(solid line) and Equation 4-2 (dashed line).

When steam is introduced, further porosity is developed by steam gasification, conversion of the

chemical agent to a more active form either by reactions with steam (e.g., H2O + K2CO3 ⇌ K2O

+3CO + H2 [155]) and/or a change in the environment around the particles shifting the

equilibrium composition. We were unable to find any data on the diffusion of steam through

liquid hydroxide. If the chemical layer around the particles is very thin or incomplete (i.e., for the

lowest chemical ratios), the steam could more easily gasify the particles (Figures 4-10a and b).

The higher the ratio of chemical agent to carbon, the thicker the layer and the lower the

gasification rate of the particle that reduces the yield with little to no increases in the pore

volume. With the exception of the 0.43 molar ratio (Figure 4-10c), the pore volume did not

Page 86: Conversion of Petroleum Coke to Porous Materials

72

increase significantly after 15 min of steam exposure. At this point, it is not clear why there were

differences with a molar ratio of 0.43.

Figure 4-17. The relationship between steam time and the pore volume change in the presence of

steam at 800 ˚C with different amounts of KOH. The change of total pore volume equals to (pore

volume at steam time X) – (pore volume at steam time 0). 0.11KOH/St (), 0.21KOH/St (),

0.43KOH/St (), 0.64KOH/St ().

The pore sizes increased with steam exposure after KOH activation (Figures 4-8 and 4-11) with a

corresponding decrease in yield suggesting that carbon from the pore walls was consumed.

Introducing steam at a lower temperature of 120 ˚C (Table 4-1) resulted in both a lower yield

(24%) and a lower total pore volume (0.6 cm3/g) than introducing steam at 800 ˚C (48% yield

and 1.1 cm3/g total pore volume), suggesting that porosity and/or the appropriate species/phase

of chemical was not established before steam introduction. A similar trend was observed when

introducing steam at 120 ˚C with NaOH (Table 4-1).

Page 87: Conversion of Petroleum Coke to Porous Materials

73

After NaOH activation, steam exposure decreased the pore volume (Figure 4-2) without

changing the pore size distribution (Figure 4-9) suggesting that the particles were gasified

(consistent with particle size distributions seen in SEM [96]). As sodium is a less effective

activation agent, the porosity likely extended less far into the particles, and exposure to steam

would remove relatively more of the developed pores than when potassium was used. At this

point, however, we do not have direct evidence for the actual extent of porosity development

within the particles. In a previous study [10], petcoke particles less than 45 µm developed

significantly more microporosity than particles of 300-600 µm size activated with CO2 at 900 ˚C,

suggesting the porosity was developed following a shrinking core model. Further investigations

are underway to better understand the pore development within the particles including

interconnectivity of pores and distribution of pores throughout the particles.

In summary, there is a not simple relationship between the amount of chemical agent used and

the properties of the resulting AC. As steam is not effective at activating non-porous materials

[96], activation with a chemical agent is required. If introduced together, significant gasification

of the carbon occurs. Thus, introducing the chemical agent first is a better approach. Before

steam addition, the amount of porosity developed is limited by the amount of oxygen in the

chemical agent (Figure 4-16) and the mobility of the chemical agent - potassium is more mobile

than sodium and hence develops more porosity (Figure 4-1) and also pores of different sizes and

shapes (compare Figure 4-8 and 4-9). When steam is introduced after KOH activation, the pores

are widened (Figure 4-8 and 4-11), but the particles are also consumed (Figure 4-1b and 4-10). A

potassium rich layer around the particle (Figure 4-5) slows this consumption. An in situ

Page 88: Conversion of Petroleum Coke to Porous Materials

74

technique is required to determine the thickness at activation conditions, as well as the particle

size distribution in order to calculate the appropriate chemical agent to petcoke ratio.

4.4 Economic estimation of petcoke activation

As summarized in Figure 4-18, similar pore volume (1.1 cm3/g) were achieved with a lower

chemical amount by co-activation (0.43KOH/St) and a higher chemical amount by KOH

activation (0.64KOH). However, co-activated sample has a slightly smaller yield (47%) than the

chemical activated sample (67%). With the consideration of lower yield by co-activation, the

increased pore volume of co-activated sample was offset by the lower yield. Regarding to

porosity, there is no need to consume additional energy by injection of steam for 30 min (5 cm3

steam). However, the extra cost to generate steam (3 g steam) should be compared with the

saving from less chemical agents (2.5 g KOH). Steam cost is difficult to estimate because there

are multiple ways to generate steam. All parameters in the entire process (e.g. condensate, power

generator, cooling system) need to be considered carefully [156]. The net low-pressure steam

cost generated from conventional methods range from 1 USD/Klb to 3 USD/Klb [156]. So the

cost of extra 30 min steam is from 1.36 USD to 4.08 USD. 1 kg KOH (>84%) is 45.8 USD,

therefore, using 2.5 g less KOH saves 0.11 USD. The cost of steam is over 10 to 40 times of the

savings from reduction of the chemical amounts. Thus, co-activation of steam and KOH require

higher expense in the entire process.

However, comparing between the same amounts of chemical agents in Figure 4-13, the

maximum pore volume was developed by introducing steam for only 15 min with KOH to

petcoke ratios of 0.43 or 0.64. In both case, an additional 0.1 cm3/g pore volume per gram of

Page 89: Conversion of Petroleum Coke to Porous Materials

75

petcoke were formed when introducing 15 min steam. The cost of 15 min steam ranged from

0.68 USD to 2.04 USD. However, the value of 0.1 cm3/g pore volume increasing is heavily

dependant on the sale price of the final AC. As petcoke-derived AC has not been ultilized as a

commercial AC product yet, it is not easy to evaluate the value of 0.1 cm3/g pore volume

increasing, and then compare with the expense of steam generation.

Figure 4-18. Schematic graph for KOH activation and co-activation of petcoke.

Page 90: Conversion of Petroleum Coke to Porous Materials

76

Chapter Five: PORE DEVELOPMENT DURING CHEMICAL ACTIVATION

The development of pores in petcoke is explained in more depth in this chapter. The impact of

mixing methods before activation was investigated at the beginning. The diffusivity of different

chemical agents (KOH/NaOH) was compared with cross-sectional areas of non-washed AC, in

order to understand the pore development at a high activation temperature of 800 ˚C. For lower

energy requirement and increased yield, activation was also performed at lower activation

temperatures with various holding times. A calculated bulk density of petcoke was compared

with the measured value to study the pore structure of raw petcoke. Finally, a pore development

mechanism was proposed.

Page 91: Conversion of Petroleum Coke to Porous Materials

77

5.1 Experimental methods – Cross-sectioning

A similar approach as that described in Chapter 3 was applied for non-washed AC samples

initially but the approach was not successful. The particles did not break and so cross-sections

were not exposed. Therefore, an extra polishing step was required. The polishing used an

automatic polishing machine (LaboSystem, Struers Ltd, Ontario, Canada), operated by a

technician in the Univeristy of Calgary’s Department of Geoscience. Non-washed AC samples

were fixed into a resin as described in Chapter 3, and then samples were polished with the

machine for ~ 5 min.

5.2 Results

5.2.1 Impact of mixing method – ball-milling

As shown in Figure 5-1, the size of the petcoke particles decreased from 50-250 µm to mostly <

10 µm after ball-milling. The peak of particle size distribution of raw petcoke was approximately

150 µm, while the peak decreased to 5 µm after ball-milling. The ball-milled petcoke particles in

Figure 5-1 (b) included not only tiny particles but also a few large particles, compared to raw

petcoke (Figure 5-1 (c)). Thus, ball-mill technique reduced the particle size for the majority of

the petcoke particles.

Page 92: Conversion of Petroleum Coke to Porous Materials

78

Figure 5-1 Particle size change before and after ball-milling petcoke. (a) particle size of ball-

milled petcoke for 14 h in a rotary tumbler and raw petcoke. The corresponding SEM images are

displayed in insets (b) and (c), respectively.

5.2.2 Physically mixing and ball-milling with KOH activation

Different mixing methods were used to determine the impact of the initial dispersion of KOH

with petcoke (Figure 5-2). The least dispersed sample contained KOH pellets with petcoke (150-

300 µm), followed by physically mixed KOH and petcoke (150-300 µm), and then ball-milled

KOH and petcoke (0-150 µm) which was ball-milled together for 14 h with zirconia balls (5

mm).

Particle Size (µm)

0 50 100 150 200 250 300

Fre

qu

en

cy (

%)

0

10

20

30

40

50

60

P-Petcoke B-Petcoke

c b a

Raw

petcoke

Page 93: Conversion of Petroleum Coke to Porous Materials

79

The results in Figure 5-2 showed that directly using KOH pellet and petcoke produced similar

pore volume (~0.88 cm2/g) and yield (72%) as physically mixed KOH powder with petcoke, but

less time and energy consumed when placing the pellets directly on the top/bottom of petcoke.

Figure 5-2. The impact of mixing method for KOH and petcoke on pore volume (micropore

volume indicated by black bar, mesopore volume by grey bar) and yield(solid red circle) of AC.

The total height of the bar in the figure represents total pore volume. AC was prepared from

KOH to petcoke molar ratio of 0.43at 800 ˚C for 30 min. The samples from left to right were AC

from KOH pellet, KOH powder by physically mixing (PM) and ball-milling (BM). The digital

photos show the corresponding mixture of KOH with petcoke before activation.

Although AC produced by ball-milling had the highest pore volume (1.3 cm3/g), the yield was

only 53%. The low yield was mainly from the loss of fine particles by explosion during

activation (similar photo as Figure 5-3c). As KOH melted at 360 ˚C, the carbon oxygen reaction

Pellet PM BM

Po

re V

olu

me (

cm

3/g

)

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

Yie

ld (%

)

0

20

40

60

80

100

Page 94: Conversion of Petroleum Coke to Porous Materials

80

transferred from solid-solid reaction to liquid-solid. As petcoke has a smaller density (~1 g/cm3)

than KOH (2.12 g/cm3), petcoke after ball milling became fine particles which floated on the top

layer of melting KOH. The oxidation reaction released gases, such as H2 and CO, resulting in

sample explosion. Therefore, ball-milling is not required to mix petcoke with KOH. Aside from

being time consuming, the separation of the sample for the zirconia balls is difficult, creating a

higher possibility of sample loss.

In order to quantify the KOH amount in the ball-milled mixture (5 g KOH and 2.5 g petcoke),

the sample was titrated with HCl (0.1g of mixture was put in 20 mL of DI water). The KOH

concentrations of the solutions from these samples were 0.028 mol/L, 0.036 mol/L and 0.039

mol/L. Each sample was titrated at least twice to obtain reliable data (the replicates were within

0.1 mL difference at the end point). The variation indicated ball-balling mixing could not result

in a homogenous mixture of KOH and petcoke. These KOH concentrations by titration were

compared with theoretical KOH concentrations when KOH and petcoke were physically mixed

at mass ratios of 1 and 2. The theoretical values were calculated with the same experimental

conditions as the titration solution (0.1 KOH [85%] in 20 mL DI water). The KOH

concentrations by titration were in the range of 0.037-0.05 mol/L (theoretical values in Table 5-

1). Additionally, the corresponding pore volume (Table 5-1) of AC prepared from ball-milled

KOH and petcoke at a mass ratio of 2 was 0.55 cm3/g, which was in the range of pore volumes

by AC from physically mixed KOH and petcoke with mass ratios of 1 and 2. As discussed in

Chapter 4, the pore volumes were increased with higher amount of KOH. Both titration data and

pore volumes indicated the separation process after ball-milling reduced the initial KOH to

petcoke mass ratio from 2 to nearly 1. The powder mixture of KOH and petcoke stuck to the

Page 95: Conversion of Petroleum Coke to Porous Materials

81

zirconia balls after ball-milling because the washing solution to clean zirconia balls was basic

(pH >10). Also there were a great number of KOH pellets remaining with the zirconia balls after

separating through sieving (mesh size of the sieve is No. 35). A better separation method is

required for further investigation of ball-milling mixing method.

Table 5-1 KOH concentration after separating zirconia balls measured by titration.

Samples Ball-milling (KOH:PC = 2) Ball-milling (KOH:PC = 3) Theoretical values

#1 #2 #3 #1 #2 #3 KOH:PC = 1 KOH:PC = 2

KOH amount in

solution (mol/L) 0.028 0.036 0.039 0.042 0.054 0.049 0.037 0.05

AC pore volume

(cm3/g) 0.55 1.33 0.37 0.74

To increase KOH to petcoke ratio after ball-milling, 10 g of mixture (7.5 g KOH and 2.5 g

petcoke, KOH to petcoke mass ratio of 3) was initially put into the milling jar. However only 6 g

of the mixture was collected after removing the zirconia balls. Therefore, the sample loss was

40% after separating out the zirconia balls. The same titration method was used to measure KOH

amount in the ball-milled mixture. The average KOH concentration in 20 mL DI water for three

samples was 0.048 mol/L, which was similar to the theoretical KOH concentration (0.05 mol/L)

when physically mixed KOH to petcoke at a mass ratio of 2. Thus, this set of data was used for a

fair comparison of different mixing methods with a similar KOH amount.

5.2.3 Physically mixing and ball-milling with NaOH activation

In order to discuss the impact of particle size for pore development, petcoke was ground

physically in a mortar, and then sieved for different particle sizes ranging from 20 to 500 µm

Page 96: Conversion of Petroleum Coke to Porous Materials

82

(20-63 µm, 63-150 µm, 150-300 µm, 300-500 µm). Only few particles smaller than 20 µm could

be obtained by physical grinding and sieving. Therefore, particles smaller than 20 µm in this set

of experiment were obtained via ball-milling, but it was noted that the presence of a few large

particles (63-300 µm) after ball-milling (see Figure 5-3b). The sample was not sieved after ball-

milling because of the limited amount of petcoke obtained.

The pore volumes and yields obtained by chemical activation of NaOH and petcoke with a molar

ratio of 0.71 are given in Figure 5-3a. The change of particle sizes of petcoke did not influence

the mesopore volume when particles were smaller than 300 µm. When the particle size exceeded

300 µm, the mesopore volume decreased slightly by 0.03 cm3/g. The change in particle size,

however, had a significant impact on the micropore volume. It reached a maximum value of 0.46

cm3/g at the particle size range of 20-63 µm. The results were identical to the previous findings

by Karimi who activated petcoke with CO2 at 900 ˚C for 2 - 15 h [10]. That is, higher surface

areas (175 - 650 m2/g) were achieved for 20 - 45 µm particles compared to 50 - 125 m2/g for 300

- 600 µm particles.

As seen in the SEM image (Figure 5-3b), ball-milled petcoke had a mixture of fine particles

(particle size < 20 µm) and larger petcoke particles (63 µm < particle size < 300 µm). The pore

volume measured by N2 adsorption was the average pore volume of the mixture, which led to

similar pore volumes for ball-milled AC and physically ground AC (20-63 µm). AC prepared

from ball-milled mixture was shown in Figure 5-3c. It was likely that a portion of fine particles

exploded when heating up. During activation of the ball-milled sample, a certain amount of fine

particles may remain on the wall of the ceramic tube, be deposited down stream, or blow away

Page 97: Conversion of Petroleum Coke to Porous Materials

83

with the N2 flow. So yield for ball-milled AC reduced by ~ 20% compared to physical ground

AC (20-63 µm). The yields for the physically-ground petcoke samples were similar (~ 40%)

regardless of particle size range.

Figure 5-3. The impact of petcoke particle size (0-500 µm) on (a) pore volume(grey bars are

micropores, black bars are mesopores) and yield(solid red circle). (b) the corresponding SEM

images of five different petcoke particle sizes ranging in size from 0 to 500 µm. (c) digital photo

of ball-milled sample after taking out of furnace.

20-63 µm Ball-milled 63-150 µm 150-300 µm 300-500 µm

b

Ball-milled (majority < 20 µm)

mic

rop

ore

s

me

so

po

res

Page 98: Conversion of Petroleum Coke to Porous Materials

84

Similar experiments with NaOH and petcoke were also done with various mixing methods. The

results in Figure 5-4 showed that despite better mixing, the pore volumes (0.55 cm2/g) and yields

(56%) were similar for all three mixtures. These results suggest that diffusion of the Na to the

petcoke was not the limiting step in the activation. Compared to results of KOH mixing in Figure

5-2, pore volumes of AC activated by NaOH had larger variations (larger error bars). The

difference may be related to different phase behavior of Na and K species during activation. As

shown in Table 5-2, KOH and NaOH melt at 360 ˚C and 400 ˚C, respectively. They are liquid at

activation temperature of 800˚C. As the carbonates (Na2CO3/K2CO3) were produced, the

activation system become complex which includes multiple chemical species listed in Table 5-2.

According to the phase diagrams for the hydroxide and the carbonate (Na2CO3/NaOH and

K2CO3/KOH) [153], the petcoke is in a liquid mixture for both NaOH and KOH activation if

there is hydroxide and carbonate in the activation system. Other products in NaOH activation,

such as Na2S and Na2O, are solid substances at 800 ˚C. While for KOH activation, K2O and K2S

are likely to be liquid phase. Therefore, more solid species (Na2S, Na2O) were possibly presented

in NaOH activation system, which limited the diffusivity of chemical species during activation.

However, it is noted that K2O and Na2O were not stable during activation, and they are

intermediate for the formation of metallic K or metallic Na reported in the literature [5, 157].

Page 99: Conversion of Petroleum Coke to Porous Materials

85

Table 5-2. Physical properties of the pure K and Na species in activation process [158, 159].

Chemical

species

Melting

point (˚C)

Boiling

point (˚C)

Decomposition

temperature

(˚C)

KOH 360 1594 High

temperature

K2S 840 912

K2O 740 300

K2CO3 891 1200

K 63.5 759

NaOH 320 1388 High

temperature

Na2S 1176 920

Na2O 1132 700

Na2CO3 851 >1700

Na 98 883

The surface areas of raw petcoke measured by N2 adsorption were similar before (< 2 m2/g) and

after (< 4 m2/g) ball-milling. A different trend was observed for porous corn-stover as the

feedstock. By controlling the factors during the planetary ball-milling process, the surface area of

biochar from corn-stover feedstock increased from 60 to 194 m2/g [110]. In a different study,

seven commercial AC samples from several types of bituminous coal and biomass were ball-

milled in various conditions, and both surface area and pore volume increased after ball-milling

[160]. The different trend between petcoke and other materials mentioned above are their initial

porosity before ball-milling. As petcoke is a non-porous material while commercial AC in the

aforementioned study had at least 0.35 cm3/g pore volume and 500 m2/g surface area.

Page 100: Conversion of Petroleum Coke to Porous Materials

86

Figure 5-4. The impact of mixing method for NaOH and petcoke on pore volume (micropore

volume indicated by black bar, mesopore volume by grey bar) and yield(solid red circle) of AC

prepared from NaOH to petcoke molar ratio of 0.71at 800 ˚C for 30 min. The samples from left

to right are AC from NaOH pellet, NaOH powder by physically mixing (PM) and ball-milling

(BM) with the molar ratio of 0.71 NaOH to petcoke. The digital photos are the mixtures of

NaOH with petcoke before activation.

Ball-milling reduced the particle size and improved the mixing at the same time. In order to

exclude the factor of the change in particle size, a similar particle size range (0-150 µm in Figure

5-5a) was controlled for physically mixed and ball-milled samples. A mixture of NaOH and

petcoke (150-300 µm) was ball-milled together in rotary tumbler for 14 h with a NaOH to

petcoke molar ratio of 0.71. Particle size of petcoke after ball-milling was measured once NaOH

was washed away with DI water. The particle size distribution in Figure 5-5a (blue dash curve)

showed petcoke particles after ball-milling were in a range of 0-150 µm. Therefore, a similar

particle size range (0-150 µm) was controlled for physically mixed petcoke and NaOH sample.

Page 101: Conversion of Petroleum Coke to Porous Materials

87

Thereby, raw petcoke (150-300 µm) was manually ground in a mortar with a pestle for 5 min,

and the ground petcoke was then sieved to particles smaller than 150 µm. Extra NaOH with a

molar ratio of NaOH to petcoke at 0.71 was mixed with the petcoke (0-150 µm) for ~1 min.

Particle size distribution of the petcoke after sieving (black solid curve) in Figure 5-5a confirmed

that a similar particle size range (0-150 µm) was achieved for both ball-milled and physically

mixed samples. However, AC prepared after ball-milling had a lower pore volume (both

micropore and mesopore volume) and yield compared to physically mixed AC in Figure 5-5b.

Page 102: Conversion of Petroleum Coke to Porous Materials

88

Particle size (µm)

0 50 100 150 200 250 300

Fre

qu

en

cy (

%)

0

5

10

15

20

25

30

PM_0-150 µm

BM_0-150 µm

a

AC_B_0.68Na AC_P_0.68Na

Po

re v

olu

me

(c

m3

/g)

0.0

0.1

0.2

0.3

0.4

0.5

0.6

Yie

ld (%

)

0

20

40

60

80

100b

Figure 5-5. Mixing impact on porosity and yield with the same particle size range (0-150 µm).

(a) particle size distribution of petcoke by physically mixed and ball-milled mixture with NaOH

at a molar ratio (NaOH:petcoke) of 0.71 before activation; (b) micropore volume (dark blue bar

for ball-milled AC, black bar for physically mixed AC), mesopore volume (light blue bar for

ball-milled AC, grey bar for physically mixed AC) and yield (●) of corresponding AC activated

at 800 ˚C for 30 min.

BM_0.71NaOH PM_0.71NaOH

Page 103: Conversion of Petroleum Coke to Porous Materials

89

Previous findings of Raman analysis performed before and after ball-milling treatment indicated

that the mechanical energy during ball-milling introduced defects and created disordered

structure on petcoke [161, 162]. These newly developed defects were easier to adsorb oxygen

atoms and become carbon active site C(O), resulting in an improvement of CO2 gasification rate

of petcoke [161]. However, the pore volume of ball-milled AC was lower than physically mixed

AC in Figure 5-5b. An increasing number of defects after ball-milling did not improve the

reactivity of carbon oxygen reactions for higher pore volumes obtained. Therefore, a possible

explanation for a lower pore volume obtained with ball-milled AC could be less NaOH collected

after separating out zirconia balls from the ball-milled mixture. A similar discussion was covered

with KOH mixing (see Figure 5-2 and Table 5-1). But further investigation with NaOH mixing

need to be carried on.

5.2.4 Physically mixing and ball milling with different chemical ratios

Consistent with earlier findings [96] reported by Virla et al, more mesopores were created by

physically mixing when NaOH was used as the chemical agent compared to KOH. However, the

mechanism of pore development is still unknown when activating petcoke with different mixing

methods and various chemical amount. The following experiment was performed to understand

the impact of the amount of NaOH on AC porosity and yield when using either method of

physically mixing or ball-milling before activation.

The possible reactions between NaOH and petcoke are [5, 8, 81]:

Page 104: Conversion of Petroleum Coke to Porous Materials

90

𝑁𝑎𝑂𝐻 + 𝐶 → 𝐶𝑂 + 𝑁𝑎 +1

2𝐻2 Equation 5-1

6𝑁𝑎𝑂𝐻 + 𝐶 → 2𝑁𝑎 + 3𝐻2 + 2𝑁𝑎2𝐶𝑂3 Equation 5-2

4𝑁𝑎𝑂𝐻 + 𝐶𝐻2 → 𝑁𝑎2𝐶𝑂3 + 𝑁𝑎2𝑂 + 3𝐻2 Equation 5-3

𝑁𝑎2𝑂 + 𝐶 → 2𝑁𝑎 + 𝐶𝑂 Equation 5-4

𝑁𝑎2𝐶𝑂3 + 2𝐶 → 2𝑁𝑎 + 3𝐶𝑂 Equation 5-5

The mesopore volumes in Figure 5-6 decreased linearly with sodium to petcoke ratios. However,

the micropore volumes and yields were not linearly related to the ratios. Assuming the

occurrence of an ideal reaction during activation process, one mole of carbon consumes one

mole of sodium hydroxide (Equation 5-1). The theoretical yields should be 7%, 37% and 67%

with decreasing NaOH to petcoke molar ratios. However, the experimental yields were 53%,

57% and 74% respectively. The experimental yields were higher than the theoretical yields for

all of the samples, and the difference between experimental and the corresponding theoretical

yields became smaller (i.e., 46%, 20%, 7%) with decreasing NaOH amount. This indicated

undesired reactions occurred and reacted with NaOH rather than consuming carbon for pore

formation. For instance, one mole of carbon required two mole of NaOH to produce CO2

(Equation 5-2). In this case, less carbon was consumed with a same amount of NaOH as that

used in the ideal reaction. In addition, Na2O and Na2CO3 were considered as intermediates. The

incomplete reactions (from Equation 5-3 to Equation 5-5) reacted with a greater amount of

NaOH, which also wasted NaOH for the purpose of developing porosity. Since the biggest

difference between theoretical and experimental yields was presented at the highest NaOH to

petcoke molar ratio of 1.06, a greater number of undesired reactions were likely involved during

its activation process.

Page 105: Conversion of Petroleum Coke to Porous Materials

91

In order to make full use of the NaOH, ball-milling in a rotary tumbler was applied to mix

chemicals and petcoke more homogenously with various NaOH to petcoke molar ratios (ranging

from 0.35-1.06). However, for all the NaOH to petcoke molar ratios tested, the pore volumes and

yields of ball-milled samples did not change significantly compared to physically mixed

samples. Since the color of mixture became darker after ball-milling, the homogenous mixture of

NaOH and petcoke was achieved by ball-milling NaOH and petcoke together. However, when

petcoke was ball-milled with different amounts of NaOH, the petcoke particle size changed at

the same time.

Page 106: Conversion of Petroleum Coke to Porous Materials

92

PM-1.06NaOH

BM-1.02NaOH

PM-0.71NaOH

BM-0.71NaOH

PM-0.35NaOH

BM-0.35NaOH

Po

re V

olu

me

(c

m3

/g)

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Yie

ld (%

)

0

20

40

60

80

100

Figure 5-6. The impact of mixing methods and amount of NaOH on pore volumes (represented

by bars) and yields (represented by solid red circles) during NaOH activation. AC prepared with

1.06/0.71/0.35 NaOH to petcoke molar ratios by physically mixing (PM) or ball-milled (BM).

The black bar is micropore volume, and the grey bar is mesopore volume.

After 14 h ball-milling treatment, DI water was used to remove NaOH in the mixture to measure

particle size of petcoke. Summarized in Figure 5-7, after ball-milling with NaOH to petcoke at

molar ratios of 1.06/0.71/0.35, the majority of the particles became smaller. However, the peaks

of particle size distribution at highest (1.06) and lowest (0.35) NaOH to petcoke molar ratios

shifted from 150 µm to 40 µm, while less particle size reduction was observed for NaOH to

petcoke molar ratio of 0.71. The reduction of particle size before and after ball-milling could also

Page 107: Conversion of Petroleum Coke to Porous Materials

93

be clearly observed in SEM images by comparing between them (Figure 5-8) and with raw

petcoke (Figure 5-8a). Further activation (Figure 5-8d, f, h) did not change the particle size

significantly, but created more fine particles. In contrast to AC prepared from ball-milled

mixture with the same NaOH to petcoke ratio of 0.71, physically mixed AC had much larger

particles with distinct cracks on those particles.

Figure 5-7. Particle size distribution of raw petcoke and ball-milled samples mixed with NaOH

to petcoke molar ratios of 1.06/0.71/0.35 after washing but before activation.

Particle Size (µm)

0 50 100 150 200 250 300

Fre

qu

en

cy (

%)

0

10

20

30

40

50

60

B-Na

B-2Na

B-3Na

P-Petcoke

Raw petcoke

BM-0.71NaOH

BM-0.35NaOH

BM-1.06NaOH

Page 108: Conversion of Petroleum Coke to Porous Materials

94

Figure 5-8. SEM images of (a) physically mixed petcoke, (c, e, g)ball-milled petcoke with NaOH

to petcoke molar ratios at 0.35/0.71/1.06, and(d, f, h) AC after NaOH activation. (b) AC prepared

by physically mixed with NaOH to petcoke at molar ratio of 0.71.

a. Petcoke

c. BM_0.34NaOH

e. BM_0.71NaOH

g. BM_1.06NaOH h. AC_BM_1.06NaOH

b. AC_PM_0.71NaOH

f. AC_BM_0.71NaOH

d. AC_BM_0.34NaOH

Page 109: Conversion of Petroleum Coke to Porous Materials

95

5.2.5 Measurement of chemical diffused into petcoke cores

In order to study if the chemical agent reached the core of the AC particles, the activation of

large chunk of petcoke was completed by previous postdoc Dr. Montes in our group. The digital

photos in Figure 5-9 (b and d) clearly showed that petcoke activated with KOH and NaOH

formed a grey layer around the petcoke, and less chemicals remained at the bottom of ceramic

boat. In contrast, the K2CO3 powder stayed beneath the petcoke after activation, and there was a

distinct boundary between the chemicals and petcoke. The different phenomena may result from

the formation of a molten phase when increasing temperature during KOH and NaOH activation.

Therefore, petcoke at the top layer adhered with fluid chemicals, but the solid phase in K2CO3

prohibited the interaction with petcoke at the top surface. K2CO3 was recognized as a less

effective chemical agent to create porosity on petcoke compared to KOH or NaOH.

The pore volume of AC core and edge was determined by N2 adsorption. Only big chunks of

petcoke activated by KOH developed porosity on both edge and core. The average pore volume

on the edge was 0.5 cm3/g, which was higher than pore volume at the core (0.1 cm3/g). As

petcoke is a non-porous material, the porosity at the core of petcoke was developed by KOH.

Thus, KOH was able to diffuse and reach the core. However, it was hard for NaOH and K2CO3

to reach the core. The surface area was similar to raw petcoke (< 2 m2/g) both on the edge and at

the core for either NaOH or K2CO3 activation.

Page 110: Conversion of Petroleum Coke to Porous Materials

96

Figure 5-9. Large particles of petcoke mixed with (a) KOH, (c) NaOH, and (e) K2CO3, and

corresponding images (b, d and f) after activation at 800 ˚C for 30 min with a chemical to

petcoke mass ratio of 2.

To understand whether or not the pore formed without carbon consumption during activation, the

difference of the experimental pore volume and the pore volume by carbon oxygen reactions was

calculated in Table 5-3 (KOH) and Table 5-4 (NaOH). By assuming carbon consumption was the

only way to create porosity, the pore volumes through carbon oxygen reactions were calculated

by yield (mass of carbon consumed) according to Equation 5-6. The same bulk density of AC (1

g/cm3 petcoke-derived AC prepared at default activation condition) was used for all samples. The

experimental pore volumes in Table 5-3 were higher than estimated pore volumes for both KOH

activated and co-activated AC prepared from higher molar ratios of 0.43 and 0.64. The larger

difference between estimated and experimental pore volume indicated a greater possibility of

developing porosity without consuming carbon. However, all NaOH activated carbons either

with or without steam addition have a higher estimated pore volume than experiment pore

Page 111: Conversion of Petroleum Coke to Porous Materials

97

volume (Table 5-4). It may be an indication of NaOH developed pores in shrink core model. The

different results of KOH and NaOH suggested 1) NaOH developed pores based on the shrink

core model, which was also indicated by the data in Figure 4-2 and Figure 4-9; 2) K species were

more active than Na species when developing porosity without consuming carbon during

activation, which may also be a reason for KOH to reach the core but not for NaOH as the

discussion mentioned above (in Figure 5-9).

𝐸𝑠𝑡𝑖𝑚𝑎𝑡𝑒𝑑 𝑝𝑜𝑟𝑒 𝑣𝑜𝑙𝑢𝑚𝑒 = 1𝑔∗𝑌𝑖𝑒𝑙𝑑

𝜌𝐴𝐶 Equation 5-6

Table 5-3. Estimated and actual pore volumes for petcoke activated with KOH.

Sample

Estimated pore

volume by carbon

oxygen reactions

(cm3/g)

Experimental pore

volume (cm3/g) Difference

0.11KOH 0.43 0.18 -0.25

0.21KOH 0.50 0.37 -0.14

0.43KOH 0.58 0.67 0.09

0.64KOH 0.83 1.10 0.27

0.11KOH/St 1.35 0.15 -1.20

0.21KOH/St 1.40 0.57 -0.83

0.43KOH/St 1.33 1.08 -0.25

0.64KOH/St 1.30 1.38 0.08

Page 112: Conversion of Petroleum Coke to Porous Materials

98

Table 5-4. Estimated and actual pore volumes for petcoke activated with NaOH.

Sample

Estimated pore

volume based by

carbon oxygen

reactions (cm3/g)

Experimental pore

volume (cm3/g) Difference

0.18NaOH 0.23 0.01 -0.22

0.35NaOH 0.65 0.14 -0.51

0.71NaOH 1.03 0.55 -0.48

1.06NaOH 1.15 0.73 -0.42

0.18NaOH/St 1.05 0.05 -1.00

0.35NaOH/St 1.10 0.10 -1.00

0.71NaOH/St 1.20 0.43 -0.77

1.06NaOH/St 1.60 0.71 -0.89

5.2.6 Cross-section analysis

Cross-sections of non-washed sample were obtained by fracturing the solidified resin without the

extra polishing step. However, the cross-section has a tendency to break through the chemical

layers instead of the middle of AC particles as shown in Figure 5-10a and b. Therefore, no cross-

section was successfully made.

The EDX mappings (in Figure 5-10b and d) found chemicals (either K species or Na species)

around the particles. However, sulfur distribution on cross-sections of NaOH and KOH activated

carbon were different in Figure 5-10b and d. KOH activated carbon had a homogenous S cross-

section (Figure 5-10b), but appeared to be more rod-shaped with NaOH activated carbon (Figure

5-10d). The co-existence of Na and S could also be an indication of the formation of Na2S. A

higher melting temperature of Na2S (1176˚C) compared to K2S (840˚C) could be a possible

explanation for different sulfur distribution between the cross-sections of NaOH and KOH

Page 113: Conversion of Petroleum Coke to Porous Materials

99

activated carbon since the solid phase during NaOH activation limited the diffusivity of active

chemical species.

Figure 5-10. SEM images and potassium, sodium and sulfur mapping of non-washed AC from

petcoke with (a, b) 0.43KOH, and (c, d) 0.71NaOH. Note: the elements in image (b) were

alumina, sulfur and potassium, but (c) involved alumina, sulfur and sodium. The activation was

performed at 800 ºC for 30 min with chemical to petcoke mass ratio of 2 without washing step.

From the SEM images of the cross-section of non-washed AC activated with KOH in Figure 5-

11, the chemicals (in grey) were present in the cracks through the cross-section or around the

particles. However, not all of the cracks were occupied with chemicals since a certain amount of

Page 114: Conversion of Petroleum Coke to Porous Materials

100

chemicals were likely to be removed or dissolved into the oil solution while polishing with an

automatic machine. As the potassium mapping in Figure 5-11d was highly related to the grey

area in SEM image in Figure 5-11c, the evidence of K diffusion into the core of the particles was

present. Moreover, the presence of K was also related to S when comparing Figure 5-11d and e.

This result could imply the formation of chemical compounds of K and S. For instance, K2S was

reported as a product during KOH activation of high sulfur petcoke [27, 30].

Figure 5-11. SEM analysis of petcoke activated with KOH before wash.a-c) SEM images of

0.43KOH with different magnifications, (d) potassium mapping and (e) sulfur mapping. Petcoke

was activated with KOH to petcoke molar ratio of 0.43 at 800 ºC for 30 min.

Page 115: Conversion of Petroleum Coke to Porous Materials

101

After removing chemical layers with HCl (1M) and DI water, the cross-section of AC was

achieved by fracturing solidified resin. Compared to the smooth cross-section of raw petcoke

(Figure 5-12b), AC developed large cracks on the cross-section when activation with KOH to

petcoke molar ratios at 0.21 (Figure 5-12c and d) and 0.43 (Figure 5-12 e and f). As chemical

ratio reached 0.64, a distinct ‘onion-like’ structure could be observed in Figure 5-12g and h. A

similar structure was also discovered by Niasar et al. when activating petcoke at 750 ˚C for 3 h

with KOH to petcoke mass ratio of 3 (corresponding molar ratio of KOH to petcoke was 0.64)

[102]. The author considered that those cracks and the ‘onion-like’ structure benefitted pore

development at higher chemical ratios.

As shown in Figure 5-13b, the large cracks on cross-section of AC prepared during KOH

activation transformed into ‘onion-like’ structure when adding 60 min steam into the same

system. The change to an ‘onion-like’ structure is related to pore volume and pore size increment

with the addition of steam, since the addition of steam is like to widen the cracks and form the

‘onion-like’ structure. However, the structure did not change in this way at lower chemical

ratios; the structure transformation only occur if there were a sufficient number of cracks before

steam addition, in which case the expansion of the whole particle size led to the formation of an

‘onion-like’ structure. The structure transformation may be explained by the release of gases

through the reactions of steam with carbon (C + H2O → CO + H2). Gases trapped in the thicker

chemical layers; led to expanding particles. The expansion of particles exposed larger contact

surface area for K species to develop further porosity, leading to higher possibility for pore

formation.

Page 116: Conversion of Petroleum Coke to Porous Materials

102

Figure 5-12. SEM images of (a, b) raw petcoke and (c-h) AC prepared from KOH activation at

800 ˚C for 30 min with various KOH to petcoke molar ratios of 0.21/0.43/0.64.

a. Petcoke b. Petcoke

c. 0.21KOH d. 0.21KOH

e. 0.43KOH f. 0.43KOH

g. 0.64KOH h. 0.64KOH

Page 117: Conversion of Petroleum Coke to Porous Materials

103

Figure 5-13. The SEM images of (a, b) KOH activated petcoke at 800 ˚C for 30 min, and (c, d)

KOH steam co-activated petcoke at 800 ˚C for 60 min. The KOH to petcoke molar ratio was of

0.43 for both KOH and KOH steam activation.

5.2.7 Activation at lower temperatures

The total pore volumes and yields of the petcoke samples activated at different temperatures are

shown in Figure 5-14. At 400 ˚C, there was essentially no activation. As the activation

temperature increased from 400 to 800 ˚C, the yield appears to decrease linearly from 98% to

75%. The total pore volume, however, increased more rapidly to a temperature of 600 ˚C

(0.01375 cm3/g*min) than between 600 and 800 ˚C (0.00475 cm3/g*min). The pore volumes

a. 0.43KOH b. 0.43KOH

c. 0.43KOH/St_60 min d. 0.43KOH/St_60 min

Page 118: Conversion of Petroleum Coke to Porous Materials

104

were 0.55 cm3/g and 0.74 cm3/g at 600 ˚C and 800 ˚C, respectively. As shown in Figure 5-15, the

AC activated at temperature lower than 600 ˚C contained only micropores, mainly below 1.5 nm

in size. When the activation temperature increased to 700 ˚C and 800 ˚C, pores between 1.5 and

2 nm in size appeared.

Figure 5-14. Pore volume (●) and yield (×) of AC prepared from petcoke with increasing

temperatures. KOH activation of petcoke was performed with KOH to petcoke molar ratio of

0.43 at certain temperature without any holding time.

Temperature (˚C)

400 600 800

To

tal p

ore

vo

lum

e (

cm

3/g

)

0.0

0.2

0.4

0.6

0.8

Yie

ld (%

)

0

20

40

60

80

100

Page 119: Conversion of Petroleum Coke to Porous Materials

105

Figure 5-15. Pore size distributions of raw petcoke and AC (0.43KOH) prepared from petcoke

activated with increasing temperatures.KOH activation of petcoke was performed with KOH to

petcoke molar ratio of 0.43 at certain temperature without any holding time.

Within the same activation time (134 min) in Table 5-5, total pore volume was 0.27 cm3/g higher

but yield was 7% lower for AC prepared at 700 ˚C in contrast to that activated at 500 ˚C.

However, with the same activation temperature (500 ˚C), the similar pore volume (~0.36 cm3/g)

and yield (~94%) were obtained with 0 min (Figure 5-14) and 40 min holding time (Table 5-5) at

500 ˚C. Similar results were observed with AC activated at 600 and 800 ˚C within the same

activation time (Table 5-5). Therefore, the pore volumes and yields are more dependent on

activation temperature than activation time.

Pore width (nm)

0.0 0.5 1.0 1.5 2.0 2.5 3.0

dV

/dw

po

re v

olu

me (

cm

3/(

g*n

m))

0.0

0.5

1.0

1.5

2.0

Petcoke0.43KOH_400_00.43KOH_500_00.43KOH_600_00.43KOH_700_00.43KOH_800_0

0.0 0.5 1.0 1.5 2.0 2.5 3.0

0.0000

0.0002

0.0004

0.0006

0.0008

0.0010

0.0012

Page 120: Conversion of Petroleum Coke to Porous Materials

106

Table 5-5. Total pore volume and yield versus temperature for same activation time.

Temperature

(˚C)

Holding time

(min)

Total activation time*

(min)

Pore volume

(cm3/g)

Yield

(%)

500 40 134 0.37 93

700 0 134 0.64 86

600 40 154 0.59 85

800 0 154 0.74 75

*Total activation time indicates the entire experiment time, which includes the time for

increasing the temperature to activation temperature as well as the time for holding at certain

activation temperatures

Low temperature activation is a more efficient way to create pores for a limited time period. To

understand how the pores were developed on non-porous material petcoke at lower activation

temperature, XRD, Raman and FTIR analysis were performed for structural and surface analysis.

The XRD spectra in Figure 5-16 confirmed that raw petcoke had a graphitic structure, since there

is a sharp peak at ~25˚in the spectra, which is related to the (002) face of graphite. The graphite

peak disappeared after activation at 800 ˚C. The absence of the graphite peak can not confirm

AC became amorphous, since XRD has a limitation of crystallite size.

Page 121: Conversion of Petroleum Coke to Porous Materials

107

Figure 5-16. XRD patterns of (a) raw petcoke and (b) AC activated with KOH to petcoke molar

ratio of 0.43 at 800 ˚C without any holding time.

Raman spectra for petcoke and AC activated at 500 ˚C and 800 ˚C (Figure 5-17) had G band

located at 1581 cm-1 and D band located at 1350 cm-1. G band is also called single band, which is

a common feature for graphitic materials [163]. D band is for disordered (defects) materials

when a sample has a small crystallite size (La < 0.5 µm). The intensity of D band is proportional

to the amount of disorder in the sample, thus the ratio of ID/IG can be used for disorder [101] and

defect quantity in graphitic materials [102]. The ID/IG ratios beside Figure 5-17 were calculated

by using the peak area of these two bands. The maximum value of 4.74 appeared when activating

AC at 500˚C without any holding time. The highest ratio of ID/IG means the most disordered

0 10 20 30 40 50 600

5000

10000

15000

20000

25000

30000

b. 0.43KOH_800_0

In

ten

sit

y (

a.u

.)

2 Theta (degree)

a. Raw petcoke

Page 122: Conversion of Petroleum Coke to Porous Materials

108

structure and defects were present on the sample (0.43KOH_500_0). ID/IG = A/La, A is a constant

for a fixed laser excitation energy by using Tuinstra Koenig law [164]. An increased ratio also

indicates a smaller La (crystallite size). Therefore, when petcoke was activated at 500 ˚C, the

structure became disordered and crystallite size decreased at the same time. As reported in the

literature [43, 165], a disordered structure developed pores more easily during KOH activation

than an ordered structure. When higher activation temperature were applied, a more ordered

structure was formed with larger crystallite size. Thus, the pore volume increased in a slower rate

at higher activation temperatures (600-800 ˚C) in contrast to lower activation temperatures (400-

600 ˚C) in Figure 5-14.

Figure 5-17. Evolution of the Raman spectra of (a) raw petcoke and AC (0.43KOH) activated at

(b) 500 ˚C and (c) 800 ˚C. KOH activation of petcoke was performed with KOH to petcoke

molar ratio of 0.43 at certain temperature without any holding time.

500 1000 1500 2000 2500 3000 3500

0

400

800

1200

1600

2000

2400

2800

3200

3600

c. 0.48K_800

a. Raw petcoke

Inte

nsit

y (

a.u

.)

Raman shift (cm-1)

b. 0.48K_500

c. 0.43K_800_0

b. 0.43K_500_0

a. Raw petcoke

3.72

4.74

3.79

ID/IG ratios

Page 123: Conversion of Petroleum Coke to Porous Materials

109

Both Raman and XRD are able to observe a crystalline structural change of AC from petcoke.

The petcoke had a graphtic structure which was characterized by both Raman and XRD.

However, the graphitic structure was damaged when petcoke was activated at 500 ˚C. By

increasing the activation temperature to 800 ˚C, the structure became more ordered on the

Raman spectra, but the ordered structure was not seen on the XRD pattern.

The change of surface functional groups while activating petcoke with increasing temperature

(400-700 ˚C) was analyzed by FTIR. C-H groups (peaks at 700-900 cm-1) disappeared when

activation temperature increased from 400 ˚C to 500 ˚C. These groups were reported as very

reactive peripheral hydrocarbons in coal and coke. The absence of C-H groups indicated that

carbonization occurred at 400-500 ˚C [30]. The oxygen groups (broad peak at 1200-1300 cm-1)

were formed at lower activation temperatures (400-600 ˚C) [30]. Also within the same activation

temperatures from 400-600 ˚C, C=C in aromatics (peak at 1600 cm-1) become intense [166]. The

existence of these functional groups provided active sites during KOH activation. As the

intensity of each peak reduced when activation temperature was over 600 ˚C, the corresponding

functional groups disappeared. The lack of surface functional groups as active sites could be one

of the reasons why pores volumes increased slower when activating at higher temperature from

600-800 ˚C (Figure 5-14). Another reason for the slower rate of the pore volume growth may be

explained by the structure transformation from a disordered one to an ordered one, which slows

down the carbon oxygen reaction. Also, the reduction of surface groups at activation

temperatures from 600 to 800 ˚C resulted in an increase in carbon composition (80 wt% to 90

wt%) of AC (See Appendix C).

Page 124: Conversion of Petroleum Coke to Porous Materials

110

4000 3500 3000 2500 2000 1500 1000

0.00

0.05

0.10

0.15

0.20

Raw petcoke

0.43KOH_400_0

0.43KOH_500_0

0.43KOH_600_0

A

bs

orb

an

ce

Wavenumbers (cm-1)

0.43KOH_700_0

Figure 5-18. FTIR spectra (ATR detector) of raw petcoke and AC (0.43KOH) prepared with

increasing activation temperature. KOH activation of petcoke was performed with KOH to

petcoke molar ratio of 0.43 at certain temperatures without any holding time.

5.2.7 Density calculations

Porosity may be developed at lower activation temperatures by opening existing closed pores on

petcoke. In order to assess this hypothesis, the density of the petcoke was measured in the lab,

and compared it to calculated values. The bulk density of petcoke was calculated by Equation 5-

6 using pore volume, yield and the density of graphite (2.16 g/cm3). For AC activated at 500 ˚C

with KOH to petcoke molar ratio of 0.43, 0.3 cm2/g pore volume and 94% yield were obtained.

Assuming all the pores developed at 500 ˚C were by opening the initially closed pores of

Page 125: Conversion of Petroleum Coke to Porous Materials

111

petcoke, in other words, 1 g petcoke had 0.3 cm2 initial pore volume and graphite structure

occupied the rest of space. The calculated bulk density of petcoke is 1.36 g/ cm3. A similar

assumption was made for pores developed at 600 ˚C. The calculated bulk density of petcoke is

1.05 g/ cm3.

Equation 5-7

The measured bulk density of petcoke is 1.26 g/cm3. Since the measured petcoke density is

closed to the average of calculated bulk density (1.21 g/cm3), the pores developed at 500-600 ˚C

were mainly by opening the initial closed pores of petcoke.

As reported in literature [167], petcoke has a structure with graphite-like crystallites embedded in

an amorphous matrix. Petcoke in our lab had a density of 1.256 g/cm3, which was lower than the

density of graphite (2.16 g/cm3). In the modeling structure of petcoke proposed by Zhong et al.

[28], the obvious space within the petcoke structure could be observed. These empty spaces may

be the closed pores mainly in ultra-micropore range, which can only be detected by CO2

adsorption. Thus, the surface area of raw petcoke by CO2 adsorption was higher than N2

adsorption (see data in Table 5-6).

Both surface areas and pore volumes detected by CO2 adsorption increased once petcoke was

activated with KOH to 400 or 450 ˚C (See Table 5-6). It implied that ultra-micropores started to

develop even at low temperature before 400 ˚C. It is noted that N2 has a diffusion problem into

micropores smaller than 0.45 nm. It was hard for 0.43KOH_450_0 to get to equilibrium at a low

Page 126: Conversion of Petroleum Coke to Porous Materials

112

relative pressure range. The analysis could not start even after 48 h. Thus, CO2 adsorption was

suggested to detect micropores smaller than 0.7 nm [140].

Table 5-6. Surface area from N2 adsorption and CO2 adsorption.

Sample DFT surface area by N2

adsorption

DA surface area by CO2

adsorption (m2/g)

Total pore volume

by CO2 adsorption

(cm3/g)

Raw petcoke 2 20 0.0006

0.43KOH_400_0 2 55 0.0033

0.43KOH_450_0 N/A 88 0.0056

5.2.8 Total activation time

Figure 5-19. Pore volume (●) and yield (×) of AC (0.43KOH) prepared with extended holding

time at various activation temperatures from 500 to 800 ˚C. Note: AC (0.43 KOH) was prepared

with KOH to petcoke molar ratio of 0.43.

Activation time (min)

0 50 100 150 200 250 300

To

tal p

ore

vo

lum

e (

cm

2/g

)

0.0

0.2

0.4

0.6

0.8

Activation time (min)

0 50 100 150 200 250 300

Yie

ld (%

)

0

20

40

60

80

To

tal p

ore

vo

lum

e (

cm

2/g

)

0.0

0.2

0.4

0.6

0.8

1.0

Yie

ld (%

)

0

20

40

60

80

100

a. 0.43KOH_500 b. 0.43KOH_600

c. 0.43KOH_700 d. 0.43KOH_800

Page 127: Conversion of Petroleum Coke to Porous Materials

113

The total pore volumes and yields changes with extended holding time at various activation

temperatures, as shown in Figure 5-19. As the holding time increased from 0 to 240 min, the

largest pore volume increment (0.18 cm3/g) and biggest yield drop (10%) were achieved at 700

˚C activation. The pore volume increased by ~0.1 cm3/g without reduction of yield at 500 ˚C

activation by extending holding time from 0 to 240 min. A similar trend was followed by AC

activated at 800 ˚C. Pore volume increased ~0.1 cm3/g, but only 4% yield dropped within the

240 min holding time. These results indicated that possible ways of pore formation may exist

without carbon consumption.

One mechanism of pore formation could be intercalation. Intercalation is beneficial for pore

development, because no carbon is consumed with an expanded carbon layer by active species

like metallic K or Na. As a consequence, the pore volume increased with a high product yield.

However, the in-situ techniques are limited to detect these intercalated species formed at certain

temperatures from 400-800 ˚C, especially when considering high sulfur releasing with the

involvement of a corrosive chemical agent like KOH or NaOH. Therefore, there is no direct

evidence that intercalation is actually happening during activation of petcoke. Since I have only

observed sparks while washing AC samples prepared from 800 ˚C, the formation of K or K2O

happened at high temperature activation. Similar results were reported in several research groups

that the intercalated species formed at temperature higher than 700˚C by analyzing XRD of non-

washed AC after cooling down [78, 81].

The other way to develop pores with only a limited amount of carbon consumed is exposing

closed pores on raw petcoke. As reported in literature by Xiao et al [168], the presence of H2 in

Page 128: Conversion of Petroleum Coke to Porous Materials

114

the carbonization process reacted with surface heteroatom species (S, O, N) of petcoke, and then

formed –CH- and –CH2- species which are ‘active sites’ for further developed pore volume.

When activation temperature reached 500 ˚C, the active species were formed for further pore

development. At lower temperatures, pores could develop from opening the initially closed pores

of raw petcoke. When extending holding time at 800 ˚C, the closed pores of petcoke may be

continuously opened from where originally covered by heteroatom species. Another reason for

increasing pore volume without consuming too much carbon is that the adequate cracks provided

more channels for large amount of K active species to diffuse into the petcoke core, therefore, a

growing number of carbon surface were exposed and attached to K species. Then, pores could be

developed by either oxidation reaction or intercalation.

The corresponding pore size distribution was listed in Figure 5-20. For lower temperature

activation (400-600 ˚C), the pore size distribution kept similar shape. But the intensity of the

peak increased with increasing activation holding time. During higher temperature activation

(600-800 ˚C) and extended activation time, pores with sizes larger than 1.5 nm appeared with.

However, the intensity of peak increased within the first 40 min of activation time, and then

decreased when the activation time was extended to 240 min. The smaller pores may transform

to bigger pores with sizes larger than 1.5 nm.

Page 129: Conversion of Petroleum Coke to Porous Materials

115

Figure 5-20. The pore size distribution of AC (0.43KOH) activated at (a) 500 ˚C, (b) 600 ˚C, (c)

700 ˚C and (d) 800 ˚C with holding time from 0-240 min. (without any holding time (), 40 min

holding time (), and 240 min holding time ()).

5.3 Pore development with chemical activation of high sulfur petcoke

As KOH melted at approximately360 ˚C, the oxidation reaction of KOH and carbon were mainly

solid-liquid reactions. Fixed sulfur (S, 6.5 wt%) was removed from the petcoke before ultra-

micropores started developing on the petcoke at 400 ˚C. KOH was first reacted with S to

transform fixed organic sulfur into the inorganic sulfur species (sulfate, sulfite, K2S) or gases

(H2S, SO2). The remaining KOH in liquid phase was then diffused into petcoke particles.

Though petcoke is a non-porous material, the partially ordered structure of petcoke (reported in

Pore width (nm)

0 1 2 3

dV

/dw

po

re v

olu

me (

cm

3/(

g*n

m))

0.0

0.2

0.4

0.6

0.8

1.0

0 min hold time40 min hold time 240 min hold time

Pore width (nm)

0 1 2 3 4

0 hold time40 min hold time240 min hold time

dV

/dw

po

re v

olu

me (

cm

3/(

g*n

m))

0.0

0.2

0.4

0.6

0.8

1.0

1.2

0 min hold time 40 min hold time240 min hold time

0 min hold time40 min hold time 240 min hold time

a. 0.43KOH_500 b. 0.43KOH_600

c. 0.43KOH_700 d. 0.43KOH_800

Page 130: Conversion of Petroleum Coke to Porous Materials

116

the literature) may provide some defects acting as ‘active sites’ for the reactions to start from.

Pores (0.3 cm3/g) were developed with almost no carbon consumption at 500 ˚C. The opening of

initial closed pores on petcoke is a possible way develop to porosity on petcoke at low activation

temperature (400-600 ˚C). It should be noted that the existence of initial closed pores on petcoke

was assumed according to the experiment data and calculated petcoke density. There is no direct

evidence in the literature that mentions the closed pores within the petcoke structure. Further

study is required. Functional groups (oxygen groups, C=C) appeared with the development of a

higher disordered carbon structure when activating petcoke at low activation temperatures (400-

600 ˚C). These functional groups could be ‘active sites’ for continuous oxidation reactions. The

disordered carbon structure made it easier for the petcoke to react compared to the ordered

graphitic structure. This explained the reason why AC prepared at low activation temperature

(400-600 ˚C) has a higher rate of pore volume increase than AC from high activation

temperatures (600-800 ˚C).

An obvious yield drop was observed when petcoke was activated at temperatures from 600 to

800 ˚C. The oxidation reactions started to take the lead in this stage. The majority product of

carbon oxygen reactions was K2CO3. Other K species including metallic K, K2S, K2O may be

present as well. From the literature it was reported that the metallic K, as an active intercalation

species, formed at temperature above 700 ˚C. However, the AC samples used for determination

of K species were collected after cooling and exposure to air. The active species are likely to

change under these conditions. There currently is no in-situ technique available to determine the

formation of K species during petcoke activation due to the high experimental temperature and

Page 131: Conversion of Petroleum Coke to Porous Materials

117

the corrosive chemicals (e.g. KOH). The oxidation reactions and the reaction to form metallic K

are thermodynamically favored even at low temperatures, from 500 to 600 ˚C. The standard

reduction potential at 25 ˚C for K is -2.92 V, which implies that metallic K is a very active

reducing agent [77]. Therefore, metallic K may not be stable at these activation temperatures,

and so it is still not clear whether metallic K formed at activation temperatures from 400 to 800

˚C. Furthermore, whether the intercalation occurred through metallic K during petcoke activation

is even more difficult to investigate. As the surface functional groups have already formed at

temperatures from 400 to 600 ˚C, the heteroatoms on the surface may react more easily with

KOH rather than ordered carbon on petcoke. The pores may be continuously exposed by opening

of the closed pores on petcoke which are originally covered by the surface heteroatom species.

The particle expansion, resulting from adequate cracks created on petcoke, is another possible

reason for pore development at 800˚C with high yields (>70%), because the additional carbon

surface increased the possibility for carbon ‘active sites’ to attach active K species.

5.4 Summary

In summary, the C-S bond on petcoke broke before pores started to develop at 400 ˚C, indicated

by the loss of S content on EDX mapping. Meanwhile, the accessible channels were formed.

When KOH melted at around 360 ˚C, KOH became mobile and diffused into these channels.

Higher initial sulfur content on petcoke suggested that more accessible channels could be

created. A greater amount of KOH was likely to be diffused into the petcoke core. When

increasing activation temperature from 400-450 ˚C, the original closed ultra-micropores of

petcoke were gradually exposed, and the surface area was increased and then determined by CO2

Page 132: Conversion of Petroleum Coke to Porous Materials

118

adsorption. Only if the pores were accumulated to sizes larger than 1 nm at the activation

temperature of 500 ˚C, pore volume and surface area could be characterized by N2 adsorption.

However, the pores developed at 500 ˚C were smaller than 1.5 nm. A high yield (94%) was

obtained with high pore volume (0.3 cm3/g) and surface area (800 m2/g). Since almost no carbon

was consumed by carbon oxygen reactions, the most pores develop at 500 ˚C were by opening

initially closed pores on petcoke. At the same activation temperature, the disordered structure

and an increasing number of defects developed, resulting in a smaller crystallite size.

Additionally, when activation temperatures increased from 400-600 ˚C, several types of surface

functional groups (oxygen groups and C=C) appeared. Both functional groups and defects may

perform as ‘active sites’ for reactions to develop more porosity. The interfacial reactions are

possible to adhere onto these ‘active sites’ at a lower temperature, resulting in a rapid increasing

of pore volume in a limited time. Carbon oxygen reactions began to take the lead of the pore

development when activation temperature was increased to 700 ˚C. The pore volume

continuously increased to 0.64 cm3/g, but yield reduced to 82%. As the temperature reached 800

˚C and holding time was extended from 0-240 min, the pore volume increased by 0.1 cm3/g with

only 4% carbon consumed. The pores created during extended holing time may from either

intercalation or continuously opening of the closed pores on petcoke (usually originally covered

by the surface heteroatom species). The particle expansion with highest chemical amount or with

addition of steam increased the possibility for additional carbon surface to attach to K active

species, and thereafter, higher porosity were developed.

Page 133: Conversion of Petroleum Coke to Porous Materials

119

Chapter Six: HIERACHICAL POROUS CARBON FROM ASPHALTENES AND

PETCOKE

In this chapter, the microporous AC prepared from petcoke in Chapter 4 and Chapter 5 were

combined into a carbon foam material with nanoscale and macroscale pores. To develop

macroscale pores in interparticle space, a salt template was used. A specific role for each

component (salt, asphaltene or carbon source) in the salt template method was investigated. The

experimental conditions using a salt template, including ball-milling environment, carbon source

and carbonization temperature, were also studied. Asphaltenes were used as natural binders for

the first time to fabricate carbon foam and hierarchical porous carbon (HPC). The following is

the schematic of preparation procedure.

Ball-mill

in isopropanol 9h

Press

5 MPa 20 s

Carbonization

5 MPa 20 s

Wash

H2O with/without HCl

Carbon foam

Carbon source (ball-milled petcoke/ CB)

Binder (Asphaltene)

Pore former (NaCl)

Mixture

Ball-mill

in isopropanol 9h

Press

5 MPa 20 s

Carbonization

5 MPa 20 s

Wash

H2O with/without HCl

HPC

Carbon source (Petcoke-derived AC)

Binder (Asphaltene)

Pore former (non-washed K salts)

Mixture

Page 134: Conversion of Petroleum Coke to Porous Materials

120

6.1 Initial Experiments

Emulsion template was the first synthesis approach for carbon foam preparation (See Appendix

D). The water droplets were suspended in the oil phase which comprised of asphaltenes and

toluene. A stable mixture was then formed in the aqueous phase. However, the mixture became

instable when adding petcoke as the carbon source into the emulsion system. Furthermore, the

carbon structure collapsed after the sample was dried in the vacuum oven. Therefore, the

emulsion template was not a successful approach to prepare carbon foam.

In order to study if asphaltenes could work as a binder, the mixture of 1 g petcoke (after ball-

mill) and 4 g salt (NaCl) was wet ball-milled with zirconia balls in isopropanol for 9 h either

with or without asphaltene (1 g). After removing the zirconia balls, the mixture was dried in a

muffle furnace at 70 ˚C for 12 h. The powder was then applied with a hydraulic pressure of 5

MPa for 20 s. A monolith could only be formed with the addition of asphaltenes (in Figure 6-1a).

Otherwise, the mixture was still powdered (in Figure 6-1b). Therefore, asphaltenes were

introduced as binders for the first time to synthesize the monolith.

Page 135: Conversion of Petroleum Coke to Porous Materials

121

Figure 6-1. (a) A pellet of ball-milled petcoke, asphaltene and NaCl (mass ratio is 1:1:4), and (b)

mixed powder of ball-milled petcoke and NaCl (mass ratio of 1:4) after applying a hydraulic

force of 5 MPa for 20 s.

To study the role of salt in carbon foam preparation, 1 g petcoke and 1 g asphaltenes were mixed

and carbonized in N2 at 400 ˚C for 2 h either with 4 g of NaCl (in Figure 6-2a) or without NaCl

(in Figure 6-2b). Even though both mixtures with or without NaCl formed similar monoliths

after applying a hydraulic pressure of 5 MPa for 20 s, the monolith was only retained after

carbonization when NaCl was present. Thus, salt (NaCl) held the monolithic structure during

carbonization process. As reported by Gray et al., Athabasca asphaltenes were melted at an

average temperature of 224 ˚C, which was lower than asphaltenes decomposition temperature

(350 ˚C) [169]. NaCl remained as solid at 400 ˚C during carbonization. And NaCl did not

decompose because of its chemical stability. Asphaltenes in molten phase provided a mobile

environment for tiny petcoke particles to move around (particle size of 1.5-2.0 µm), but not for

large NaCl particles (initial particle sizes various from 100-500 µm). The monolith structure was

fixed by large NaCl particles in a well-dispersed molten mixture. However, for the system

without NaCl, only a thin layer was formed after releasing gases through asphaltenes

decomposition process.

Page 136: Conversion of Petroleum Coke to Porous Materials

122

Figure 6-2. Carbon materials after carbonizing (a) ball-milled petcoke, asphaltenes, NaCl (mass

ratio is 1:1:4) and (b) ball-milled petcoke and asphaltene(mass ratio of 1:1) in N2 at 400 ˚C for 2

h. The corresponding carbon foams were (c) with NaCl or (d) without NaCl after removing

NaCl in hot water, respectively.

Comparing to the SEM images of carbon foam with NaCl (Figure 6-2c) and without NaCl

(Figure 6-2d) after washing with hot water, the macroscale pores in carbon foam were only

formed with the presence of NaCl after removing the salt template (NaCl).

6.2 Parameters in salt template method

6.2.1 Ball-mill environment (Wet ball-mill vs Dry ball-mill)

Wet ball-mill was achieved by introducing isopropanol in a milling jar for 9 h, whileas dry ball-

mill was achieved with the absence of isopropanol. After pressing the mixtures of ball-milled

petcoke, NaCl and asphaltenes into a monolith, the monolith was then carbonized at 400 ˚C for 2

h. The SEM images of monoliths after washing were carbon foam (Figure 6-3). The macropores

Page 137: Conversion of Petroleum Coke to Porous Materials

123

in the interparticle space of carbon foam was larger via dry ball-mill process (Figure 6-3b) than

wet ball-milling (Figure 6-3a). This indicated that larger NaCl crystals remained after the dry

ball-milling. Jung et al. also found a much broader size reduction by using the dry ball-mill

technique in contrast to the wet ball-mill [106]. The wet ball-mill technique provided a better

dispersion of NaCl and asphaltenes around petcoke particles in the molten phase. The

electrostatic force between particles was weaker when isopropanol presented at the same time,

which avoided the agglomeration of particles. A homogenous mixture was obtained by wet ball-

milling [170]. Additionally, the isopropanol in the wet ball-milling also reduced the local

temperature during the ball-mill process. Thus, wet ball-milling was applied in the following

experiment in this section.

Page 138: Conversion of Petroleum Coke to Porous Materials

124

Figure 6-3. SEM images of carbon foam prepared by salt template with ball-milled petcoke,

asphaltenes and NaCl mixing at a mass ratio of 1:1:4 in (a, c) wet ball-mill (b, d) dry ball-mill.

6.2.2 Carbon source (carbon black vs ball-milled petcoke)

Carbon black (Monarch 120 from Cabot Cop, Boston, MA) was selected as another carbon

precursor compared to petcoke, because it has over 99 wt% carbon content and regular shape.

a b

c d

Page 139: Conversion of Petroleum Coke to Porous Materials

125

Larger macroscale pores between carbon particles were observed on the cross-section SEM

images of ball-milled petcoke (in Figure 6-4 b, e, f), which may be due to the bigger particle

sizes of ball-milled petcoke. Carbon black has a particle size in nanometer scale (8-100 nm);

however, the particle size for ball-milled petcoke is on the micrometer scale (1.5-2.0 µm). Bigger

particle sizes of petcoke also resulted in more cracks (shown in Figure 6-3 b) and less

mechanical strength (tested by hand) of the carbon foam structure.

Page 140: Conversion of Petroleum Coke to Porous Materials

126

Figure 6-4. SEM image of carbon foam using (a, c, e) carbon black as a carbon source, or (b, d,

f) ball-milled petcoke as a carbon source. The carbon foam was prepared from a mass ratio of

carbon precursor, asphaltenes and NaCl of 1:1:4, and then carbonized in N2 at 400 ˚C for 2 h.

6.2.3 Carbonization temperature (400 vs 600 ˚C)

a b

c d

e f

Page 141: Conversion of Petroleum Coke to Porous Materials

127

The variation of carbonization temperature from 400-600 ˚C did not have an impact on carbon

foam structure (Figure 6-5). Therefore, 400 ˚C was selected for further experiments to prepare

the carbon foam with less energy consumption.

Figure 6-5. SEM image of carbon foam by carbonizing carbon black, asphaltenes and NaCl (a, c)

at 600 ˚C for 2 h, or (b, d) 400 ˚C for 2 h The carbon foam was prepared from a mass ratio of

carbon black, asphaltenes and NaCl of 1:1:4.

a b

c d

Page 142: Conversion of Petroleum Coke to Porous Materials

128

6.3 HPC preparation

6.3.1 Carbon foam structure by using petcoke derived AC as carbon precursor

By using petcoke derived AC as a starting material, the hierarchical structure was built according

to the synthesis approach of carbon foam. NaCl was the pore-forming agent for interconnected

porosity. Asphaltenes work as binders to build interparticle space between AC particles.

However, the AC particles (Figure 6-6 b, d, e) were too big (25-250 µm) for a mechanically

strong monolith.

Page 143: Conversion of Petroleum Coke to Porous Materials

129

Figure 6-6. SEM image of HPC created by carbonizing 400 ˚C for 2 h (a, c, e) ball-milled

petcoke, asphaltenes and NaCl (mass ratio = 1:1:4) or (b, d, f) petcoke-derived AC (0.43KOH/St)

with asphaltenes and NaCl (mass ratio = 1:1:4). Note: AC (0.43 KOH/St) was prepared with

KOH steam co-activation at 800 ˚C for 30 min with KOH to petcoke molar ratio of 0.43.

a b

c d

e f

Page 144: Conversion of Petroleum Coke to Porous Materials

130

It was hypothesized that by introducing micro-mesoporous AC derived from petcoke, the

macroscale pore sizes of carbon foam would be integrated with nanometer pores on AC particles.

The initial total pore volume of AC was 0.98 cm3/g with a surface area of 2100 m2/g (DFT).

However, only 0.1 cm3/g pore volume and 180 m2/g surface area remained after the formation of

HPC. When comparing pore size distribution of HPC and the original AC (Figure 6-7), it was

seen that most pores in micropores (< 2 nm) and small mesopores (2-4 nm) on AC were blocked

by asphaltenes during wet ball-milling or carbonization process.

Figure 6-7. Pore size distribution of AC (●) prepared from petcoke through co-activation of

KOH and steam (0.43KOH/St), and the corresponding HPC (○) by using the same AC as carbon

precursor.Note: AC (0.43 KOH/St) was prepared with KOH steam co-activation at 800 ˚C for 30

min with KOH to petcoke molar ratio of 0.43.

6.3.2 HPC by using non-washed AC as carbon precursor

Wu et al. reported that a chemical layer (mainly KOH and K2CO3) was formed around the

microporous AC during KOH activation of petcoke [13]. In order to avoid the asphaltene

Pore width (nm)

0 1 2 3 4 5

dV

/dw

po

re v

olu

me (

cm

3/(

g*n

m))

0.0

0.1

0.2

0.3

0.4

0.5

AC HPC

Page 145: Conversion of Petroleum Coke to Porous Materials

131

blockage and make use of the chemical species inside micropores, non-washed AC was used as

an alternative carbon precursor for HPC preparation because the micropores of AC were

protected by chemical species before HPC structures was built. Since the literature showed that

asphaltene has a preference to adsorb into mesopores rather than micropores [96]. Highly

microporous AC without any mesopores (0.43KOH) was selected for the following experiments.

The exclusion of mesopores on AC lowered the possibility of asphaltene blockage into pores

during ball-milling.

The surface area for non-washed AC (0.43KOH) was lower than 2 m2/g, which confirmed that

all the micropores (Vmicro = 0.84 cm3/g) were blocked with K salts. After wet ball-milling in

isopropanol with asphaltene, the surface area was similar (1.8 m2/g) to what it was before ball-

milling. EDX mapping in Figure 6-8g showed 54 wt% K was dominant on the surface of the

particle. Oxygen and sulfur content were 42 wt% and 4 wt%, respectively. A monolith was

formed when pressing at 5 MPa for 20 s, and the physical shape of the monolith (Figure 6-8b)

did not change after carbonizing at 400 ˚C for 2 h. The additional pressure compressed the

powder mixture into the monolith firmly, which was hard to break manually. The EDX mapping

of the monolith (h) indicated a lower potassium content (46 wt%) but a higher oxygen content

(51 wt%) presented on the surface after carbonization. After 1 h of sonication and hot water

washing, the monolith broke into several small chunks. Only 16 wt% of K remained on the

coarse surface (Figure 6-8f) of chunks after washing. However, a high potassium amount (43

wt%) was detected on the smooth surface (Figure 6-8f), which is an indication of remaining K

salts in the HPC structure. Therefore, by continuous acid washing, almost all K was washed

away (< 0.5 wt%). Over 90 wt% C was exposed on the external surface. The sulfur left was only

Page 146: Conversion of Petroleum Coke to Porous Materials

132

2.2 wt%. The surface area for HPC after the acid wash increased from 3.9 m2/g (only water

wash) to 74 m2/g. However, the surface area after continuous acid washing was still lower than

original AC (2300 m2/g for 0.43KOH).

After wet BM After press and carbonize After H2O wash and dry

Figure 6-8. Digital photos (a-c) and SEM images (d-f) with EDX mapping (g-i) of HPC prepared

from non-washed AC (0.43KOH) with 1:1 mass ratioof AC and asphaltenes. Note: Red

indicates potassium in EDX mapping. AC (0.43 KOH) was prepared with KOH activation at 800

˚C for 30 min with KOH to petcoke molar ratio of 0.43.

5 mm 5 mm 5 mm

200 µm 200 µm 200 µm

50 µm 50 µm 50 µm

a

)

b c

d e f

g h i

Page 147: Conversion of Petroleum Coke to Porous Materials

133

As most of the micropores may have been blocked by asphaltene, the mass ratio of asphaltene to

AC was then reduced from 1:1 to 0.1:1. A similar experiment was performed with lower

asphaltene amount. The result showed that monolith shape stayed the same after carbonization at

400 ˚C for 2 h (Figure 6-9b). With a water wash only, potassium was removed from the surface

of HPC by using lower asphaltene amount (< 2 wt% K detected by EDX mapping). And the

sulfur content was also reduced to 1.3 wt%. The surface area after water washing increased to

830 m2/g. Meanwhile, the total pore volume increased to ~ 0.3 cm3/g.

AC + Asp (R=1:1) AC + Asp (R=1:0.1) Asp without AC (R=0:1)

Figure 6-9 Digital photos and SEM images of HPC prepared from non-washed AC (0.43KOH)

and asphaltene at mass ratios of 1:1 (a, d) and 1:0.1 (b, e). Note: AC (0.43 KOH) was prepared

with KOH activation at 800 ˚C for 30 min with KOH to petcoke molar ratio of 0.43.

Without any AC involved in the entire process, asphaltene was directly activated with KOH at a

mass ratio of 2 (Figure 6-9c and f). But there was no porosity developed after activation at 400

a b c

d e f

50 µm 50 µm 50 µm

1 cm 1 cm 1 cm

Page 148: Conversion of Petroleum Coke to Porous Materials

134

˚C for 2 h (the same experiment condition was applied as HPC preparation). The monolith did

not maintain its physical shape when the temperature increased. Only a layer was left in the

ceramic boat (Figure 6-9c).

The pore size distribution in Figure 6-10 contained extra peaks that appeared at 0.5 nm and 1.3

nm for the lower asphaltene amount (R=1:0.1). As for the high asphaltene amount (R=1:1), there

was only one peak at 0.5 nm after water and acid washing. Therefore, the occupied pores could

be exposed by either continuous acid washing or lowering the asphaltene amount. The pore

exposure was also related to the DFT surface area detected by N2 adsorption. After continuous

acid washing, the DFT surface area increased from 1.8 m2/g to 74 m2/g. And the DFT surface

area for lower asphaltene amount was almost 830 m2/g. A schematic of pore development is

given in Figure 6-11.

Page 149: Conversion of Petroleum Coke to Porous Materials

135

Figure 6-10. Pore size distribution of HPC prepared from non-washed AC (0.43KOH) and

asphaltenes at mass ratios of 1:1 (● for water wash, ○ for water and acid wash) and 1:0.1( for

water wash). Note: AC (0.43 KOH) was prepared with KOH activation at 800 ˚C for 30 min with

KOH to petcoke molar ratio of 0.43.

Pore width (nm)

0.0 0.5 1.0 1.5 2.0

dV

/dw

po

re v

olu

me (

cm

3/(

g*n

m))

0.0

0.2

0.4

0.6

0.8

1.0HPC after water wash (R=1:1)HPC after water and acid wash (R=1:1)HPC after water wash (R=1:0.1)

Page 150: Conversion of Petroleum Coke to Porous Materials

136

Figure 6-11. The schematic graph of pore development for HPC structure.

Salt (K species) AC Aphaltene

Page 151: Conversion of Petroleum Coke to Porous Materials

137

Chapter Seven: Conclusions and Recommendations

7.1 Conclusions

AC was successfully prepared using either NaOH or KOH activation with petcoke (6.5 wt%).

The mixing method for KOH/NaOH and petcoke before activation did not have a significant

impact on produced AC properties (pore volume and yield). Therefore, directly putting

NaOH/KOH pellets (about 6 mm diameter and 2 mm length in the form of a cylinder) on the top

of petcoke particles gave the same results as physically mixing by hand, while ball-milling

resulted in high dispersion, but separation of the material from the zirconia balls was difficult. In

both NaOH and KOH activation processes, the pore volume was positively related to an

increasing amount of chemical agent, but the yield showed a negative trend. The total pore

volumes of KOH activated samples (0.16-1.1 cm3/g) were higher than that activated with NaOH

(0.01-0.73 cm3/g). Also, activation with KOH resulted in highly microporous material, while

activation with NaOH resulted in micro-mesopores. By introducing steam as another oxygen

source, the chemical amount was reduced without compromising pore volume when using KOH,

but was not for NaOH. The AC yields with co-activation decreased to < 45% because of steam

gasification at 800 ˚C. A linear decrease in the yield was observed with increasing steam time (0

to 120 min) for all KOH to petcoke ratios.

A chemical layer was seen on the cross-sections of particles after activation but before washing

through SEM/EDX mapping. Large cracks on both external surface and cross-section of AC

(SEM images) were exposed once chemical layers were washed. An ‘onion-like’ structure was

formed under extensive activation conditions: for instance, chemical activation with highest

Page 152: Conversion of Petroleum Coke to Porous Materials

138

KOH to petcoke ratio of 0.64 or co-activation for 60 min with KOH to petcoke ratio of 0.43. The

transformation from cracks to the ‘onion-like’ structure may be related to the release of gases by

the interactions between steam and chemical species. The thickness of chemical layers around

particles corresponded with the steam gasification rate, and thus the amount of carbon consumed

by the addition of steam decreased with an increasing amount of chemical agent.

The study of KOH activation at various activation temperatures ranging from 400-800 ˚C

showed ultra-micropores started to develop at 400 ˚C on petcoke. An increasing pore volume

with almost no yield drop was achieved at 500 ˚C. At this temperature of activation, pores were

possibly developed by opening initially closed pores of petcoke. Carbon-oxygen reaction started

to take the lead when temperature increased to 700 ˚C. By extending holding time at 800 ˚C,

pores may also be developed continuously either by intercalation or further exposure of the

initially closed pores.

Finally, HPC was successfully fabricated by using asphaltenes as binders between petcoke-

derived AC. The macroscale pores within interparticle space were combined with the already

produced nanoscale pores on petcoke by using a salt template.

7.2 Recommendations

Considering the findings of this study, the following work is suggested

(a) Sulfur evolution of activation process

For petcoke with higher sulfur content, it is always appealing to consider. To begin,

understanding how sulfur species were transformed during the activation process will be crucial.

Page 153: Conversion of Petroleum Coke to Porous Materials

139

Sulfur evolution for the entire activation process is needed. SEM/EDX is only a surface

characteristic technique. Instead, elemental analysis is suggested for an accurate sulfur

determination of bulk materials. The specific sulfur species could be analyzed by X-ray

photoelectron spectroscopy (XPS) on the surface. Additionally, the amount of sulfur in gases

will be trapped in water. The sulfur in filtrates could be detected by inductively coupled plasma

(ICP). In this way, it may be possible to study the approximate temperature for C-S breakage and

possible reactions involving several sulfur species. The understanding of these fundamentals will

guide ways to utilize sulfur in either petcoke or asphaltenes.

(b) HPC with petcoke and asphaltenes

In order to optimize experiment conditions of HPC preparation, systematic experiments should

be designed. For instance, the mechanical strength of the monolith is required to be measured by

tensile test. The washing step needs to be modified in order to wash most of K species but

maintain the monolith shape. As Chapter 6 showed, the amount of asphaltene is also an

important parameter for the exposure of the nanoscale pores on petcoke. Further investigation of

the relationship of the amount of asphaltene and micropore volumes occupied by salts (K

species) will also be required. Other parameters are suggested as well. For instance, lower

activation temperature (200 ˚C).

(c) Life-cycle assessment for KOH activation and KOH and steam co-activation of petcoke

From an economic evaluation on AC production from several feedstocks, petcoke was proposed

as the most cost-effective raw material [25]. However, the most effective petcoke activation

approach is still not clear. Regarding pore volume, KOH and steam co-activation proved to be an

Page 154: Conversion of Petroleum Coke to Porous Materials

140

effective approach to reduce KOH amount. However, the addition of steam consumed a greater

amount of carbon and required more energy as well. Therefore, performing a life-cycle

assessment for different activation methods will provide information on the most effective way

to activate petcoke.

(d) Petcoke structure

As shown in XRD and HRTEM, petcoke is a partially ordered material with no porosity on the

surface. The existence of closed pores on petcoke was suggested in Chapter 5, because high

yields (>86%) of AC were prepared from petcoke activated at low temperatures. However, there

is no direct evidence to illustrate the porous structure in petcoke. Micro-computed tomography

(micro-CT) is a 3D imaging technique with high resolution (~ 2 µm). The internal structure of

petcoke can be observed without destroying the sample. Micro-CT has already been applied in

various geological applications [171, 172]. Several researchers captured micrometer-sized

features in coal using Micro-CT [173-175]. The density and porosity of large petcoke particles

are also suggested in order to understand better the initial porous structure of the petcoke.

(e) K species on AC

Determining the formation of specific K species during petcoke activation is essential for the

understanding of reactions involved and whether intercalation happened in the process to

develop pores. As a result, designing petcoke-derived AC with desirable properties will be

easier. XRD and XPS are suggested techniques to determine K species on petcoke before

washing at each activation temperature.

Page 155: Conversion of Petroleum Coke to Porous Materials

141

(f) Ball milling mixing with NaOH and petcoke

The titration data in Table 5-1 showed that ball milling reduced KOH amount before activation.

Similar trend may be followed with ball milling mixing of NaOH and petcoke as well. The initial

NaOH to petcoke ratio should be increased to 3. The same titration experiment is required to

measure the exact amount of NaOH in ball-milled mixture before activation. Then, AC

properties after ball milling with NaOH to petcoke of 3 can compare with the other two mixing

methods (physical mixing and no mixing) with a lower NaOH to petcoke of 2.

Page 156: Conversion of Petroleum Coke to Porous Materials

142

Reference

[1] CAPP, Crude Oil Forecast, Market and Transportation report, in, 2019.

[2] D.T. Lorne Stockman, Stephen Kretzmann, Petroleum coke: The Coal Hiding in The Tar

Sands, in, Washington DC, USA, 2013.

[3] P. Redelius, Asphaltenes in Bitumen, What They Are and What They Are Not, Road

Materials and Pavement Design, 10 (2009) 25-43.

[4] K.J. Leontaritis, J.O. Amaefule, R.E. Charles, A Systematic Approach for the Prevention and

Treatment of Formation Damage Caused by Asphaltene Deposition.

[5] T. Otowa, R. Tanibata, M. Itoh, Production and adsorption characteristics of MAXSORB:

High-surface-area active carbon, Gas Separation & Purification, 7 (1993) 241-245.

[6] C.C. Small, Z. Hashisho, A.C. Ulrich, Preparation and characterization of activated carbon

from oil sands coke, Fuel, 92 (2012) 69-76.

[7] M.F. Yardim, E. Ekinci, V. Minkova, M. Razvigorova, T. Budinova, N. Petrov, M.

Goranova, Formation of porous structure of semicokes from pyrolysis of Turkish coals in

different atmospheres☆, Fuel, 82 (2003) 459-463.

[8] M.A. Lillo-Ródenas, D. Cazorla-Amorós, A. Linares-Solano, Understanding chemical

reactions between carbons and NaOH and KOH: An insight into the chemical activation

mechanism, Carbon, 41 (2003) 267-275.

[9] M. Wu, Q. Zha, J. Qiu, X. Han, Y. Guo, Z. Li, A. Yuan, X. Sun, Preparation of porous

carbons from petroleum coke by different activation methods, Fuel, 84 (2005) 1992-1997.

[10] A. Karimi, O. Thinon, J. Fournier, J.M. Hill, Activated carbon prepared from Canadian oil

sands coke by CO2 activation: I. Trends in pore development and the effect of pre-oxidation,

The Canadian Journal of Chemical Engineering, 91 (2013) 1491-1499.

[11] C. Chen, H. Huang, Y. Yu, J. Shi, C. He, R. Albilali, H. Pan, Template-free synthesis of

hierarchical porous carbon with controlled morphology for CO2 efficient capture, Chemical

Engineering Journal, 353 (2018) 584-594.

[12] L. Pan, X. Li, Y. Wang, J. Liu, W. Tian, H. Ning, M. Wu, 3D interconnected honeycomb-

like and high rate performance porous carbons from petroleum asphalt for supercapacitors,

Applied Surface Science, 444 (2018) 739-746.

[13] J. Wu, V. Montes, L.D. Virla, J.M. Hill, Impacts of amount of chemical agent and addition

of steam for activation of petroleum coke with KOH or NaOH, Fuel Processing Technology, 181

(2018) 53-60.

[14] H. Predel, Petroleum Coke, in: Ullmann's Encyclopedia of Industrial Chemistry, 2000.

[15] J.M. Hill, A. Karimi, M. Malekshahian, Characterization, gasification, activation, and

potential uses for the millions of tonnes of petroleum coke produced in Canada each year,

Canadian Journal of Chemical Engineering, 92 (2014) 1618-1626.

[16] R.G. Etter, Production and use of a premium fuel grade petroleum coke, in, Google Patents,

2001.

[17] AER, ST98-2017: Alberta’s Energy Reserves 2016 Supply/Demand Outlook, in, Calgary:

Alberta Energy Regulator, 2017.

[18] Z. Xing, K. Du, Particulate matter emissions over the oil sands regions in Alberta, Canada,

Environmental Reviews, 25 (2017) 432-443.

[19] R. Kozicki, G. Wrightson, Prevention of Airborne Dust from Petroleum Coke Stockpiles, in:

S.J. Ikhmayies, B. Li, J.S. Carpenter, J.-Y. Hwang, S.N. Monteiro, J. Li, D. Firrao, M. Zhang, Z.

Page 157: Conversion of Petroleum Coke to Porous Materials

143

Peng, J.P. Escobedo-Diaz, C. Bai (Eds.) Characterization of Minerals, Metals, and Materials

2016, Springer International Publishing, Cham, 2016, pp. 319-325.

[20] J.M. Hill, Sustainable and/or waste sources for catalysts: Porous carbon development and

gasification, Catalysis Today, 285 (2017) 204-210.

[21] N. Rambabu, R. Azargohar, A.K. Dalai, J. Adjaye, Evaluation and comparison of

enrichment efficiency of physical/chemical activations and functionalized activated carbons

derived from fluid petroleum coke for environmental applications, Fuel Processing Technology,

106 (2013) 501-510.

[22] M. Tan, P. Li, J. Zheng, T. Noritatsu, M. Wu, Preparation and modification of high

performance porous carbons from petroleum coke for use as supercapacitor electrodes, New

Carbon Materials, 31 (2016) 343-351.

[23] Y. Zhang, T. Cai, J. Huang, W. Xing, Z. Yan, Functionalized activated carbon prepared

form petroleum coke with high-rate supercapacitive performance, Journal of Materials Research,

31 (2016) 3723-3730.

[24] R. Fong, U. von Sacken, J.R. Dahn, Studies of Lithium Intercalation into Carbons Using

Nonaqueous Electrochemical Cells, Journal of The Electrochemical Society, 137 (1990) 2009-

2013.

[25] G.G. Stavropoulos, A.A. Zabaniotou, Minimizing activated carbons production cost, Fuel

Processing Technology, 90 (2009) 952-957.

[26] H. Al-Haj-Ibrahim, B.I. Morsi, Desulfurization of petroleum coke: a review, Industrial &

Engineering Chemistry Research, 31 (1992) 1835-1840.

[27] J. Shan, J.J. Huang, J.Z. Li, G. Li, J.T. Zhao, Y.T. Fang, Insight into transformation of

sulfur species during KOH activation of high sulfur petroleum coke, Fuel, 215 (2018) 258-265.

[28] Q. Zhong, Q. Mao, J. Xiao, A. van Duin, J.P. Mathews, Sulfur removal from petroleum

coke during high-temperature pyrolysis. Analysis from TG-MS data and ReaxFF simulations,

Journal of Analytical and Applied Pyrolysis, 132 (2018) 134-142.

[29] P.-j. Zhao, C. Ma, J.-t. Wang, W.-m. Qiao, L.-c. Ling, Almost total desulfurization of high-

sulfur petroleum coke by Na2CO3-promoted calcination combined with ultrasonic-assisted

chemical oxidation, New Carbon Materials, 33 (2018) 587-594.

[30] S.H. Lee, C.S. Choi, Chemical activation of high sulfur petroleum cokes by alkali metal

compounds, Fuel Processing Technology, 64 (2000) 141-153.

[31] J. Xiao, Q. Zhong, F. Li, J. Huang, Y. Zhang, B. Wang, Modeling the Change of Green

Coke to Calcined Coke Using Qingdao High-Sulfur Petroleum Coke, Energy & Fuels, 29 (2015)

3345-3352.

[32] R. Alcántara, J.M. Jiménez-Mateos, P. Lavela, J. Morales, J.L. Tirado, Microstructure and

intercalation properties of petrol cokes obtained at 1400°C, Materials Science and Engineering:

B, 39 (1996) 216-223.

[33] I.C. Popovici, S. Birghila, G. Voicu, V. Ionescu, V. Clupina, G. Prodan, Morphological and

microstructural characterization of some petroleum cokes as potential anode materials in lithium

ion batteries, J. Optoelectron. Adv. Mater., 12 (2010) 1903-1908.

[34] Q. Zhong, Q. Mao, L. Zhang, J. Xiang, J. Xiao, J.P. Mathews, Structural features of

Qingdao petroleum coke from HRTEM lattice fringes: Distributions of length, orientation,

stacking, curvature, and a large-scale image-guided 3D atomistic representation, Carbon, 129

(2018) 790-802.

Page 158: Conversion of Petroleum Coke to Porous Materials

144

[35] K. Gharbi, K. Benyounes, M. Khodja, Removal and prevention of asphaltene deposition

during oil production: A literature review, Journal of Petroleum Science and Engineering, 158

(2017) 351-360.

[36] A.T. Mohd Din, B.H. Hameed, A.L. Ahmad, Batch adsorption of phenol onto

physiochemical-activated coconut shell, Journal of Hazardous Materials, 161 (2009) 1522-1529.

[37] H.S. Niasar, H. Li, T.V.R. Kasanneni, M.B. Ray, C. Xu, Surface amination of activated

carbon and petroleum coke for the removal of naphthenic acids and treatment of oil sands

process-affected water (OSPW), Chemical Engineering Journal, 293 (2016) 189-199.

[38] J. Choi, Z.G. Barnard, S. Zhang, J.M. Hill, Ni catalysts supported on activated carbon from

petcoke and their activity for toluene hydrogenation, The Canadian Journal of Chemical

Engineering, 90 (2012) 631-636.

[39] Y. Shi, J. Chen, J. Chen, R.A. Macleod, M. Malac, Preparation and evaluation of

hydrotreating catalysts based on activated carbon derived from oil sand petroleum coke, Applied

Catalysis A: General, 441-442 (2012) 99-107.

[40] S.C. Liew, J.M. Hill, Impacts of vanadium and coke deposits on the CO2 gasification of

nickel catalysts supported on activated carbon from petroleum coke, Applied Catalysis a-

General, 504 (2015) 420-428.

[41] J. Wang, S. Kaskel, KOH activation of carbon-based materials for energy storage, Journal

of Materials Chemistry, 22 (2012) 23710-23725.

[42] S. Roldán, I. Villar, V. Ruíz, C. Blanco, M. Granda, R. Menéndez, R. Santamaría,

Comparison between Electrochemical Capacitors Based on NaOH- and KOH-Activated

Carbons, Energy & Fuels, 24 (2010) 3422-3428.

[43] B. Xu, Y. Chen, G. Wei, G. Cao, H. Zhang, Y. Yang, Activated carbon with high

capacitance prepared by NaOH activation for supercapacitors, Materials Chemistry and Physics,

124 (2010) 504-509.

[44] C. Akmil Başar, A. Karagunduz, B. Keskinler, A. Cakici, Effect of presence of ions on

surface characteristics of surfactant modified powdered activated carbon (PAC), Applied Surface

Science, 218 (2003) 170-175.

[45] Ö. Aktaş, F. Çeçen, Bioregeneration of activated carbon: A review, International

Biodeterioration & Biodegradation, 59 (2007) 257-272.

[46] J.-W. Lim, Y. Choi, H.-S. Yoon, Y.-K. Park, J.-H. Yim, J.-K. Jeon, Extrusion of honeycomb

monoliths employed with activated carbon-LDPE hybrid materials, Journal of Industrial and

Engineering Chemistry, 16 (2010) 51-56.

[47] E. Ayranci, N. Hoda, Adsorption kinetics and isotherms of pesticides onto activated carbon-

cloth, Chemosphere, 60 (2005) 1600-1607.

[48] H.L. Chiang, C.P. Huang, P.C. Chiang, The surface characteristics of activated carbon as

affected by ozone and alkaline treatment, Chemosphere, 47 (2002) 257-265.

[49] A.M. Puziy, O.I. Poddubnaya, R.P. Socha, J. Gurgul, M. Wisniewski, XPS and NMR

studies of phosphoric acid activated carbons, Carbon, 46 (2008) 2113-2123.

[50] R.A. Arnold, J.M. Hill, Effect of calcium and barium on potassium-catalyzed gasification of

ash-free carbon black, Fuel, 254 (2019) 115647.

[51] B. Spitzer, C. Mende, M. Menner, T. Luck, Determination of the carbon content of

biomass—A prerequisite to estimate the complete biodegradation of polymers, Journal of

environmental polymer degradation, 4 (1996) 157-171.

Page 159: Conversion of Petroleum Coke to Porous Materials

145

[52] B. Jiang, Y. Zhang, J. Zhou, K. Zhang, S. Chen, Effects of chemical modification of

petroleum cokes on the properties of the resulting activated carbon, Fuel, 87 (2008) 1844-1848.

[53] S.B. Lyubchik, H. Benaddi, V.V. Shapranov, F. Beguin, Activated carbons from chemically

treated anthracite, Carbon, 35 (1997) 162-165.

[54] C. Moreno-Castilla, J. Rivera-Utrilla, Carbon Materials as Adsorbents for the Removal of

Pollutants from the Aqueous Phase, MRS Bulletin, 26 (2012) 890-894.

[55] T. Budinova, E. Ekinci, F. Yardim, A. Grimm, E. Björnbom, V. Minkova, M. Goranova,

Characterization and application of activated carbon produced by H3PO4 and water vapor

activation, Fuel Processing Technology, 87 (2006) 899-905.

[56] M. Molina-Sabio, F. Rodríguez-Reinoso, F. Caturla, M.J. Sellés, Development of porosity

in combined phosphoric acid-carbon dioxide activation, Carbon, 34 (1996) 457-462.

[57] D. Lozano-Castelló, M.A. Lillo-Ródenas, D. Cazorla-Amorós, A. Linares-Solano,

Preparation of activated carbons from Spanish anthracite: I. Activation by KOH, Carbon, 39

(2001) 741-749.

[58] M.A. Lillo-Ródenas, D. Lozano-Castelló, D. Cazorla-Amorós, A. Linares-Solano,

Preparation of activated carbons from Spanish anthracite: II. Activation by NaOH, Carbon, 39

(2001) 751-759.

[59] M. Krol, G. Gryglewicz, J. Machnikowski, KOH activation of pitch-derived carbonaceous

materials-Effect of carbonization degree, Fuel Processing Technology, 92 (2011) 158-165.

[60] B. Cagnon, X. Py, A. Guillot, F. Stoeckli, G. Chambat, Contributions of hemicellulose,

cellulose and lignin to the mass and the porous properties of chars and steam activated carbons

from various lignocellulosic precursors, Bioresource Technology, 100 (2009) 292-298.

[61] S. Hirunpraditkoon, N. Tunthong, A. Ruangchai, K. Nuithitikul, Adsorption capacities of

activated carbons prepared from bamboo by KOH activation, World Academy of Science,

Engineering and Technology, 78 (2011) 711-715.

[62] V. Minkova, M. Razvigorova, E. Bjornbom, R. Zanzi, T. Budinova, N. Petrov, Effect of

water vapour and biomass nature on the yield and quality of the pyrolysis products from

biomass, Fuel Processing Technology, 70 (2001) 53-61.

[63] M. Olivares-Marín, J.A. Fernández, M.J. Lázaro, C. Fernández-González, A. Macías-

García, V. Gómez-Serrano, F. Stoeckli, T.A. Centeno, Cherry stones as precursor of activated

carbons for supercapacitors, Materials Chemistry and Physics, 114 (2009) 323-327.

[64] W. Chen, H. Zhang, Y. Huang, W. Wang, A fish scale based hierarchical lamellar porous

carbon material obtained using a natural template for high performance electrochemical

capacitors, Journal of Materials Chemistry, 20 (2010) 4773-4775.

[65] W. Huang, H. Zhang, Y. Huang, W. Wang, S. Wei, Hierarchical porous carbon obtained

from animal bone and evaluation in electric double-layer capacitors, Carbon, 49 (2011) 838-843.

[66] X. Li, W. Xing, S. Zhuo, J. Zhou, F. Li, S.-Z. Qiao, G.-Q. Lu, Preparation of capacitor’s

electrode from sunflower seed shell, Bioresource Technology, 102 (2011) 1118-1123.

[67] D. Kalpana, S.H. Cho, S.B. Lee, Y.S. Lee, R. Misra, N.G. Renganathan, Recycled waste

paper—A new source of raw material for electric double-layer capacitors, Journal of Power

Sources, 190 (2009) 587-591.

[68] L. Wei, M. Sevilla, A.B. Fuertes, R. Mokaya, G. Yushin, Hydrothermal Carbonization of

Abundant Renewable Natural Organic Chemicals for High-Performance Supercapacitor

Electrodes, Advanced Energy Materials, 1 (2011) 356-361.

Page 160: Conversion of Petroleum Coke to Porous Materials

146

[69] X. Li, C. Han, X. Chen, C. Shi, Preparation and performance of straw based activated

carbon for supercapacitor in non-aqueous electrolytes, Microporous and Mesoporous Materials,

131 (2010) 303-309.

[70] S. Zhao, C.-Y. Wang, M.-M. Chen, J. Wang, Z.-Q. Shi, Potato starch-based activated

carbon spheres as electrode material for electrochemical capacitor, Journal of Physics and

Chemistry of Solids, 70 (2009) 1256-1260.

[71] M. Mastalerz, L. He, Y.B. Melnichenko, J.A. Rupp, Porosity of Coal and Shale: Insights

from Gas Adsorption and SANS/USANS Techniques, Energy & Fuels, 26 (2012) 5109-5120.

[72] C.D. Monk, G.I. Child, S.A. Nicholson, Biomass, Litter and Leaf Surface Area Estimates of

an Oak-Hickory Forest, Oikos, 21 (1970) 138-141.

[73] R. Pietrzak, XPS study and physico-chemical properties of nitrogen-enriched microporous

activated carbon from high volatile bituminous coal, Fuel, 88 (2009) 1871-1877.

[74] J.I.P. F. Suárez-García, M. Pérez-Mendoza, J. Nauroy, A. Martínez-Alonso, J.M.D. Tascón,

Porosity Development in Carbon Nanofibers by Physical and Chemical Activation, Journal of

Nano Research, 17 (2012) 211-227.

[75] E. Raymundo-Piñero, P. Azaïs, T. Cacciaguerra, D. Cazorla-Amorós, A. Linares-Solano, F.

Béguin, KOH and NaOH activation mechanisms of multiwalled carbon nanotubes with different

structural organisation, Carbon, 43 (2005) 786-795.

[76] R. Xue, Z. Shen, Formation of graphite-potassium intercalation compounds during

activation of MCMB with KOH, Carbon, 41 (2003) 1862-1864.

[77] P. Vanysek, "Electrochemical series" in Handbook of Chemistry and Physics: 92nd Edition,

in, Chemical Rubber Company, 2011.

[78] M. Yuan, Y. Kim, C.Q. Jia, Feasibility of recycling KOH in chemical activation of oil-sands

petroleum coke, The Canadian Journal of Chemical Engineering, 90 (2012) 1472-1478.

[79] J. Dı́az-Terán, D.M. Nevskaia, J.L.G. Fierro, A.J. López-Peinado, A. Jerez, Study of

chemical activation process of a lignocellulosic material with KOH by XPS and XRD,

Microporous and Mesoporous Materials, 60 (2003) 173-181.

[80] P. Ehrburger, A. Addoun, F. Addoun, J.-B. Donnet, Carbonization of coals in the presence

of alkaline hydroxides and carbonates: Formation of activated carbons, Fuel, 65 (1986) 1447-

1449.

[81] C. Lu, S. Xu, C. Liu, The role of K2CO3 during the chemical activation of petroleum coke

with KOH, Journal of Analytical and Applied Pyrolysis, 87 (2010) 282-287.

[82] D. Lozano-Castelló, J.M. Calo, D. Cazorla-Amorós, A. Linares-Solano, Carbon activation

with KOH as explored by temperature programmed techniques, and the effects of hydrogen,

Carbon, 45 (2007) 2529-2536.

[83] A. S.Kim, Fundamentals of Irreversible Thermodynamics for Coupled Transport,

IntechOpen, 2019.

[84] J. Ågren, Thermodynamics and Diffusion Coupling in Alloys—Application-Driven Science,

Metallurgical and Materials Transactions A, 43 (2012) 3453-3461.

[85] D.S. Scott, F.A.L. Dullien, Diffusion of ideal gases in capillaries and porous solids, AIChE

Journal, 8 (1962) 113-117.

[86] J.-G. Choi, D.D. Do, H.D. Do, Surface Diffusion of Adsorbed Molecules in Porous Media: 

Monolayer, Multilayer, and Capillary Condensation Regimes, Industrial & Engineering

Chemistry Research, 40 (2001) 4005-4031.

Page 161: Conversion of Petroleum Coke to Porous Materials

147

[87] V. Vattipalli, X. Qi, P.J. Dauenhauer, W. Fan, Long Walks in Hierarchical Porous Materials

due to Combined Surface and Configurational Diffusion, Chemistry of Materials, 28 (2016)

7852-7863.

[88] A. Linares-Solano, M.A. Lillo-Rodenas, J.P. Marco Lozar, M. Kunowsky, A.J. Romero

Anaya, NaOH and KOH for preparing activated carbons used in energy and environmental

applications, International Journal of Energy, Environment and Economics, 20 (2012) 59-91.

[89] N. Byamba-Ochir, W.G. Shim, M.S. Balathanigaimani, H. Moon, Highly porous activated

carbons prepared from carbon rich Mongolian anthracite by direct NaOH activation, Applied

Surface Science, 379 (2016) 331-337.

[90] R.L. Tseng, Physical and chemical properties and adsorption type of activated carbon

prepared from plum kernels by NaOH activation, Journal of Hazardous Materials, 147 (2007)

1020-1027.

[91] R.L. Tseng, Mesopore control of high surface area NaOH-activated carbon, Journal of

Colloid and Interface Science, 303 (2006) 494-502.

[92] Z. Hu, H. Guo, M.P. Srinivasan, N. Yaming, A simple method for developing mesoporosity

in activated carbon, Separation and Purification Technology, 31 (2003) 47-52.

[93] D.C. Huang, Q.L. Liu, W. Zhang, J.J. Ding, J.J. Gu, S.M. Zhu, Q.X. Guo, D. Zhang,

Preparation of high-surface-area activated carbon from Zizania latifolia leaves by one-step

activation with K2CO3/rarefied air, Journal of Materials Science, 46 (2011) 5064-5070.

[94] V. Montes, J.M. Hill, Pore enlargement of carbonaceous materials by metal oxide catalysts

in the presence of steam: Influence of metal oxide size and porosity of starting material,

Microporous and Mesoporous Materials, 256 (2018) 91-101.

[95] T. Kawano, M. Kubota, M.S. Onyango, F. Watanabe, H. Matsuda, Preparation of activated

carbon from petroleum coke by KOH chemical activation for adsorption heat pump, Applied

Thermal Engineering, 28 (2008) 865-871.

[96] L.D. Virla, V. Montes, J. Wu, S.F. Ketep, J.M. Hill, Synthesis of porous carbon from

petroleum coke using steam, potassium and sodium: Combining treatments to create

mesoporosity, Microporous and Mesoporous Materials, 234 (2016) 239-247.

[97] M.G. Deng, R.Q. Wang, Effect of petroleum coke expanding by perchloric acid on the

performance of the resulted activated carbon, Functional Materials Letters, 07 (2014) 1350066.

[98] X. Yuan, S.W. Choi, E. Jang, K.B. Lee, Chemically activated microporous carbons derived

from petroleum coke: Performance evaluation for CF4 adsorption, Chemical Engineering

Journal, 336 (2018) 297-305.

[99] E. Jang, S.W. Choi, S.-M. Hong, S. Shin, K.B. Lee, Development of a cost-effective CO2

adsorbent from petroleum coke via KOH activation, Applied Surface Science, 429 (2018) 62-71.

[100] T. Mochizuki, M. Kubota, H. Matsuda, L.F. D'Elia Camacho, Adsorption behaviors of

ammonia and hydrogen sulfide on activated carbon prepared from petroleum coke by KOH

chemical activation, Fuel Processing Technology, 144 (2016) 164-169.

[101] H. Zhang, J. Chen, S. Guo, Preparation of natural gas adsorbents from high-sulfur

petroleum coke, Fuel, 87 (2008) 304-311.

[102] H.S. Niasar, H. Li, S. Das, T.V.R. Kasanneni, M.B. Ray, C. Xu, Preparation of activated

petroleum coke for removal of naphthenic acids model compounds: Box-Behnken design

optimization of KOH activation process, Journal of Environmental Management, 211 (2018) 63-

72.

Page 162: Conversion of Petroleum Coke to Porous Materials

148

[103] M. Wu, Y. Wang, D. Wang, M. Tan, P. Li, W. Wu, N. Tsubaki, SO3H-modified petroleum

coke derived porous carbon as an efficient solid acid catalyst for esterification of oleic acid,

Journal of Porous Materials, 23 (2016) 263-271.

[104] P. Pourghahramani, E. Forssberg, Changes in the structure of hematite by extended dry

grinding in relation to imposed stress energy, Powder Technology, 178 (2007) 30-39.

[105] N. Kotake, K. Suzuki, S. Asahi, Y. Kanda, Experimental study on the grinding rate

constant of solid materials in a ball mill, Powder Technology, 122 (2002) 101-108.

[106] H.J. Jung, Y. Sohn, H.G. Sung, H.S. Hyun, W.G. Shin, Physicochemical properties of ball

milled boron particles: Dry vs. wet ball milling process, Powder Technology, 269 (2015) 548-

553.

[107] M.V. Antisari, A. Montone, N. Jovic, E. Piscopiello, C. Alvani, L. Pilloni, Low energy

pure shear milling: A method for the preparation of graphite nano-sheets, Scripta Materialia, 55

(2006) 1047-1050.

[108] R.M. Thankachan, M.M. Rahman, I. Sultana, A.M. Glushenkov, S. Thomas, N. Kalarikkal,

Y. Chen, Enhanced lithium storage in ZnFe2O4–C nanocomposite produced by a low-energy

ball milling, Journal of Power Sources, 282 (2015) 462-470.

[109] T.-T.-N. Nguyen, J.-L. He, Preparation of titanium monoxide nanopowder by low-energy

wet ball-milling, Advanced Powder Technology, 27 (2016) 1868-1873.

[110] S.C. Peterson, M.A. Jackson, S. Kim, D.E. Palmquist, Increasing biochar surface area:

Optimization of ball milling parameters, Powder Technology, 228 (2012) 115-120.

[111] X.-h. Yang, P. He, Y.-y. Xia, Preparation of mesocellular carbon foam and its application

for lithium/oxygen battery, Electrochemistry Communications, 11 (2009) 1127-1130.

[112] H. Imamura, M. Kusuhara, S. Minami, M. Matsumoto, K. Masanari, Y. Sakata, K. Itoh, T.

Fukunaga, Carbon nanocomposites synthesized by high-energy mechanical milling of graphite

and magnesium for hydrogen storage, Acta Materialia, 51 (2003) 6407-6414.

[113] P. Fratzl, R. Weinkamer, Nature’s hierarchical materials, Progress in Materials Science, 52

(2007) 1263-1334.

[114] Y. Li, Z.-Y. Fu, B.-L. Su, Hierarchically Structured Porous Materials for Energy

Conversion and Storage, Advanced Functional Materials, 22 (2012) 4634-4667.

[115] S. Dutta, A. Bhaumik, K.C.W. Wu, Hierarchically porous carbon derived from polymers

and biomass: effect of interconnected pores on energy applications, Energy & Environmental

Science, 7 (2014) 3574-3592.

[116] R.-w. Fu, Z.-h. Li, Y.-r. Liang, F. Li, F. Xu, D.-c. Wu, Hierarchical porous carbons:

design, preparation, and performance in energy storage, New Carbon Materials, 26 (2011) 171-

179.

[117] S. Li, Y. Liang, D. Wu, R. Fu, Fabrication of bimodal mesoporous carbons from petroleum

pitch by a one-step nanocasting method, Carbon, 48 (2010) 839-843.

[118] C. Xue, B. Tu, D. Zhao, Facile fabrication of hierarchically porous carbonaceous

monoliths with ordered mesostructure via an organic organic self-assembly, Nano Research, 2

(2009) 242-253.

[119] M.C. Gutiérrez, M.L. Ferrer, F. del Monte, Ice-Templated Materials: Sophisticated

Structures Exhibiting Enhanced Functionalities Obtained after Unidirectional Freezing and Ice-

Segregation-Induced Self-Assembly, Chemistry of Materials, 20 (2008) 634-648.

Page 163: Conversion of Petroleum Coke to Porous Materials

149

[120] A.-H. Lu, W.-C. Li, W. Schmidt, F. Schüth, Fabrication of hierarchically structured carbon

monoliths via self-binding and salt templating, Microporous and Mesoporous Materials, 95

(2006) 187-192.

[121] S.H. Kim, B.Y.H. Liu, M.R. Zachariah, Ultrahigh surface area nanoporous silica particles

via an aero-sol-gel process, Langmuir, 20 (2004) 2523-2526.

[122] X. Zheng, X. Cao, X. Li, J. Tian, C. Jin, R. Yang, Biomass lysine-derived nitrogen-doped

carbon hollow cubes: Via a NaCl crystal template: An efficient bifunctional electrocatalyst for

oxygen reduction and evolution reactions, Nanoscale, 9 (2017) 1059-1067.

[123] D. Qiu, T. Cao, J. Zhang, S.-W. Zhang, D. Zheng, H. Wu, W. Lv, F. Kang, Q.-H. Yang,

Precise carbon structure control by salt template for high performance sodium-ion storage,

Journal of Energy Chemistry, 31 (2019) 101-106.

[124] D. Lozano-Castelló, D. Cazorla-Amorós, A. Linares-Solano, D.F. Quinn, Activated carbon

monoliths for methane storage: influence of binder, Carbon, 40 (2002) 2817-2825.

[125] J. Alcañiz-Monge, J.P. Marco-Lozar, M.Á. Lillo-Ródenas, CO2 separation by carbon

molecular sieve monoliths prepared from nitrated coal tar pitch, Fuel Processing Technology, 92

(2011) 915-919.

[126] E. Vilaplana-Ortego, M.A. Lillo-Ródenas, J. Alcañiz-Monge, D. Cazorla-Amorós, A.

Linares-Solano, New insights on the direct activation of isotropic petroleum pitch by alkaline

hydroxides, Fuel Processing Technology, 91 (2010) 145-149.

[127] J. Li, S. Wang, Y. Ren, Z. Ren, Y. Qiu, J. Yu, Nitrogen-doped activated carbon with

micrometer-scale channels derived from luffa sponge fibers as electrocatalysts for oxygen

reduction reaction with high stability in acidic media, Electrochimica Acta, 149 (2014) 56-64.

[128] S. Wei, H. Zhang, Y. Huang, W. Wang, Y. Xia, Z. Yu, Pig bone derived hierarchical

porous carbon and its enhanced cycling performance of lithium-sulfur batteries, Energy &

Environmental Science, 4 (2011) 736-740.

[129] Y. Liang, D. Wu, R. Fu, Carbon Microfibers with Hierarchical Porous Structure from

Electrospun Fiber-Like Natural Biopolymer, Scientific Reports, 3 (2013) 1119.

[130] F. Zhang, K.-X. Wang, G.-D. Li, J.-S. Chen, Hierarchical porous carbon derived from rice

straw for lithium ion batteries with high-rate performance, Electrochemistry Communications, 11

(2009) 130-133.

[131] J. Zhou, Z. Qiu, J. Zhou, W. Si, H. Cui, S. Zhuo, Hierarchical porous carbons from

alkaline poplar bark extractive-based phenolic resins for supercapacitors, Electrochimica Acta,

180 (2015) 1007-1013.

[132] Z. Zheng, Q. Gao, Hierarchical porous carbons prepared by an easy one-step carbonization

and activation of phenol–formaldehyde resins with high performance for supercapacitors,

Journal of Power Sources, 196 (2011) 1615-1619.

[133] B. Xu, S. Hou, G. Cao, M. Chu, Y. Yang, Easy synthesis of a high surface area,

hierarchical porous carbon for high-performance supercapacitors, RSC Advances, 3 (2013)

17500-17506.

[134] H. Zhang, L. Zhang, J. Chen, H. Su, F. Liu, W. Yang, One-step synthesis of hierarchically

porous carbons for high-performance electric double layer supercapacitors, Journal of Power

Sources, 315 (2016) 120-126.

[135] Y. Gong, Z. Wei, J. Wang, P. Zhang, H. Li, Y. Wang, Design and Fabrication of

Hierarchically Porous Carbon with a Template-free Method, Scientific Reports, 4 (2014) 6349.

Page 164: Conversion of Petroleum Coke to Porous Materials

150

[136] H. Zhong, F. Xu, Z. Li, R. Fu, D. Wu, High-energy supercapacitors based on hierarchical

porous carbon with an ultrahigh ion-accessible surface area in ionic liquid electrolytes,

Nanoscale, 5 (2013) 4678-4682.

[137] C. Zou, D. Wu, M. Li, Q. Zeng, F. Xu, Z. Huang, R. Fu, Template-free fabrication of

hierarchical porous carbon by constructing carbonyl crosslinking bridges between polystyrene

chains, Journal of Materials Chemistry, 20 (2010) 731-735.

[138] V. Kurian, N. Mahapatra, B. Wang, M. Alipour, F. Martens, R. Gupta, Analysis of Soot

Formed during the Pyrolysis of Athabasca Oil Sand Asphaltenes, Energy & Fuels, 29 (2015)

6823-6831.

[139] S.C. Liew, Impacts of Vanaddium and Coke Deposition on CO2 Gasification of Nickel

Catalysts Supported on Activated Carbon from Petroleum Coke, in, Univeristy of Calgary,

Calgary, AB, Canada, 2014.

[140] R.G. Kaldenhoven, J.M. Hill, Determining the pore structure of activated carbon by

nitrogen gas adsorption, in: Catalysis: Volume 30, The Royal Society of Chemistry, 2018, pp.

41-63.

[141] D. Lozano-Castelló, D. Cazorla-Amorós, A. Linares-Solano, Usefulness of CO2

adsorption at 273 K for the characterization of porous carbons, Carbon, 42 (2004) 1233-1242.

[142] D. Cazorla-Amorós, J. Alcañiz-Monge, A. Linares-Solano, Characterization of Activated

Carbon Fibers by CO2 Adsorption, Langmuir, 12 (1996) 2820-2824.

[143] A. Guillot, F. Stoeckli, Reference isotherm for high pressure adsorption of CO2 by carbons

at 273 K, Carbon, 39 (2001) 2059-2064.

[144] S. Lowell, J.E. Shields, M.A. Thomas, M. Thommes, Characterization of Porous Solids

and Powders: Surface Area, Pore Size and Density, Springer Netherlands, 2012.

[145] P.P.R. David M. Levine, Robert K. Smidt, Applied Statistics for Engineers and Scientists,

Prentice Hall, Inc, 2001.

[146] K.S.W. Sing, Reporting physisorption data for gas/solid systems with special reference to

the determination of surface area and porosity (Recommendations 1984), in: Pure and Applied

Chemistry, 1985, pp. 603.

[147] L. Ding, Z. Dai, J. Wei, Z. Zhou, G. Yu, Catalytic effects of alkali carbonates on coal char

gasification, Journal of the Energy Institute, 90 (2017) 588-601.

[148] P. Parthasarathy, K.S. Narayanan, S. Ceylan, N.A. Pambudi, Optimization of Parameters

for the Generation of Hydrogen in Combined Slow Pyrolysis and Steam Gasification of Biomass,

Energy & Fuels, 31 (2017) 13692-13704.

[149] N. Sadhwani, S. Adhikari, M.R. Eden, Z. Wang, R. Baker, Southern pines char gasification

with CO2—Kinetics and effect of alkali and alkaline earth metals, Fuel Processing Technology,

150 (2016) 64-70.

[150] A. Karimi, N. Semagina, M.R. Gray, Kinetics of catalytic steam gasification of bitumen

coke, Fuel, 90 (2011) 1285-1291.

[151] T. Wigmans, R. Elfring, J.A. Moulijn, On the mechanism of the potassium carbonate

catalysed gasification of activated carbon: the influence of the catalyst concentration on the

reactivity and selectivity at low steam pressures, Carbon, 21 (1983) 1-12.

[152] K.S.W. Sing, D.H. Everett, R.A.W. Haul, L. Moscou, R.A. Pierotti, J. Rouquerol, T.

Siemieniewska, Reporting Physisorption Data for Gas/Solid Systems, in: Handbook of

Heterogeneous Catalysis, Wiley-VCH Verlag GmbH & Co. KGaA, 2008.

Page 165: Conversion of Petroleum Coke to Porous Materials

151

[153] E.M. Levin, C.R. Robbins, H.F. McMurdie, S. American Ceramic, Phase diagrams for

ceramists, American Ceramic Society, Columbus, Ohio, 1964.

[154] J. Kopyscinski, M. Rahman, R. Gupta, C.A. Mims, J.M. Hill, K2CO3 catalyzed CO2

gasification of ash-free coal. Interactions of the catalyst with carbon in N2 and CO2 atmosphere,

Fuel, 117, Part B (2014) 1181-1189.

[155] D.W. McKee, Gasification of graphinte in carbon dioxide and water vapor-the catalytic

effects of alkali metal salts, Carbon, 20 (1982) 59-66.

[156] U.S.D.o. Energy, How to calculate the true cost of steam, in, Industrial Technoloties

Program, Washington. DC, 2003.

[157] D.W. McKee, D. Chatterji, The catalytic behavior of alkali metal carbonates and oxides in

graphite oxidation reactions, Carbon, 13 (1975) 381-390.

[158] O.J. Foust, SODIUM-NaK ENGINEERING HANDBOOK. VOLUME I. SODIUM

CHEMISTRY AND PHYSICAL PROPERTIES, 1972, United States, 1972.

[159] I. Barin, F. Sauert, Thermochemical Data of Pure Substances, VCH, 1993.

[160] E. Partlan, K. Davis, Y. Ren, O.G. Apul, O.T. Mefford, T. Karanfil, D.A. Ladner, Effect of

bead milling on chemical and physical characteristics of activated carbons pulverized to

superfine sizes, Water Research, 89 (2016) 161-170.

[161] J. Zou, B. Yang, K. Gong, S. Wu, Z. Zhou, F. Wang, Z. Yu, Effect of mechanochemical

treatment on petroleum coke–CO2 gasification, Fuel, 87 (2008) 622-627.

[162] R. Nandhini, P.A. Mini, B. Avinash, S.V. Nair, K.R.V. Subramanian, Supercapacitor

electrodes using nanoscale activated carbon from graphite by ball milling, Materials Letters, 87

(2012) 165-168.

[163] M.S. Dresselhaus, A. Jorio, A.G.S. Filho, R. Saito, Defect characterization in graphene and

carbon nanotubes using Raman spectroscopy, Philosophical Transactions of the Royal Society A:

Mathematical, Physical and Engineering Sciences, 368 (2010) 5355-5377.

[164] P. Mallet-Ladeira, P. Puech, C. Toulouse, M. Cazayous, N. Ratel-Ramond, P. Weisbecker,

G.L. Vignoles, M. Monthioux, A Raman study to obtain crystallite size of carbon materials: A

better alternative to the Tuinstra–Koenig law, Carbon, 80 (2014) 629-639.

[165] J. Song, W. Shen, J. Wang, W. Fan, Superior carbon-based CO2 adsorbents prepared from

poplar anthers, Carbon, 69 (2014) 255-263.

[166] E. Jang, S.W. Choi, K.B. Lee, Effect of carbonization temperature on the physical

properties and CO2 adsorption behavior of petroleum coke-derived porous carbon, Fuel, 248

(2019) 85-92.

[167] A.G.V. Berry, R. Edgeworth-Johnstone, Petroleum Coke.Formation and Properties,

Industrial & Engineering Chemistry, 36 (1944) 1140-1144.

[168] R. Xiao, S. Xu, Q. Li, Y. Su, The effects of hydrogen on KOH activation of petroleum

coke, Journal of Analytical and Applied Pyrolysis, 96 (2012) 120-125.

[169] M.R. Gray, G. Assenheimer, L. Boddez, W.C. McCaffrey, Melting and Fluid Behavior of

Asphaltene Films at 200−500 °C, Energy & Fuels, 18 (2004) 1419-1423.

[170] M. Yi, Z. Shen, A review on mechanical exfoliation for the scalable production of

graphene, Journal of Materials Chemistry A, 3 (2015) 11700-11715.

[171] D. Wildenschild, A.P. Sheppard, X-ray imaging and analysis techniques for quantifying

pore-scale structure and processes in subsurface porous medium systems, Advances in Water

Resources, 51 (2013) 217-246.

Page 166: Conversion of Petroleum Coke to Porous Materials

152

[172] F. Mees, R. Swennen, M. Van Geet, P. Jacobs, Applications of X-ray computed tomograpy

in the geosciences, in: Geological Society Special Publication, 2003, pp. 1-6.

[173] S. Mazumder, K.H.A.A. Wolf, K. Elewaut, R. Ephraim, Application of X-ray computed

tomography for analyzing cleat spacing and cleat aperture in coal samples, International Journal

of Coal Geology, 68 (2006) 205-222.

[174] M.N. Heriawan, K. Koike, Coal quality related to microfractures identified by CT image

analysis, International Journal of Coal Geology, 140 (2015) 97-110.

[175] H.L. Ramandi, P. Mostaghimi, R.T. Armstrong, M. Saadatfar, W.V. Pinczewski, Porosity

and permeability characterization of coal: a micro-computed tomography study, International

Journal of Coal Geology, 154-155 (2016) 57-68.

Page 167: Conversion of Petroleum Coke to Porous Materials

153

Appendix A: Mixing efficiency on activated carbon from petroleum coke

Table A-1. Element composition of petcoke and ACs by SEM/EDX and ultimate analysis

Sample Carbon (wt%) Oxygen (wt%) Sulfur (wt%)

Petcoke

Petcoke cross-section from ES 86.60 + 3.77 6.96 + 3.81 6.44 + 2.11

Petcoke cross-section from

common lab 86.43 + 2.11 9.13 + 1.29 4.42 + 0.96

Petcoke exterior from common lab 85.02 + 3.89 6.61 + 1.18 8.37 + 4.48

Ultimate analysis 83.9 1.4 6.4

ACs

0.43K_500 cross-section from

common lab 74.41+ 8.86 23.89 + 7.45 1.70 + 2.39

0.43K/St_500 cross-section from

common lab 78.23+ 4.43 20.71+ 3.26 1.08+ 1.17

The ultimate analysis not only detect carbon, oxygen, sulfur content of petcoke, but

hydrogen (3.6 wt%), nitrogen (2.0 wt%), and ash (2.6%) was included into 100% of the

material composition.

The oxygen content was indirectly calculated in ultimate analysis by the difference:

O = 100 − (C + H + N + S)

I only consider carbon, oxygen, sulfur content for SEM/EDX analysis, since the other

elements (such as hydrogen, nitrogen) peaks were not obvious to detect on the spectra.

Since the analysis process is different with ultimate analysis and SEM/EDX, the data was

not equally comparable within this two measurement.

Considering the variation in sulfur and oxygen content by SEM/EDX, other method to

measure the elemental analysis need to be found.

Elemental analysis in Chemistry department or ICP analysis in common lab could be

considered.

Considering the variation in each element content, the composition was similar for AC

with or without steam addition.

Page 168: Conversion of Petroleum Coke to Porous Materials

154

Table A-2. Surface area and pore volume of raw petcoke by CO2 adsorption

Petcoke particle size

(µm)

P/P0 SA by DA method

(m2/g)

Pore volume by DA method

(cm3/g)

300-600 0-0.035 38 0.014

150-300 (Previous) 0-0.3 20 0.006

150-300 (Corrected) 0-0.035 84 0.08

<150 (Melanie’s

data)

0-0.03 29 0.02

Same relative pressure and Dubinin-Astakhov (DA) method were used to compared with

data in Arash’s paper.

My data after correction was following the tread with Arash’s data.

I also double checked the raw file of Melanie’s data. Since her data was only collected

from 0-0.03 P/P0, she had a lower value of surface area and pore volume for petcoke

particle smaller than 150 µm.

Page 169: Conversion of Petroleum Coke to Porous Materials

155

Appendix B: Preparation of activated carbon from petcoke

Figure B-1. Pore size distribution (PSD) and cumulative pore volume on NaOH ( for PSD,

solid line for cumulative pore volume) and NaOH/St ( for PSD, dash line for cumulative pore

volume) activated samples at different NaOH to petcoke molar ratio (a.1.06, b. 0.71, c. 0.35)

NaOH activated samples are micro-mesoporous materials.

Pore Width (nm)

0 5 10 15 20 25 30

dV

/dlo

g(w

) P

ore

Vo

lum

e (

cm

3/g

)

0.0

0.2

0.4

0.6

0.8 Cu

mu

lativ

e P

ore

Vo

lum

e (c

m3/g

)

0.0

0.2

0.4

0.6

0.8

dV

/dlo

g(w

) P

ore

Vo

lum

e (

cm

3/g

)

0.0

0.2

0.4

0.6

0.8 Cu

mu

lativ

e P

ore

Vo

lum

e (c

m3/g

)

0.0

0.2

0.4

0.6

0.8

dV

/dlo

g(w

) P

ore

Vo

lum

e (

cm

3/g

)

0.0

0.2

0.4

0.6

0.8 Cu

mu

lativ

e P

ore

Vo

lum

e (c

m3/g

)

0.0

0.2

0.4

0.6

0.8a

b

c

Page 170: Conversion of Petroleum Coke to Porous Materials

156

Larger mesopores appears at around 4 nm and 20 nm.

The impact of steam addition on NaOH activated samples are similar.

Steam destroyed almost half of the micropore volume.

The lines of cumulative pore volume are parallel at the same NaOH ratio regardless of

the steam addition. Pores can not be enlarged during the process of adding steam.

.

Page 171: Conversion of Petroleum Coke to Porous Materials

157

Appendix C: Low temperature activation

Table C-1. Elemental analysis

Samples C (%) H (%) N (%) Other

0.43KOH_400 78.54 3.38 1.67 16.42

0.43KOH_450 79.59 3.14 1.74 15.53

0.43KOH_500 72.05 2.56 1.76 23.64

0.43KOH_600 77.60 2.09 1.29 19.03

0.43KOH_700 80.95 1.19 0.91 16.96

0.43KOH_800 90.53 0.58 0.83 8.06

The composition of carbon, hydrogen and nitrogen is similar for AC activated at lower

temperature from 400-600.

An increasing of carbon content was observed with slight decreasing content of hydrogen

and nitrogen.

Page 172: Conversion of Petroleum Coke to Porous Materials

158

Appendix D: Carbon foam by emulsion template

Figure D-1. The emulsion template to produce carbon foam

The carbon structure collapse after drying in the vacuum over

Page 173: Conversion of Petroleum Coke to Porous Materials

159

Appendix E. Reprint Permission Letters

Permission for Figure 2-1

Page 174: Conversion of Petroleum Coke to Porous Materials

160

Permission for Figure 2-2

Page 175: Conversion of Petroleum Coke to Porous Materials

161

Permission for Figure 2-3

Page 176: Conversion of Petroleum Coke to Porous Materials

162

Permission for Figure 2-4

Page 177: Conversion of Petroleum Coke to Porous Materials

163

Permission for Figure 2-5

Page 178: Conversion of Petroleum Coke to Porous Materials

164

Page 179: Conversion of Petroleum Coke to Porous Materials

165

Permission for Figure 2-6

Page 180: Conversion of Petroleum Coke to Porous Materials

166

Permission for Figure 2-7

Page 181: Conversion of Petroleum Coke to Porous Materials

167

Permission for Figure 2-8

Page 182: Conversion of Petroleum Coke to Porous Materials

168

Permission for Figure 2-9

Page 183: Conversion of Petroleum Coke to Porous Materials

169

Permission for Chapter 4

Page 184: Conversion of Petroleum Coke to Porous Materials

170

Permission letter from co-author

Library

University of Calgary

2500 University Dr. NW, Calgary, AB, T2N 3Y7

Canada

March 05, 2019

RE: Copyright permission

Calgary, Alberta, Canada

I, Vicente Montes Jiménez, hereby provide the permission to Jingfeng Wu (PhD Candidate,

Dept. of Chemical & Petroleum Engineering, University of Calgary) to include the journal paper

'Impacts of amount of chemical agent and addition of steam for activation of petroleum coke

with KOH or NaOH' in her PhD thesis. The paper is published in Fuel Processing Technology

(2018, vol. 181, pp. 53-60). I am the second author of this paper and performed SEM, organized

the data to discuss the mechanism of co-activation. I am happy to authorize her to include this

article in her PhD thesis.

Sincerely

Vicente Montes Jimenez

Córdoba, Spain, 2019/03/05

Page 185: Conversion of Petroleum Coke to Porous Materials

171

Permission letter from co-author

Library

University of Calgary

2500 University Dr. NW, Calgary, AB, T2N 3Y7

Canada

January 25, 2019

RE: Copyright permission

I, Luis Daniel Virla Alvarado, hereby provide the permission to Jingfeng Wu (PhD Candidate, Dept.

of Chemical & Petroleum Engineering, University of Calgary) to include the journal paper ‘Impacts

of amount of chemical agent and addition of steam for activation of petroleum coke with KOH or

NaOH’ in her PhD thesis. The paper is published in Fuel Processing Technology (2018, vol. 181, pp.

53-60). I am the third author of this paper and contributed with the discussion of the data analysis and

the explanation of the mechanism of co-activation of petcoke. I am happy to authorize her to include

this article in her PhD thesis.

Sincerely,

Luis Virla, PhD

Post Doctoral Fellow R&D

Industrial Climate Solutions, Inc.


Recommended