adsorption of single-ring model naphthenic acid from oil ... · intra-particle diffusion rate...

135
Adsorption of Single-ring Model Naphthenic Acid from Oil Sands Tailings Pond Water Using Petroleum Coke- Derived Activated Carbon by Bithun Sarkar A thesis submitted in conformity with the requirements for the degree of Master of Applied Science Department of Chemical Engineering and Applied Chemistry University of Toronto © Copyright by Bithun Sarkar 2013

Upload: haxuyen

Post on 04-Jun-2018

219 views

Category:

Documents


0 download

TRANSCRIPT

Adsorption of Single-ring Model Naphthenic Acid from Oil Sands Tailings Pond Water Using Petroleum Coke-

Derived Activated Carbon

by

Bithun Sarkar

A thesis submitted in conformity with the requirements for the degree of Master of Applied Science

Department of Chemical Engineering and Applied Chemistry University of Toronto

© Copyright by Bithun Sarkar 2013

ii

Adsorption of Single-ring Model Naphthenic Acid from Oil Sands

Tailings Pond Water Using Petroleum Coke-Derived Activated

Carbon

Bithun Sarkar

Master of Applied Science

Department of Chemical Engineering and Applied Chemistry

University of Toronto

2013

ABSTRACT

Petroleum coke-derived activated carbons were prepared and used for the adsorptive removal of

a single-ring naphthenic acid (NA) from synthetic oil sands tailings pond water (TPW). The

overall adsorption process was found to be intra-particle diffusion-controlled. The Weber-Morris

intra-particle diffusion rate constants decreased from 7.43 to 1.23 mg/g min0.5

after activated

carbon was post-oxidized with oxygen, suggesting a hindering effect of oxygen surface groups.

The Freundlich model fit of the equilibrium adsorption isotherms and the small negative ΔHo

pointed to a physisorption-dominated process and the importance of specific surface area. It was

estimated that about 2.7 g/L of basic CO2-activated carbon is needed to reduce NA concentration

from 120 mg/L to 2.5 mg/L (~98% removal) in synthetic TPW. However, equilibrium adsorption

capacity was found to vary significantly after oxygen or nitrogen groups were introduced onto

the surface. Therefore, there is a potential for enhanced adsorption by chemical functionalization

of carbon.

iii

Acknowledgements

I would like to express my deepest gratitude to my thesis supervisor Professor Charles Q. Jia for

providing consistent guidance, support, supervision and financial support throughout my

research. I have thoroughly enjoyed the challenging research environment fostered at the Green

Technology Group.

I would also like extend my sincere appreciation to the rest of my committee members,

Professors Don W. Kirk and Edgar Acosta, for their valuable suggestions and feedbacks to my

research.

Professor Shitang Tong has been one of the most resourceful individuals that I have met during

my graduate studies. I would like thank him for all his advices and for performing the post-

treatment of activated carbon with gaseous ammonia. A special thanks to Dr. Eric Morris, who

has been a mentor when I first started at the Green Tech. lab. I would like to express my most

sincere thanks to Siyu (Jois) Xie, who had been an outstanding summer research assistant, and

made many contributions to this work.

Special thanks to John Caguiat, Derek Huynh and Jocelyn Zuliani for their contribution to this

research project; Rose Balazs and Dan Mathers of the Analest Facility (U of T) for sharing their

expertise on FTIR spectroscopy. I would like to thank the rest of the Green Technology Group

for their companionship over the last twenty months.

Finally, I would like to dedicate this work to an old friend of mine.

iv

TABLE OF CONTENTS

CHAPTER 1 INTRODUCTION………………….…………………………..………1

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

1.1.1 The oil sands industry……………………………………...…………....1

1.1.2 Oil sand naphthenic acids……………………………………………….3

1.1.3 Adsorption of naphthenic acids by activated carbon………………........4

1.2 Research Objectives…………………...………..……………………….…......6

CHAPTER 2 OIL SANDS NAPHTHENIC ACIDS AND ACTIVATED

CARBON: A LITERATURE REVIEW……………..…….................8

2.1 Oil Sands Naphthenic Acids…………………………………………..….........8

2.1.1 Origin………………………………………………………………........8

2.1.2 Structure and classification………………………………………….......8

2.1.3 Mobility and toxicity………………………………………………......10

2.1.4 Chemical properties………………………………………………........11

2.1.5 Analytical techniques for detection and quantification………………..12

2.1.6 Remediation techniques………………………………………………..14

2.2 Activated Carbon……..…………………………………………………........15

2.2.1 Applications as an adsorbent…………………………………………..15

2.2.2 Precursor materials for activated carbon production……………..........17

2.2.2.1 Oil sands petroleum coke………………………………............18

2.2.3 Activated carbon production processes…………………………….......19

2.2.3.1 Physical activation……………………………………………..20

2.2.3.2 Chemical activation………………………………………........21

2.2.3.3 Post-modification of activated carbon……………………........22

2.2.3.4 Factors influencing the physical and chemical properties

of activated carbon………………………………………......…23

CHAPTER 3 THEORETICAL OVERVIEW OF ADSORPTION……………….25

3.1 Physisorption vs. Chemisorption.……………………………..………….......25

3.2 Adsorption Kinetics……………………………………………………….......26

3.2.1 Empirical models……………………………………………………....26

3.2.1.1 Pseudo-first order model……………………………………….26

3.2.1.2 Pseudo-second order model……………………………………27

3.2.1.3 Elovich model………………………………………………….28

3.2.2 Mechanistic models…………………………………………………….28

3.2.2.1 Weber-Morris intraparticle diffusion model…………………...30

3.3 Equilibrium Adsorption Isotherms………………………………………….30

v

3.3.1 Langmuir model………………………………………………………..31

3.3.2 Freundlich model……………………………………………………….32

3.3.3 Temkin model…………………………………………………….……33

CHAPTER 4 EXPERIMENTAL METHODOLOGY…......………………...…….34

4.1 Preparation of Activated Carbon……………………………........................34

4.1.1 Materials and instruments……………………………………………...35

4.1.2 Physical activation of delayed petroleum coke………………………...36

4.1.2.1 CO2 activation………………………………………………….36

4.1.2.2 Steam activation……………………………………………......37

4.1.3 Post-treatment of petroleum coke-derived activated carbon…………..39

4.1.3.1 Post-oxidation……………………………………………….....39

4.1.3.2 Post-treatment with gaseous ammonia………………………....41

4.2 Characterization of Activated Carbon………………………........................42

4.2.1 Materials and instruments……………………………………………...42

4.2.2 Pore structure analysis………………………………………………….43

4.2.3 Surface chemistry analysis. ……………………………………….…...44

4.2.3.1 FTIR analysis…………………………………………………...44

4.2.3.2 Boehm titration………………………………………………....44

4.2.3.3 Elemental analysis……………………………………………...46

4.3 Adsorption Study……………………………………………….......................47

4.3.1 Materials and instruments………………………………………….…..47

4.3.2 Model naphthenic acid...…………………………………………….…48

4.3.3 Quantification of the model naphthenic acid and total acid-extractable

organics from aqueous phase…………………………………………..48

4.3.3.1 Sample extraction from aqueous phase……………………..….48

4.3.3.2 FTIR-based quantification…………………………………..….49

4.3.4 Batch-type adsorption kinetics study……………………………..……51

4.3.5 Equilibrium adsorption (isotherm) study…………………………..…..52

4.3.6 Adsorption data quality…….……………………………………..……54

CHAPTER 5 RESULTS AND DISCUSSIONS……..…...………………...….……56

5.1 Characterization of Activated Carbon…………………………………...….56 5.1.1 Pore structure analysis……………………………………………….…56

5.1.2 Surface chemistry analysis………………………………………….….58

5.1.2.1 FTIR analysis……………………………………………….….58

5.1.2.2 Boehm titration……………………………………….………..61

5.1.2.3 Elemental analysis……………………………………….……..62

5.2 Adsorption Study…………………………………………………………...…65 5.2.1 Adsorption kinetics…………………………….……………………….65

5.2.2.1 Effect of pore structure………….…….…………………...…...67

5.2.2.2 Effect of initial adsorbate concentration…………………….….72

vi

5.2.2.3 Effect of temperature…………………………………………..75

5.2.2.4 Effect of surface functional groups……………………………77

5.3.2 Equilibrium adsorption study……………………………………….…82

5.3.2.1 Equilibrium adsorption isotherms……………………………...82

5.3.2.2 Effect of surface functional groups………………………….…85

5.3.2.3 Thermodynamic analysis of the model naphthenic acid

adsorption…………………………………………………..….88

5.3.2.4 Equilibrium adsorption of total acid-extractable organics

From real tailings pond water………….……………………....92

CHAPTER 6 CONCLUSIONS AND RECOMMENDATIONS……………….…95

6.1 Conclusions………………………………………………………….….......…95

6.2 Recommendations………………….……………………………….…….......99

REFERENCES………………………………………………………….……………….…101

APPENDICES ………………………………………………………….………………..…111

vii

LIST OF FIGURES

Figure 2-1: Examples of acyclic, monocyclic, and bicyclic naphthenic acid structure......…9

Figure 3-1: Schematic representation of a three-step conceptual model for solid-liquid

adsorption process……………………………………………………………..29

Figure 4-1: Schematic diagram of the experimental apparatus used for CO2-activation….37

Figure 4-2: Schematic diagram of the experimental apparatus used for steam

activation......…………………………………………………………………..38

Figure 4-3: Schematic diagram of the post-oxidation process with humidified-oxygen…..40

Figure 4-4: Experimental flow diagram of the adsorption study………………………….53

Figure 5-1: FTIR spectra of petroleum coke-derived activated carbon adsorbents….…….59

Figure 5-2: FTIR spectrum of ammonia post-treated CO2-activated carbon………………60

Figure 5-3: Amount of surface oxygen groups on activation carbon adsorbents….……….61

Figure 5-4: Elemental ‘organic’ oxygen content of activated carbon adsorbents…..……...63

Figure 5-5: The effect of shaking speed on the model naphthenic acid adsorption………..66

Figure 5-6: Time-dependent adsorption of mNA on adsorbents of different porosity…….67

Figure 5-7: Correlation between mesoporous volume fraction and pseudo-2nd

order rate

constant…………………………………………………………………….….70

Figure 5-8: Molecular dimensions of the model naphthenic acid………………………….71

Figure 5-9: Time-dependent adsorption of mNA onto CO2-ADC-6h with different

initial concentrations………………………………………….………….……72

Figure 5-10: Initial rate of adsorption as a function of initial mNA concentration…………73

Figure 5-11: Effect of temperature on time-dependent mNA adsorption…………..………75

Figure 5-12: Arrhenius plot of mNA adsorption……………………………………………76

Figure 5-13: Time-dependent adsorption of mNA onto adsorbent of different surface

functionalities…………………………………………………………….……77

Figure 5-14: Weber-Morris intraparticle diffusion plots of mNA adsorption onto surface

oxidized activated carbons…………………………………….………………78

Figure 5-15: Model naphthenic acid adsorption isotherms…………………………………83

viii

Figure 5-16: Adsorption isotherms reflecting the effect of adsorbent surface

functionalities…………………………………..……………………………...87

Figure 5-17: Plot of ln (Kc) versus 1/T for the estimation of thermodynamic

parameters………………………………………………………………...…...91

Figure 5-18: Adsorption isotherms of TAOs and mNA onto CO2-ADC-9h-pNH3………..93

ix

LIST OF TABLES

Table 3-1: Linearized forms of adsorption isotherm model………………………………33

Table 4-1: Adsorbent nomenclature……….........................................................................34

Table 4-2: Materials used in activated carbon production………......................................35

Table 4-3: Instruments and equipment used in activated carbon production………..........35

Table 4-4: Treatment conditions for physical activations..……….....................................39

Table 4-5: Experimental conditions for the post-treatments of activated carbon...............41

Table 4-6: Materials used for activated carbon characterization….....................................42

Table 4-7: Instruments and equipment used for activated carbon characterization............42

Table 4-8: Materials used for the adsorption study…………...….....................................47

Table 4-9: Instruments and equipment used for the adsorption study................................47

Table 4-10: Adsorption kinetic data quality for ‘effect of surface functional groups’

study…………………………………………………………………………...54

Table 4-11: Equilibrium adsorption data quality……………………...................................55

Table 5-1: Physical pore properties of activated carbon adsorbents...................................56

Table 5-2: Amount of surface oxygen groups on activated carbon adsorbents...................62

Table 5-3: Elemental analysis of activated carbons.............................................................64

Table 5-4: Empirical kinetic model parameters from the ‘effect of pore structure’

study...................................................................................................................69

Table 5-5: Correlation between adsorption pore properties and pseudo-2nd

order rate

parameters...........................................................................................................70

Table 5-6: Effect of initial mNA concentration on initial rate of adsorption……..............73

Table 5-7: Effect of temperature on pseudo-2nd

order rate constant...................................76

Table 5-8: Weber-Morris intraparticle diffusion parameters………………..……………79

x

Table 5-9: Correlation between adsorbent surface chemistry and Weber-Morris rate

parameters……………………………………………………………………..80

Table 5-10: Adsorption isotherm model parameters for mNA adsorption…………………84

Table 5-11: Relationship between mesoporous SSA and equilibrium adsorbed quantity… 85

Table 5-12: Effect of surface functional groups on the equilibrium adsorption of mNA…..86

Table 5-13: Effect of temperature on equilibrium constant for mNA adsorption………….89

Table 5-14: Thermodynamic parameters for the mNA adsorption onto activated carbon…91

Table 5-15: Equilibrium adsorption of TAOs and mNA onto CO2-ADC-9h-pNH3….…...93

xi

APPENDICES

Appendix A: Naphthenic acid quantification data…………………………………………..111

Appendix B: Nitrogen isotherms for adsorbent pore structure analysis…….……………….114

Appendix C: Adsorption kinetics data…………………………………….………………...115

Appendix D: Adsorption isotherm data…………………………………….……………….121

1

CHAPTER 1 INTRODUCTION

1.1 Background and Motivation

1.1.1 The Oil Sands Industry

The oil sands deposits of Northern Alberta, Canada are estimated to constitute world’s largest

bitumen reserve, containing approximately 1.7 trillion barrels of bitumen [28]. These deposits

are primarily composed of silica sand, clay minerals, water and bitumen. Bitumen and the

mineral solids have composition ranges of 6-14 wt% and 80-85 wt%, respectively, with water

making up the rest of the balance [85]. Bitumen is a viscous form of petroleum that can be

extracted from the oil sands ore and then chemically processed into synthetic crude oil (SCO).

The ore also contains a wide range of organic compounds with broad molecular weight range

[85]. These compounds are mostly hydrocarbons with different molecular structures (i.e.,

paraffinic, olefinic and aromatic) and organic acids (i.e., carboxylic and sulphonic acid) [74].

Raw bitumen can be recovered from the ore by either in-situ techniques that utilize steam

treatment or by a combination of surface mining and extraction methods. In the past, the latter

has been widely employed by companies such as Suncor Energy and Syncrude Canada Ltd.

Three successive steps may be used to describe the production of synthetic crude oil from

bitumen using this method: (1) mining of the bituminous ore, (2) extraction of bitumen from the

mineral solids, and (3) upgrading of bitumen into lighter fractions to produce SCO. The synthetic

crude oil is then distributed to oil refineries for further purification [85]. Although this method is

very efficient in separating bitumen from the ore, it produces large volumes of overburden

materials also known as tailings [85].

2

The Canadian oil sands industry is rapidly expanding to meet the increasing global and Canadian

energy demand. Paslawski et al. (2009) reported a cumulative production of 1.1 million barrels

of synthetic crude oil per day from bituminous ore by Canadian oil sands companies [72].

Despite the economic benefits of the oil sands development, the industry faces some major

challenges due to the negative environmental impacts caused by its mining, extraction, and

upgrading operations.

The management of tailings has been one of the major issues with the oil sands operations.

Currently, the oil sands companies hold regulatory permits for water diversion from nearby

sources with maximum volume set forth by governmental officials. In 2006, Syncrude diverted

approximately 33.9 million cubic metres of Athabasca River water [85]. A major portion of the

water is required for the bitumen extraction step, which utilizes an alkaline hot water extraction

process for bitumen recovery [9, 16, 26]. After extraction, the process water and the tailings

consisting of sand slurry, clay minerals, residual bitumen, and leached organic and inorganic

components of the ore are stored in containment facilities known as tailings ponds [26]. These

ponds were formerly excavated mine pits. Oil sands tailings ponds are often described as holding

basins that allow solids to settle, resulting in an overlaying water layer. This water is most

commonly known as oil sands tailings pond water (TPW). A fraction of the water is recycled for

process operations [26]. Tailings pond water has been proven to be toxic to a variety of

organisms, and thereby cannot be released into the environment. Currently, a zero discharge

policy implemented by the government prohibits the release of TPW into the environment [76].

The volume of tailings pond water is increasing with the production of the synthetic crude oil

from bitumen. According to Small (2011), a slurry waste volume of more than 4 x 108 m

3 is

stored in the tailings ponds throughout the Athabasca region of Alberta. It is evident that the

3

increasing volume of oil sands tailings pond water is a major environmental concern. Thereby,

effective remediation and reclamation of the tailings ponds are vital for the sustainable

development of the Canadian oil sands.

1.1.2 Oil Sands Naphthenic Acids

Naphthenic acids (NAs) are naturally occurring constituents of bituminous oil sands deposits,

originating from depositional and post-depositional environments [9, 71, 85]. They are

solubilized from the ore during the bitumen extraction step, and released into the tailings pond

water. NAs have been defined as complex mixtures of alkyl-substituted acyclic and

cycloaliphatic carboxylic acids. They have the general chemical formula, CnH2n+zO2, where n

specifies the number of carbon atoms and z is zero or a negative even integer, indicative of

hydrogen deficiency due to the presence of ring structure [8, 9, 16, 26,76]. Oil sands NAs

typically have molecular weight greater than 120 g/mol.

Naphthenic acids are proven to be one of main toxic components of tailings pond water,

responsible for causing acute and chronic toxicity within microorganisms [8, 9, 16, 26, 28, 76].

Due to their water solubility, they are particularly harmful to aquatic wildlife [9]. Naphthenic

acids concentration greater than 2.5 ppm is toxic to fish species [9]. The concentration of NAs in

tailings pond water ranges from 20-120 ppm, whereas the concentration in Athabasca River is

typically less than 0.01 ppm [8, 9, 16, 26, 28, 76]. Potential leaching of naphthenic acids into the

nearby environment is a concern, and consequently there is a strong environmental incentive

behind the removal of these compounds from tailings pond water.

Naphthenic acids are also natural surfactants, comprised of polar hydrophilic head groups

(containing oxygen heteroatoms) and non-polar hydrophobic tail groups (carbon chain) [9, 38,

4

42]. This property allows the compounds to stabilize residual bitumen-in-water emulsions by

reducing interfacial surface tension, limiting the potential recovery of residual bitumen from

tailings ponds [38]. NAs also diminish the settling properties of fine clay minerals through

surface activities, which leads to slower settling rates of the minerals [38]. This effect is

undesirable for water reclamation.

It is apparent that naphthenic acids contribute to the environmental burden of oil sands

operations and diminish the reclamation and remediation efforts of tailings ponds. Therefore, the

removal of these compounds from oil sands tailings pond water must be investigated.

1.1.3 Adsorption of Naphthenic Acids by Activated Carbon

Biodegradation, ozonation, photocatalytic degradation and adsorption by various media have

been among some of the techniques investigated for naphthenic acids removal from tailings pond

water [8, 9, 10, 28, 30, 36, 57]. Removal of contaminants by adsorption is a well-developed

wastewater treatment technology. Activated carbon is one of the most commonly used

adsorbents for wastewater treatment. Due to intricate porosity and high specific surface area,

activated carbon possesses high adsorption capacity for contaminant species. The removal of

naphthenic acids by activated carbon adsorption is a promising solution to the problem presented

in this study. However, commercially produced activated carbon can be expensive. The

treatment of tailings pond water using commercially activated carbon may not be an

economically feasible alternative.

Oil sands petroleum coke (PC) is a well-known waste-product of the bitumen upgrading process.

It is a carbonaceous solid (80-85 wt% carbon) produced during a de-carbonization step to

5

convert heavy petroleum residues to synthetic crude oil [49, 71]. Although petroleum coke is a

potential combustible fuel source, due to its high sulphur content (6-7 wt%), combustion would

invariably produce large quantities of sulphur dioxide (SO2), which is not environmentally

feasible [49]. Consequently, petroleum coke has been accumulating as a by-product in oil sands

industry for many years. The Energy Resources Conservation Board of Canada has reported a

petroleum coke production rate of 20,000 tonnes/day in 2009, with an estimated increase to

about 40,000 tonnes/day by 2010 [4].

Recently, researchers have investigated the potential of petroleum coke as a precursor for

activated carbon [4, 49, 71]. Raw coke is non-porous and consequently has very low specific

surface area (~0.15 m2/g) [26]. However, studies have been indicated that petroleum coke can be

activated (thermally treated with the presence of a reagent) to produce activated carbon with high

specific surface area (~500-1500 m2/g) [4, 49, 71, 85]. High specific surface area generally

results in high adsorption capacity for contaminants. In their work, Yuan (2010) and Awoyemi

(2011) demonstrated the effectiveness of petroleum coke-derived activated carbon for the

removal of polycyclic aromatic hydrocarbons.

There is a strong environmental driving force behind the removal of naphthenic acids from

tailings pond water. Petroleum coke is vastly abundant, and presents itself as an effective

activated carbon precursor. This research thesis investigates the adsorptive removal of

naphthenic acids from oil sands tailings pond water by petroleum coke-derived activated carbon.

6

1.2 Research Objectives

The ultimate research goal was to investigate the adsorptive removal of naphthenic acids from

tailings pond water by petroleum coke-derived activated carbon. It should be noted that a model

naphthenic acid was used for most of the investigation. This is so that the non-competitive

adsorption process could be elucidated by establishing a baseline understanding of a single

naphthenic acid species before undertaking more complex naphthenic acid mixtures. The

following research objectives were investigated to answer the corresponding research questions.

Synthesize activated carbon from petroleum coke by various activation methods to

enhance the adsorption capacity of adsorbents; characterize the physical and chemical

properties of the newly synthesized activated carbons;

1. How do different activation methods affect the physical pore properties of the

adsorbents?

2. How would surface oxidation and basic surface treatment alter the chemical

surface functionalities of the adsorbents?

Perform batch-type time-dependent adsorption studies with the petroleum coke-derived

activated carbons to investigate the adsorption kinetics of the model naphthenic acid;

1. Which kinetic rate models best represent the experimental data?

2. How do the specific surface area and the pore size distribution of activated carbon

influence the rate constant?

3. What are the effects of acidic surface functionalities of adsorbents on the rate

parameters?

7

4. What is the rate-controlling step of the adsorption process?

Investigate the equilibrium adsorption of the model naphthenic acid;

1. Which adsorption isotherm model best describes the experimental data?

2. How does the activated carbon pore structure influence the equilibrium adsorption

capacity?

3. What are the effects acidic and basic surface functionalities on the equilibrium

adsorption capacity?

Investigate the equilibrium adsorption of total acid-extractable organics (complex

naphthenic acid mixtures) from tailings pond water using petroleum coke-derived

activated carbon.

1. Under the same conditions, how do the adsorption of total acid-extractable

organics compare to that of the model naphthenic acid? What are the

implications?

8

CHAPTER 2 OIL SANDS NAPHTHENIC ACIDS AND ACTIVATED CARBON:

A LITERTURE REVIEW

2.1 Oil Sands Naphthenic Acids

2.1.1 Origin

Recall from the first chapter that naphthenic acids (NAs) are one of the natural components

found in bituminous oil sands deposits. Although the oil sands industry has been moving towards

using an in-situ steam-assisted gravity drainage (SAGD) process for bitumen recovery, surface

mining/extraction methods have been more commonly employed by companies over the last few

decades [85]. The latter method has been argued to be the more efficient of the two, but it also

produces the greater amount of overburden materials [85]. In this method, bitumen is recovered

from mined ore using an alkaline hot water extraction process [26]. The resulting tailings from

the extraction process are then stored in oil sands tailings ponds containing slurry of sand, silt,

clay, organic and inorganic components of bitumen and residual bitumen. It is during this

extraction step that naphthenic acids find their way into the tailings ponds water [16, 17, 28].

2.1.2 Structure and Classification

Naphthenic acids are a complex group of alkyl-substituted acyclic and cycloaliphatic carboxylic

acids with the general chemical formula CnH2n+zO2, where n indicates the total number of carbon

atoms and z (either zero or a negative even integer) represents hydrogen deficiency in a

compound due to presence of ring structure [17, 28, 47, 85]. The magnitude of z value divided

by two indicates the number of rings present in a compound. Figure 2-1 displays the chemical

structures of a few naphthenic acids with different z values.

9

Z = 0

Z = -2

Z = -4

Figure 2-1: Examples of acyclic (z= 0), monocyclic (z= -2) and bicyclic (z= -4) naphthenic acid

structures

10

Recent advances in analytical tools revealed that the naphthenic acids classified by the formula

CnH2n+zO2 can undergo mild oxidation in tailings ponds to form ‘oxygenated’ or ‘oxy-naphthenic

acids’ with the formula CnH2n+zOx, where x = 2-5 [26]. Han et al., (2009) detected mono- and

di-oxy naphthenic acids (i.e., CnH2n+zO3 and CnH2n+zO4) in Syncrude tailings pond water [28].

According to Whitby (2010), NAs found in oil sands deposits may also contain pyrrole,

thiophene and phenol groups [77]. Hence, unique classification of these carboxylic acids is rather

difficult, in that, the term ‘naphthenic acid’ has become somewhat ambiguous. As of recent

years, the NAs represented by the formula CnH2n+zO2 have been referred to as ‘classical’ NAs

[26]. The variability in the chemical structures of NAs also presents a large challenge in the

quantification and characterization of the compounds [71].

2.1.3 Mobility and Toxicity

The pH of oil sands tailings ponds water typically ranges between 8-9. Zubot (2010) reported the

average pH of Syncrude TPW to be around 8.5. This is mainly due to the use of caustic solution

in the bitumen extraction process. The solubility of naphthenic acids in TPW is pH dependent –

where the dissociation constants of NAs generally range between 10-5

and 10-6

(pka = 5-6) [17].

Hence, NAs are readily soluble in TPW, where they exist as naphthenate salts (most commonly

as sodium and calcium naphthenates) [85]. The water solubility favours the transport of NAs in

surface water and groundwater, which would allow for the uptake by plants and animals [71].

Toxicological research has proven NAs to be the main components of TPW which are

responsible for acute and chronic toxicity in a variety of organisms such as fish, amphibians,

zooplankton and some mammals (rats and guinea pigs) [71, 77]. In order to emphasize the

degree of toxicity found in tailings pond water, Clemente and Fedorak (2005) reported

11

naphthenic acids concentration greater than 2.5-5 ppm to be toxic to fish species, where the

concentration of NAs in tailings ponds water ranges between 20-120 ppm [8, 17]. In comparison,

the Athabasca River (a nearby freshwater source) has a NAs concentration of less than 0.01 ppm

[66]. Although the toxicity imposed by NAs are species specific, and the level of toxicity varies

with structural composition, the leaching of TPW into groundwater and nearby fresh water

sources can have dangerous environmental consequences [17, 40].

2.1.4 Chemical Properties

Naphthenic acids are also surfactants or surface-active agents. They contain polar (hydrophilic)

head groups comprised of oxygen heteroatoms and non-polar (hydrophobic) tail groups

comprised of long carbon chains [17, 38, 42]. This unique characteristic allows these molecules

to partition to both oil and water phase simultaneously. By reducing the interfacial surface

tension, NAs can stabilize residual bitumen-in-water emulsions in tailings ponds, limiting the

potential for residual bitumen recovery from TPW [42]. Furthermore, NAs also diminish the

settling properties of fine clay minerals in tailings ponds water through surface activities, which

leads to slower settling rate [38]. Jiang et al., (2011) demonstrated that the addition of sodium

naphthenate to synthetic TPW hinders the settling properties of kaolinite, which is one of the

most predominant clay mineral in TPW.

Recent studies have indicated that recycling NA-containing TPW for process operations may

lead to corrosion of process equipment [17, 71]. Although the corrosion mechanism is not well

understood, it has been theorized to involve chelation of metal ions by carboxylate ions

(naphthenates), which subsequently leads to the formation of hydrogen gas [17]. Corrosion may

lead to equipment failure, and consequently raises safety and economic concerns.

12

In summary, naphthenic acids contribute to the environmental and economic burden of oil sands

operations. The removal of these compounds is not only vital to the reclamation and remediation

efforts of tailings ponds, but also to the sustainable and economically feasible development of

Canadian oil sands resources.

2.1.5 Analytical Techniques for Detection and Quantification

Due to the range of structural variability, accurate quantification and characterization of oil sands

naphthenic acids have been a challenge. A number of analytical techniques have been developed

for the purpose. Some are solely used to determine the total NAs concentration from aqueous

phase while others can reveal information in regards to structural composition.

A Fourier-transform infrared (FTIR) spectroscopy-based method developed by researchers at

Syncrude Ltd. has been used most commonly for detecting and quantifying NAs [39]. The

detection is based on infrared (IR) absorbance of carboxylic acids in organic liquid phase. For

this reason, NAs are extracted into an organic phase from acidified aqueous phase using

dichloromethane [81]. In liquid or solid state, most carboxylic acids exist as hydrogen-bonded

dimers. However, in dilute solutions, equilibrium exists between the acid monomers and dimers

[66]. The absorbance peak heights of the monomeric and dimeric forms are measured at

wavenumbers 1743 and 1704 cm-1

, respectively [17, 77]. The combined peak height is directly

proportional to NAs concentration. Hence, NAs concentration of an unknown sample is

determined by comparing the combined absorbance peak height of the sample to those in a

calibration curve obtained from the FTIR analyses of solutions prepared from commercially

available NAs [17].

13

Due to the robustness and the cost-effectiveness of the FTIR technique, it has been utilized

extensively in the oil sands industry and in the scientific community [17, 66, 77]. However, the

technique has certain limitations. Since the detection method is based on the IR absorbance of all

carboxylic acids in a solution, the technique cannot differentiate between different naphthenic

acid species. In other words, it only measures the total NAs concentration without revealing any

information about the structural composition. For the same reason, the technique cannot

differentiate between classical, oxy- and unconventional naphthenic acids with phenol, thiophene

and pyrrole groups. Recently, the term ‘total acid-extractable organics’ or TAOs has been used to

describe the organic acids detected and quantified by this technique from TPW [26, 70, 86]. The

detection limit of NAs/TAOs for this method is typically around 1 ppm [16].

Gas chromatography coupled with mass spectroscopy (GC-MS) has also been used to quantify

and characterize NAs. This method requires a pre-derivatization step to yield t-butyldimethylsilyl

esters of naphthenic acids, which are then analyzed by GC-MS [16, 17, 31, 45]. The detection

limit of this method has been reported to 0.01 ppm [16]. However, the method assumes complete

derivatization of NAs by derivatizing agent, which may not always be the case.

Several high performance liquid chromatography (HPLC) -based methods have been developed

for quantifying NAs. Yen et al., (2004) developed a HPLC method where NAs are derivatized to

2-nitrophenylhydrazides. The derivatized compounds are detected using a UV-Vis diode array

detector. The detection limit of this method was 5 ppm [81]. Complete derivatization is also

assumed in this method.

For some areas of research and analyses, it is not sufficient to only determine the total

naphthenic acids concentration but also important to know the molecular structures and

14

compositions of NAs in a mixture (i.e., toxicological research). For the latter purpose, mass

spectroscopy (MS) and its derivatives provide the best information [17]. Recent technological

advances allow for the usage of sophisticated tools such as HPLC coupled with high-resolution

mass spectrometer (HRMS) or tandem mass spectrometer (i.e., quadruple time-of-flight MS) for

compositional analyses of NAs [8, 28]. However, these instruments can often be very expensive

and difficult to use, and consequently unsuitable for every day laboratorial use.

For determining the total concentration of NAs/TAOs or the concentration of a single (surrogate)

NA in a solution, the FTIR method is still one of the most common and robust analytical

techniques.

2.1.6 Remediation Techniques

Effective remediation of naphthenic acids is vital to the reclamation efforts of oil sands tailings

ponds. Biodegradation, photocatalytic degradation, ozonation, nanofiltration, sequestration using

cyclodextrin-based polymers and adsorption by various media are some of the techniques

investigated and reported in the literature for NAs remediation from TPW [17, 20, 40, 47, 48,

70].

Microbial degradation of NAs typically produces CO2, fragmented organic compounds and water

[40]. However, the extent of degradation is dependent on the molecular weight and the chemical

structure of the compounds. Scott et al., (2005) reported that the low molecular weight NAs are

more susceptible to biodegradation than high molecular weight NAs. It has also been suggested

that the rate of biodegradation is primarily affected by the chemical structures of NAs, where the

more recalcitrant NAs are the ones with higher degrees of aliphatic chains, methyl-substituted

15

cycloalkane rings and increased cyclicity [40]. The effectiveness of photocatalytic degradation is

also limited due to its dependency on the molecular weights and cyclicity of the compounds [20,

40]. The ozonation of naphthenic acids does not lead to complete degradation, and consequently

results in formation of several by-products such as aldehydes, ketones and peroxides and other

carboxylic constituents [40]. Some of these by-products may even be more hazardous than the

original compounds.

Adsorption is one of the most effective methods for remediating wastewater due to its high

contaminant removal efficiency [4]. Persistent contaminants can be fully removed, rather than

breaking down into smaller and potentially more harmful fragments as is with the remediation

techniques described above [5]. A few studies have investigated the effectiveness of organic rich

soil, zeolites, raw petroleum coke, cyclodextrin-based polymers and activated carbon as

adsorbents for NAs removal [36, 47, 48, 57, 70, 85, 86]. One of the main advantages of using

activated carbon is that its physical and chemical properties can be tailored to target specific

species of contaminant; thereby increasing the overall adsorption affinity of the adsorbent for the

contaminants. Mohamed et al., (2010) reported that the sorption capacity of granular activated

carbon (GAC) for NAs is much greater than polymeric sorbents. Wu et al., (2001) demonstrated

enhanced adsorption of carboxylic acid-based surfactants when using activated carbon.

Nonetheless, the full potential of activated carbon for oil sands naphthenic acids removal has yet

to be investigated and exploited thoroughly.

16

2.2 Activated Carbon (AC)

The term “activated carbon” refers to a wide range of processed amorphous carbonaceous

materials with a highly developed porosity and an extended internal specific surface area (SSA).

The latter typically ranges between 300-3000 m2/g [4, 6, 7]. AC is generally obtained by

combustion, partial combustion, or thermal decomposition of a variety of organic precursor

materials with high carbon content (i.e., coconut shells, wood, peat, and bituminous coal) [7, 71].

The resulting porous carbon materials have a wide range of applications. Activated carbon is

commonly used as an adsorbent in industrial and environmental applications. Current research

has demonstrated the potential for acting as a catalyst support medium in catalytic processes or

as an energy storage device in electrochemical double layer capacitors (EDLC) as well [4].

2.2.1 Applications of an Adsorbent

Historically, activated carbon has been used for drinking water prolongation and odour removal

purposes [6]. The First World War marks the starting point of activated carbon development and

production as an adsorbent for aqueous and vapour phase treatments [71]. Presently, activated

carbon is extensively used as an adsorbent in industrial and environmental remediation

applications such as removal of pollutants from groundwater and industrial wastewater,

hazardous gas scrubbing from industrial waste streams, and volatile organic compound capture

from contaminated sites [4, 6, 7, 49, 71]. Due to its high specific surface area (SSA), activated

carbon has high adsorption capacity for a wide range of adsorbates [6, 7].

Liquid phase contaminant removal by activated carbon adsorption has gained much attention

over the last few decades. Numerous publications can be found in the literature involving heavy

17

metal immobilization and subsequent removal by activated carbon [1, 2, 43, 61]. Ramana et al.,

(2010) reported effective removal of Cu2+

, Cd2+

, Pb2+

, and Zn2+

ions from aqueous solution using

activated carbon derived from agricultural solid waste. Activated carbon has also been used to

remove polycyclic aromatic hydrocarbons (PAHs) from wastewater. Valderrama et al., (2008)

and Yuan et al., (2010) are among some of the researchers who investigated PAHs adsorption by

activated carbon.

Studies have shown that activated carbon is also effective in removing surfactants from aqueous

phase [75, 78]. Activated carbon has high adsorption affinity for surfactants due to the strong

hydrophobic interaction between the AC surface and the non-polar carbon chain of the

surfactants [75]. Both Pendleton et al., (2001) and Ihara (1992) demonstrated effective removal

of anionic surfactants by activated carbon adsorption.

2.2.2 Precursor Materials for Activated Carbon Production

Since activated carbon is non-graphitic in structure, almost any carbonaceous material can be

converted into activated carbon [6]. However, it is generally desirable to use inexpensive and

abundant raw materials with high carbon content and low inorganic content for the production of

activated carbon [6, 7]. Inorganic materials are non-porous, and their presence reduces the total

SSA and the adsorption capacity of an adsorbent [6]. Readily available materials such as wood,

lignocellulosic biomass, peat, lignite, coconut shells, and different grades of bituminous coals

have been among some of the most commonly used AC precursors [4, 7]. Due to the increase in

activated carbon production, there is a high demand for inexpensive and abundant raw materials.

Hence, alternative sources of activated carbon have been explored in recent years. Oil sands

18

petroleum coke, a carbonaceous waste-product of the oil sands upgrading process, has gained

much attention as an activated carbon precursor [4, 49, 70, 71, 72, 85, 86].

2.2.2.1 Oil Sands Petroleum Coke

Oil sands petroleum coke is a well-known carbonaceous by-product of the bitumen upgrading

process [49]. Heavy petroleum residues must be broken down or ‘cracked’ into smaller units to

produce synthetic crude oil. The carbon rejection process utilized to produce lighter petroleum

fractions is known as ‘coking’ [71]. Petroleum coke (PC) is typically produced through either

dealkylation or dehydrogenation reactions during thermal cracking [71]. Coke composition often

varies because the hydrocarbon composition of petroleum itself is variable depending on

geographical location of the oilfield. Operational parameters of the coking process also affect the

morphology and the chemical/structural composition of petroleum coke. Oil sands companies

often use different coking techniques [71]. Suncor Energy’s coking process produces ‘delayed’

petroleum coke, which has a sponge-like structure. In contrast, Syncrude’s coke is known as

‘fluid’ coke, and it has highly-graphitized layers and often described to have an onion-like

structure [4, 49, 71].

Despite the variability in the production process, petroleum coke typically has a high carbon

content of 80-85 wt% [6]. It cannot be used as a combustible fuel source like conventional coal

due to its high sulphur content (~6-7%) [71]. Small (2011) estimated a cumulative annual

production of over 5 million tonnes of petroleum coke by Suncor Energy Inc. and Syncrude

Canada Ltd. Many years of stockpiling has led to substantial accumulation of the coke product at

the upgrading facilities.

19

Current research has indicated that oil sands petroleum coke can be readily utilized as a

precursor for activated carbon production [4, 13, 19, 49, 50, 68, 70, 71, 72, 82]. Activated carbon

with high specific surface area (SSA) is achievable through both the physical and chemical

activation of petroleum coke. With respect to the former, Awoyemi (2011) and Small (2012)

prepared activated carbon (SSA~500 m2/g) from oil sands petroleum coke using carbon dioxide

and steam-based physical activation methods and used them for liquid phase treatments. Yuan et

al., (2010) reported effective removal of polycyclic aromatic hydrocarbons using KOH-activated

petroleum coke (SSA~1900 m2/g) from aqueous phase. Bratu (2008) and Morris (2012) studied

gas-phase mercury vapour removal using petroleum coke-derived activated carbon.

2.2.3 Activated Carbon Production Processes

Activated carbon is prepared by altering the internal structure of precursor materials to develop

high porosity, which results in high specific surface area. According to Bansal (2005), the

preparation of activated carbon involves two basic steps: (1) thermal carbonization of raw

material in inert atmosphere and (2) activation of carbonized char by an activating agent. During

the carbonization step, non-carbon elements such as oxygen, nitrogen, hydrogen and sulphur are

removed from the precursor material as gaseous products due to pyrolytic decomposition [6, 7].

The remaining carbon atoms then randomly reorient themselves into stacks of cross-linked

aromatic sheets, consequently forming irregular interstices or pores between the sheets [4, 49,

71]. These pores are often blocked or filled by decomposed carbon products, resulting in low

surface area. Hence, an activation process is utilized to remove the disorganized carbons and

further extend and develop the porous structure [7, 71]. Consequently, the choice of activating

agent affects the total pore volume and the pore size distribution of activated carbon. Sing et al.,

20

(1985) of IUPAC divided the pore network of activated carbon into three distinct categories:

micropores (pore diameter < 2 nm), mesopores (2 nm <pore diameter < 50 nm), and macropores

(pore diameter > 50 nm). Factors influencing the porosity of activated carbon will be discussed

later in section 2.2.3.4.

Although there are many ways of producing activated carbon, and hundreds of publications can

be found in the literature, all activation methods can be clearly defined into two types: physical

and chemical activation [6].

2.2.3.1 Physical Activation

In physical activation, the first step is the thermal carbonization of raw material to remove

volatile components and produce a char rich in carbon [6]. During the second step (activation),

the pores created during carbonization are further developed by selective oxidation of the

carbonized products at high temperatures (~800-1000 oC) [4]. The activating agents (typically

oxidizing gases) react with the most reactive atoms in the carbon skeleton, removing some of the

internal surface mass from the solid and opening up previously blocked pores [6]. This leads to

the development of intricate pore structure and high SSA. In laboratorial settings, the two steps

(carbonization and activation) are often carried out simultaneously in heating furnace or

temperature-controlled fixed-bed reactors [4, 49]. The most commonly used activating agents for

physical activation are steam, carbon dioxide, and air. The rate of reactivity of the reagents can

be summarized in the following trend: O2> H2O > CO2 [71]. Despite its high reactivity, oxygen

is not generally used for activation. This is because physical activation with oxygen is an

exothermic reaction, and the carbon burn-off is often difficult to control [71]. Steam and carbon

21

dioxide activation both follow endothermic gasification reactions. The chemical reactions and

their respective heat of reactions are listed below [6]-

Steam Activation: C(s) + H2O(g) H2(g) + CO(g) ΔH = + 132 kJ mol-1

(2.1)

CO2 Activation: C(s) + CO2(g) 2CO(g) ΔH = + 159 kJ mol-1

(2.2)

Selection of an activating agent depends on the final application of activated carbon. Both Small

(2011) and Pastor-Villegas et al., (2001) found that steam-activated carbon generally displays

higher total pore volume and SSA than CO2-activated carbon. Steam activation typically

produces a homogenous pore size distribution, developing mostly micropores [6]. Micropores

are major contributors of total pore volume and specific surface area. CO2 activation can produce

an appreciable amount of mesopores as well as micropores [54]. Water molecules are a lot

smaller in dimension than carbon dioxide molecules. Hence, it is much easier for the former to

diffuse into the carbon pore structure, which leads to the development of more micropores [54].

2.2.3.2 Chemical Activation

Chemical activation is typically carried out in a single carbonization step [6]. The precursor

material is initially impregnated with a concentrated dehydrating reagent (activating agent) and

then heated up to a temperature between 500-800 oC in inert atmosphere for carbonization [6,

71]. Some of common activating agents for chemical activation include zinc chloride (ZnCl2),

phosphoric acid (H3PO4), potassium hydroxide (KOH), sulphuric acid (H2SO4), and ferric

chloride (FeCl3) [6, 71]. The activating agents dehydrate the raw material. Heat treatment of the

impregnated material results in charring and aromatization of the carbon structure [32]. Any

22

remaining activating agents are removed by water or acid washing [71]. According to Hsu and

Teng (2000), chemical activation leads to the creation of activated carbon structure with higher

SSA than physical activation.

Some activating agents are known to produce a certain type of pore size distribution. For

example, it has been suggested by Yang (2003) that KOH activation generates mostly

microporous activated carbon, where ZnCl2 primarily produces mesoporous activated carbon.

The effectiveness of chemical activation is dependent on the impregnation ratio (ratio of

activating agent to raw material) as well as the carbonization and activation temperature [71].

Chemical activation requires a lower temperature range (~500-800 oC) and generally results in

higher SSA in comparison to physical activation. However, the corrosive nature of the chemical

activating agents has negative environmental impacts and often limits its application [4].

2.2.3.3 Post-modification of Activated Carbon

Activated carbon is often post-modified using various methods to produce desired functionalities

for its final application [14]. Although post-modification influences the physical properties of

activated carbon to an extent, the treatments are generally utilized to introduce different chemical

functional groups on the carbon surface. Presence of certain surface functional groups (SFGs) on

activated carbon can enhance the adsorption affinity and capacity for corresponding groups of

adsorbates.

Selection of post-treatment methods depends on the final application of activated carbon.

Przepiorski et al., (2003) reported that post-modified activated carbon with increased basicity

enhances the adsorption of acidic pollutants. Introducing nitrogen-based surface functional

23

groups onto the surface of activated carbon can increase its basicity [58]. One of the common

post-treatment methods used for this purpose is thermal treatment of activated carbon in gaseous

ammonia environment at a temperature between (500-800 oC) [44, 58]. This treatment typically

introduces nitrogen-based SFGs such as –NH2, –CN, pyridinic, pyrrolic, and quaternary nitrogen

on the carbon surface [58].

Post-oxidation of activated carbon is another common post-treatment method. In contrast to the

ammonia treatment, post-oxidation introduces acidic surface functional groups such as lactonic,

phenolic, carboxylic, and anhydride groups [11]. Awoyemi (2011) studied the effect of post-

oxidized activated carbon on aqueous phase polycyclic aromatic hydrocarbon (PAH) adsorption.

Post-oxidation is typically carried out by mild heat treatment of activated carbon in oxygen-rich

environment [4, 33].

2.2.3.4 Factors Influencing the Physical and Chemical Properties of AC

Aside from the precursor material and the activation method (physical vs. chemical), several

other factors such as the activating agent, the activation temperature, the rate of heating, and the

reaction time also affect the physical and the chemical properties of activated carbon. Generally,

the specific surface area (SSA) increases with activation temperature and the reaction time, but at

the cost of product yield [4]. However, too high of a reaction temperature can lead to burn-off

between the walls of adjacent micropores, thereby increasing number of meso- and macropores

and reducing the total SSA [71]. During a SO2-activation of petroleum coke, Morris (2012)

found an optimal reaction temperature that yields the maximum SSA. The rate of heating also

affects the physical properties of activated carbon. According to Bansal et al., (1988), precursor

materials undergo a softening period before it begins to harden and shrink. A slow heating rate

24

results in a denser char, but at the cost of limited volatilization [7]. A slow heating rate is also

known to favour the formation of micropores [71].

Controlling the pore size distribution of activated carbon can be important for its final

application. Template methods using zeolites or silica are often used to control the pore size

distribution [4]. These methods usually involve carbonization of raw material in the nano-space

of a template and subsequent liberation of the resultant activated carbon from the template [4].

Fuertes et al., (2004) used meso-structure silica template to synthesize mesoporous activated

carbon.

Chemical properties such as the surface functional group (SFG) content and the point of zero

charge (pzc) are primarily dependent on the nature of activating agents. For example, physical

activation with sulphur dioxide (SO2) leads to an increase of sulphur-based SFGs [49]. Post-

modification with different reagents could also lead to alternation of the surface chemical

properties. Ammonia-treated activated carbon exhibits alkaline characteristics, whereas post-

oxidized activated carbon has acidic characteristics [33, 58].

25

CHAPTER 3 THEORETICAL OVERVIEW OF ADSORPTION

Adsorption is a process by which atoms, ion, biomolecules or molecules of gas, liquid or

dissolved solids adhere to the surface of a liquid or solid [4]. The species adsorbing to a surface

are known as adsorbates, while surface species are known as adsorbents. The nature of an

adsorption process can be categorized into two types: physisorption (physical adsorption) and

chemisorption (chemical adsorption).

3.1 Physisorption vs. Chemisorption

The distinction between the adsorption processes arises from different forces of attraction that

exist between the adsorbate and the adsorbent surface [1]. In physisorption, the forces involved

are intermolecular (i.e., van der Waals forces and hydrogen bonding), and there is no significant

change in the electronic orbital patterns of the adsorbate and the adsorbent species [3]. The

amount of energy released from physical adsorption is of the order of the enthalpy of

condensation. Atkins and Paula (2002) reported the enthalpy changes of physisorption process to

be ~20 kJ/mol or less. In contrast, chemisorption involves the formation of chemical (covalent)

bonds between the adsorbate and the adsorbent surface, the same kind as those operating in

formation of chemical compounds [4]. Consequently, the energy of adsorption is much higher

than that of physisorption, ranging between 40-200 kJ/mol [3, 12, 64]. In physisorption, the

adsorbed species are chemically identical with those in the fluid phase; hence the process is said

to be reversible. However, in chemisorption, the chemical nature of the adsorbate may be altered,

and the adsorption process may not be reversible. According to Everett (1971), in physisorption,

molecules can adsorb in excess to adsorbates that are in direct contact with the surface. This

26

results in multilayer adsorption. In contrast, the chemically-bonded adsorbates in chemisorption

generally form a uniform monolayer [21].

3.2 Adsorption Kinetics

Mathematical expressions or models are often formulated and utilized to understand and predict

the kinetics of an adsorption process. Numerous empirical and mechanistic kinetic models exist

in literature for liquid-solid adsorption system. Generally, empirical models are not derived from

first principles. Unlike mechanistic models, they do not offer very much insight into the

mechanism of adsorption. Nonetheless, these models provide a good understanding of adsorption

kinetics from a macroscopic perspective, and the model parameters (rate constants) can be used

for designing adsorption process. Three most commonly used empirical models cited in literature

for liquid-solid adsorption system are described below.

3.2.1 Empirical Models

3.2.1.1 Pseudo-first Order Model

The pseudo-first order model of Lagergren is described by the following equation [29, 59, 79,

83]-

where k1 is the pseudo-first order rate constant (min-1

); qe and qt are the amount of solute

adsorbed at equilibrium and at time, t, respectively (mg/g). Integrating equation 3.1 for the

boundary conditions t = 0 to t = t and qt = 0 to qt = qt yields the following expression-

27

The rate constant, k1 and the calculated qe value can be obtained by the linear regression of log

(qe – qt) vs. time. Ho and Mckay (1999) analyzed kinetic data from various adsorption studies

and found that the pseudo-first order model mostly applicable for the initial stage of the

adsorption but not whole range of reaction times. This phenomenon was also observed by Zhang

et al., (2010).

3.2.1.2 Pseudo-second Order Model

Yalcin (2004) and Zhang (2010) used the pseudo-second order model for surfactant adsorption.

According to Ho and McKay (1999), the differential form of the model can be expressed by the

following equation-

where k2 is the pseudo-second order rate constant (g/mg min). Integrating equation 3.3 with the

same boundary conditions as described for the pseudo-first order model yields-

The pseudo second order rate constant, k2 and the amount of solute adsorbed at equilibrium, qe

can obtained by the linear regression of equation 3.4. Taking the limit as time approaches zero,

an additional parameter, ho, is defined as the initial rate of adsorption (mg/g min) [29, 39]. ho is

defined by the following expression [29]-

28

3.2.1.3 Elovich Model

Elovich model is another empirical model expressed by the Elovich equation [35]-

where α is the initial adsorption rate (mg/g min) and β is the desorption constant. Equation 3.6

can be simplified by assuming αβt >> t, and by applying the boundary conditions qt = 0 at t = 0

and qt = qt at t = t the equation becomes-

Elovich constants α and β can be obtained by the linear regression of equation 3.7.

3.2.2 Mechanistic Models

It is important to understand the adsorption mechanism of a process in order to identify the rate-

limiting step. The mechanism of a solid-liquid adsorption process has been generally described

by the three following consecutive steps [59, 79]-

1) Film or external diffusion: represents the transport of the adsorbate molecules from the

bulk liquid phase to the exterior surface of the adsorbent.

2) Internal pore diffusion: involves the transport of the adsorbate molecules from the

external surface to the interior pore structure of the adsorbent.

29

3) Surface adsorption: represents the adsorption of the adsorbate molecules onto the active

sites of the interior adsorbent pores.

All three steps affect the rate of adsorption to an extent. However, some steps occur significantly

faster than others. If the surface adsorption step does not involve any chemical modification of

the adsorbent surface, in other words, if the adsorbates are bound to the surface by physical

forces (van der Waals), then the third step occurs rapidly [18]. In that case, one of the diffusion

steps (external or internal) would dominate the kinetic process. Figure 3.1 depicted the three step

conceptual model described above.

Figure 3.1: Schematic representation of a three-step conceptual model for solid-liquid

adsorption processes

30

3.2.2.1 Weber-Morris Intraparticle Diffusion Model

The Weber-Morris (W-M) intraparticle diffusion model is often used to determine the

intraparticle or internal diffusional rate constant of an adsorption process. The model is described

by the following equation [35, 59, 79]-

where qt is the amount of solute adsorbed (mg/g) at time, t; kid is the intraparticle diffusion rate

constant (mg/g/min0.5

) and C (mg/g) is a constant that reflects the extent of boundary layer

effects [79]. The model parameters are obtained by plotting qt vs. t0.5

. If a process is entirely

internal diffusion-controlled, then the y-intercept passes through the origin [59, 79]. However,

these plots often tend to be double-natured, consisting an initial curved portion and a final linear

portion. The initial curved portion represents the effect of external or film diffusion whereas the

final linear portion reflects the effect of internal pore diffusion [79]. The rate constant kid is the

slope of the linear portion of a W-M plot. The constant, C, is determined from the y-intercept of

the linear portion of the plot, and its magnitude reflects the boundary (film) layer effects [59, 79].

3.3 Equilibrium Adsorption Isotherms

In an adsorption process, there is a defined distribution of adsorbates between the fluid phase and

the adsorbent surface at equilibrium. An adsorption isotherm is a functional expression relating

the amount of solute adsorbed on an adsorbent at equilibrium, qe (mg/g), to the equilibrium

solute concentration in the fluid phase, Ce (mg/L), at a fixed temperature [4]. In other words,

equilibrium adsorption can be described by mathematical isotherm equations or models whose

31

parameters express the surface properties and affinity of an adsorbent at a fixed temperature [3].

Adsorption isotherms are often used to compare the adsorption capacities of different adsorbents

for a particular adsorbate species. Experimentally, qe and Ce values for an isotherm can be

obtained in two ways: (1) by adding a fixed adsorbent dosage to a series of solutions of varied

initial concentrations, or (2) by fixing the initial adsorbate concentration for a series of solutions

while varying the adsorbent dosage to the solutions. The following sections describe some of the

most commonly used and cited isotherm models in literature [3, 59, 79, 83].

3.3.1 Langmuir Model

The Langmuir model is a theoretical equilibrium isotherm model, initially developed for relating

the amount of gas adsorbed on a surface to the pressure of the gas. However, due to its simplicity

and its fundamental concept, the model has been widely applied for liquid phase adsorption [3,

83]. The Langmuir model is generally expressed by the following equation-

where qe is the amount of adsorbed solute on adsorbent surface at equilibrium (mg/g); Ce is the

equilibrium solute concentration in the solution phase (mg/L); qm is the Langmuir adsorption

capacity (mg/g); and kl is the affinity coefficient (L/mg) related to the energy of adsorption. The

key assumptions underlying the Langmuir isotherm model are listed below [3, 4]:

(1) Energetically equivalent active sites – the energy of adsorption has a homogeneous

distribution for all active sites,

(2) Monolayer coverage – each active site can only hold one adsorbate molecule,

32

(3) No interaction between adsorbed molecules.

The Langmuir equation is said to have a fundamental thermodynamic basis as it effectively

reduces to Henry’s Law at dilute adsorbate concentrations [3]. The linearized form of equation

3.9 is listed in Table 3-1.

3.3.2 Freundlich Model

The Freundlich model, an empirical isotherm model, is expressed by the following equation-

where kF is the Freundlich constant related to the adsorption capacity [(mg/g)(L/mg)1/n

] and 1/n

is reflects the intensity of adsorption. The model was derived by assuming that the adsorption

sites have an exponentially decaying energy distribution [3]. In other words, the model assumes

that the adsorbent surface is energetically heterogeneous. This assumption suggests a multilayer

adsorption process (i.e., free adsorbates can stack themselves onto adsorbate-adsorbent

monolayer). Many studies have proven the Freundlich model to be a good fit for experimental

data for rough (heterogeneous) surfaces [4, 59]. Unlike the Langmuir model, the Freundlich

model does not reduce to Henry’s Law at dilute solute concentrations. Consequently, it is

criticized for lacking a fundamental thermodynamics [3].

33

3.3.3 Temkin Model

According to the Temkin model, the heat of adsorption decays linearly rather than

logarithmically as implied by the Freundlich model. The Temkin isotherm is expressed as [3, 4]-

where RT/bT = B, which is the Temkin constant related to the heat of adsorption (J/mol); AT is

the equilibrium binding constant corresponding to the maximum binding energy (L/g); R is the

universal gas constant (8.3145 J/mol K) and T is the absolute solution temperature (K). The

linearized form of this isotherm equation is listed in Table 3-1.

Table 3-1: Linearized forms of adsorption isotherm models

Isotherm Models Linearized Isotherm Equations

Langmuir Model

Freundlich Model

Temkin Model

34

CHAPTER 4 EXPERIMENTAL METHODOLOGY

4.1 Preparation of Activated Carbon

The activated carbons used in this study were all synthesized from ‘delayed’ petroleum coke

provided by Suncor Energy. The nomenclature and brief descriptions of the adsorbents are listed

in Table 4-1. Some of the activation procedures and post-treatments were carried out by other

researchers from the Green Technology Group (U of Toronto). These adsorbents are denoted by

the (*) symbol in Table 4-1. The experimental details of the activation and the post-treatment

procedures are thoroughly discussed in the later sub-sections.

Table 4-1: Adsorbent nomenclature

Adsorbent Precursor

Material Description

CO2-ADC-6h Delayed coke CO2-activated delayed coke (reaction time = 6 hrs)

Steam-ADC-3h Delayed coke Steam-activated delayed coke

CO2-ADC-9h * Delayed coke CO2-activated delayed coke (reaction time = 9 hrs)

CO2-ADC-9h-pAir * Delayed coke Post-treated CO2-ADC-9h with air

CO2-ADC-9h-pO2 * Delayed coke Post-treated CO2-ADC-9h with humidified-oxygen

CO2-ADC-9h-pNH3 * Delayed coke Post-treated CO2-ADC-9h with gaseous ammonia

CAC Charcoal Commercially available steam-activated carbon

(Calgon - BPL)

* Adsorbents prepared by other researchers from the Green Technology Group (U of Toronto)

35

4.1.1 Materials and Instruments

The materials and the instruments used for activated carbon production are listed in Table 4-2

and 4-3, respectively.

Table 4-2: Materials used in activated carbon production

Materials Specification Supplier

N2 (gas) 99.99% BOC Gas Ltd.

CO2 (gas) 99.99% BOC Gas Ltd.

O2 (gas) 99.99% BOC Gas Ltd.

NH3 (gas) 99.99% BOC Gas Ltd.

Delayed Petroleum Coke Particle size range: 106-150 µm Suncor Energy Inc.

Table 4-3: Instruments and equipment used in activated carbon production

Instrument Specification Manufacturer

Mass flow controller GFC17 Aalborg

Quartz tube reactor 3/4 ” O.D MIE machine shop, U of Toronto

Split tube furnace VST 12/600 Carbolite

Drying oven 615 F Fisher Scientific

Sieves American standard testing VWR

Weighing balance 1702, analytical balance Sartorius

Muffle furnace M-62700 Barnstead Thermolyne

36

4.1.2. Physical Activation of Delayed Petroleum Coke

4.1.2.1 CO2 Activation

The apparatus used for CO2 activation is similar to the ones described by Jia (2009) and Morris

(2012), and it is illustrated in a schematic diagram in Figure 4-1. Prior to activation, the coke

sample was sieved to a particle size range of 106-150 µm. A quartz tube reactor (68 cm in length

and 2 cm in diameter) with a porous quartz disc located at the midpoint was mounted into a

temperature-controlled tube furnace. Approximately 10-12 g of coke was placed on the quartz

disc, resulting in a cylindrical bed. The ends of the reactor were connected to gas lines using

ball-and-socket joints sealed with high-temperature vacuum grease. Using mass flow controllers,

CO2 and N2 were supplied to the reactor from gas cylinders; each set to flow at 150 cm3/min.

Prior to activation, the reactor was purged with pure N2 for 15 minutes at room temperature. The

furnace was then set to 120oC for ~15 minutes to remove moisture from the coke sample. After

moisture removal, the system was heated up to the desired reaction temperature. Once this

temperature was reached, CO2 was introduced to the reactor using a three-way valve. This point

of time was marked as the starting point of the reaction (activation).

The bottom end of the reactor was connected to a tar collector, which was an L-shaped Pyrex

tube filled with aluminum oxide wool. It was used to capture elemental sulphur and tar residue

from the activation process. The exit gases from the process flowed through a carbon monoxide

(CO) scrubbing solution prior to venting to the atmosphere. The activation conditions are

tabulated in Table 4-4.

37

Figure 4-1: Schematic diagram of the experimental apparatus used for CO2 activation

4.1.2.2 Steam Activation

The experimental set-up for steam activation (Figure 4-2) was similar to the one for CO2

activation. In steam activation, a steam-generating apparatus replaced the CO2 gas tank. Steam

was generated using a hot plate and an Erlenmeyer flask filled with distilled water. Glass beads

were placed in the flask for effective and steady production of steam. The flask was connected to

the reactor using a stainless steel pipe. The pipe was carefully wrapped with heating tape to

prevent the condensation of steam. A three-way valve and a steam by-pass valve were used to

control the flow of steam into the reactor. Similar to CO2 activation, the system was first purged

with N2, and moisture was removed from the coke sample prior to introducing steam.

38

Figure 4-2: Schematic diagram of the experimental apparatus used for steam activation

Steam was introduced into the reactor once the temperature and the steam production had

stabilized. Due to the presence of sulphur in delayed coke, a small amount of hydrogen sulphide

(H2S) gas was produced during the activation. A sodium hydroxide solution was used to scrub

H2S. The off-gases passed through a carbon monoxide scrubber subsequently before venting.

The treatment conditions for steam activation are also summarized in Table 4-4.

39

Table 4-4: Treatment conditions for physical activation

Adsorbent

Initial

particle size

range, (µm)

Temperature,

(oC)

Reaction

Time, (hr)

Gaseous

Environment

Flow Rate,

(cm3/min)

CO2-ADC-6h 106-150 900 6 100% CO2 150

Steam-ADC 106-150 850 3 100% Steam -

CO2-ADC-9h 106-150 900 9 100% CO2 150

4.1.3 Post-treatment of Petroleum Coke-derived Activated Carbon

4.1.3.1 Post-oxidation

The petroleum coke-derived activated carbons were further modified by post-treatments. Post-

oxidative treatments were carried out in two different gaseous environments: humidified air and

humidified oxygen. For air post-oxidation, the activated carbon sample was thinly spread on a

watch glass and placed inside a muffle furnace, which contained a beaker of distilled water for

the purpose of humidification. The furnace door was kept open to allow air diffusion in and out

of the oxidation chamber. The beaker was refilled frequently to avoid complete evaporation of

water.

As described by Huynh (2012), post-oxidation with humidified-oxygen was carried out using an

experimental set-up similar to that of CO2 activation (Figure 4-3). Oxygen was humidified by

passing through an Erlenmeyer flask filled with distilled water. Since the reaction between

oxygen and activated carbon is an exothermic one, too high of a reaction temperature would

40

completely combust the carbon sample. Hence, the reaction temperature and the heating rate

were carefully controlled.

Figure 4-3: Schematic diagram of the post-oxidation process with humidified-oxygen

A thermocouple placed directly inside the reactor was used to monitor the temperature. The

following heating steps were executed in sequence for this treatment:

Heating the reactor to 230 oC at a rate 5

oC/min and maintaining for 40 minutes

Increasing temperature to 235 oC at a rate 0.5

oC/min and maintaining for 40 minutes

Increasing temperature to 240 oC at a rate 0.5

oC/min and maintaining for 135 minutes

Increasing temperature to 245 oC at a rate 0.5

oC/min and maintaining for 135 minutes

The conditions for both post-oxidative treatments are summarized in Table 4-5.

41

4.1.3.2 Post-treatment with Gaseous Ammonia

The CO2-ADC-9h sample was post-treated with gaseous ammonia to introduce nitrogen-based

basic surface functional groups. This process involved thermal treatment of the activated carbon

sample in an ammonia-rich nitrogen environment. The experimental apparatus was similar to

that of the CO2 activation. Here, NH3 replaced the CO2 gas tank, and additional scrubbing

solutions were utilized for scrubbing outlet gases. The reactor temperature was elevated to 850oC

with a N2 flow rate of 80 cm3/min. Once the temperature had stabilized, ammonia was

introduced to the system. During the activation process, a 4:1 ratio of nitrogen to ammonia flow

rate was maintained. The detailed experimental conditions are summarized in Table 4-5. In

addition to the CO scrubber, a hydrogen cyanide (HCN) and an ammonia (NH3) scrubber were

placed at the outlet gas stream for safety reasons.

Table 4-5: Experimental conditions for the post-treatments of activated carbon

Adsorbent Temperature,

(oC)

Reaction

Time, (hr)

Gaseous

Environment

Flow Rate,

(cm3/min)

CO2-ADC-9h-pAir 250 12 Air (~21% O2) -

CO2-ADC-9h-pO2 240-260 ~6 97% O2

+3% H2O 130-150

CO2-ADC-9h-pNH3 850 2 20% NH3

+80% N2 100

42

4.2 Characterization of Activated Carbon

4.2.1 Materials and Instruments

The materials and instruments used for activated carbon characterization are listed in Table 4-6

and 4-7, respectively.

Table 4-6: Materials used for activated carbon characterization

Materials Specification Supplier

N2 (gas) 99.99% BOC Gas Ltd.

N2 (liquid) - MSE Dept, U of Toronto

KBr Analytical grade BDH Chemicals

NaOH 0.1M Standard solution Sigma-Aldrich

HCl 0.1M Standard solution Sigma-Aldrich

Phenolphthalein indicator 0.375% in methanol Sigma-Aldrich

NaHCO3 Analytical grade Sigma-Aldrich

Na2CO3 Analytical grade ACP Chemicals

Table 4-7: Instruments and equipment used for activated carbon characterization

Instrument Specification Manufacturer

SSA analyzer Autosorb 1 Quantachrome Instruments

Elemental analyzer CE-440 Exeter Analytical

FTIR spectrometer Spectrum-BX; Spectrum-One Perkin Elmer

Water bath shaker SW22 Julabo USA, Inc.

Weighing balance 1702, analytical balance Sartorius

FTIR pellet die - Chemistry Shop, U of Toronto

Plastic syringe Luer-Lok 10 mL BD Falcon

Syringe filter 0.45 µm polysulphone membrane Acrodisc, Pall Corp.

43

4.2.2 Pore Structure Analysis

The pore structure of activated carbon was analyzed by manipulating nitrogen adsorption and

desorption (physisorption) isotherms at 77K using Quantachrome Autosorb-1-C specific surface

area analyzer. These isotherms were constructed by plotting the volume of N2 adsorbed on the

carbon sample against the relative pressure, P/Po (the ratio of the adsorbate pressure in the

sample cell to the vapour pressure of the adsorbate). For this study, the total pore volume and the

pore size distribution were determined by using the quenched solid density functional theory

(QSDFT). The QSDFT method is one of the more accurate techniques for pore size distribution

of amorphous carbons because it takes into account for physical and chemical heterogeneity [51,

60]. All samples were outgassed at 200 oC for 2 hours to remove moisture and volatile organics

prior to analysis. Forty adsorption points between relative pressures (P/Po) of 0.025 and 0.995

were obtained from the instrument and analyzed using QSDFT, which was available as an

analysis kernel within the Quantachrome ASiQwin software. The analysis kernel assumed slit-

shaped micropores (<2 nm) and cylindrical mesopores.

The specific surface area (SSA) was determined using a three-point Brauner-Emmitt-Teller or

BET method, which applies the BET equation on the linear region of an adsorption curve. Some

researchers have reported that BET-derived SSA may not be very accurate for highly

microporous adsorbents [63]. For this reason, the SSA obtained in this study is referred to as

‘apparent’ BET SSA. The theories behind physisorption analytical methods are discussed

thoroughly by Sing et al., (1985) in an IUPAC report.

44

4.2.3 Surface Chemistry Analysis

4.2.3.1 FTIR Analysis

The chemical functional groups on the activated carbon surface were qualitatively analyzed

using Fourier-Transform Infrared (FTIR) spectroscopy. Prior to analysis, the activated carbon

samples were oven-dried at 110 oC for 18 hours to remove moisture content. FTIR spectra were

obtained within a frequency band range of 4000-400 cm-1

. Finely powdered potassium bromide

(KBr) was used as reference in a KBr/carbon pellet with a mass ratio of 200:1. The samples were

purged with dry air for 4 minutes in the instrument chamber prior to the runs. The Spectrum

5.0.1 software was used in conjunction with the instrument. The software parameters were set to

carry out 64 co-added scans at a resolution of 4 cm-1

. All samples except for CO2-ADC-9h-pNH3

were analyzed using a Perkin-Elmer Spectrum-BX spectrometer. CO2-ADC-9h-pNH3 was

analyzed using a Perkin-Elmer Spectrum-One spectrometer.

4.2.3.2 Boehm Titration

The surface oxygen groups of the post-oxidized activated carbon samples were quantitatively

determined by Boehm titration, a method pioneered by HP Boehm (1994). The method works on

the principle that the oxygen-based functional groups on the carbon surface have different

acidities and can be neutralized by bases of different strengths [25, 52]. A series of back

titrations was performed to determine the concentrations of carboxylic, lactonic, and phenolic

functional groups on activated carbon surfaces. The experimental details are described in the

following paragraph.

45

In separate flasks, approximately 300 mg of activated carbon samples were suspended in 25 mL

standard solutions of each NaOH, Na2CO3, and NaHCO3 (0.1 mol/L each). The solutions were

then shaken in a temperature-controlled water bath shaker for 24 hours at 25oC with a shaking

speed of 200 rpm. The spent adsorbents were removed from the solutions by centrifugation and

subsequent filtration. 50 mL of 0.1 mol/L HCl standard solution was added to 15 mL aliquot of

each of the three filtrates. The Na2CO3 and NaHCO3 filtrates were heated for about 5 minutes to

remove dissolved atmospheric CO2. A few (3-4) drops of phenolphthalein indicator solution

were added to each of the three filtrates. 0.1 mol/L of NaOH standard solution was used to titrate

each solution until the colour turned faint pink. The following equation was used to determine

the amount of acidic groups present on the carbon surface [25] –

where [B] and VB are the concentration and the volume of the reaction base mixed with the

carbon sample; nSFG (mmol/g) represents the amount of surface functional group that reacts with

the reaction base during the mixing step; m is the mass of carbon and Va is the volume of aliquot

taken from the filtrates;. [HCl] and VHCl are the concentration and the volume of HCl standard

solution; [NaOH] and VNaOH are the concentration of the volume of NaOH standard solution. The

molar ratio (nHCl/nB) accounts for the calculations of monoprotic vs. diprotic bases.

NaHCO3 can only neutralize carboxylic groups, where Na2CO3 can neutralize both lactonic and

carboxylic groups. NaOH, the strongest base of three, can neutralize all the acidic functional

groups (carboxylic, lactonic, and phenolic groups). Hence, the following set of equations can be

used to determine the surface concentrations of the three acidic functional groups-

46

4.2.3.3 Elemental Analysis

Elemental analysis of the activated carbon samples was carried out using Exeter Analytical CE-

440 elemental analyzer. CO2-ADC-9h and the post-oxidized samples were analyzed for their

elemental oxygen content. CO2-ADC-9h-pNH3 and CO2-ADC-9h were analyzed for their

nitrogen content to evaluate the effect of ammonia post-treatment.

Prior to the analyses, all samples were oven-dried at 110 oC for 12 hours to remove moisture

content. Standard samples were used to calibrate the instrument. The oxygen content was

measured by pyrolysis at high temperature in presence of platinized carbon [22]. Carbon dioxide

formed from the elemental oxygen was detected and measured by high-precision thermal

conductivity detector. For nitrogen analysis, the samples were combusted in pure oxygen at high

temperature. The nitrogen-based combustion products (N oxides and NO2) were then measured

by the detector, yielding the elemental nitrogen content [22]. Triplicate samples of each

adsorbent were run to ensure good precision. Due to the nature of detection, inorganic oxygen

groups were not accounted for. For this reason, the oxygen content determined using this method

is referred to as ‘organic oxygen’. It should be noted that the FTIR and the elemental analysis of

the adsorbents were generously carried out by other researchers from the Green Technology

Group (U of Toronto).

47

4.3 Adsorption Study

4.3.1 Materials and Instruments

The materials and the instruments used for the adsorption study are listed in Table 4-8 and 4-9,

respectively.

Table 4-8: Materials used for the adsorption study

Materials Specification Supplier

NaCl 99.99% Sigma-Aldrich

Na2SO4 Analytical grade ACP Chemicals

H2SO4 0.1M Standard solution Sigma-Aldrich

NaOH 0.1M Standard solution Sigma-Aldrich

CH2Cl2 (DCM) 99.5% Sigma-Aldrich

CaCl2 99.99% Sigma-Aldrich

MgCl2 > 99.0% Sigma-Aldrich

NaHCO3 Analytical grade Sigma-Aldrich

Tailings pond water pH ~8 Suncor Energy Inc.

Table 4-9: Instruments and equipment used for the adsorption study

Instrument Specification Manufacturer

FTIR spectrometer Spectrum-One; Spectrum-BX Perkin Elmer

FTIR liquid sample cell 5 mm KBr cell Pike Technologies

Water bath shaker SW22 Julabo USA, Inc.

Centrifuge tubes Polypropylene conical bottom BD Falcon

pH meter M-630 Fisher Inc.

Plastic syringe Luer-Lok 10 mL BD Falcon

Syringe filter 0.45 µm polysulphone membrane Acrodisc, Pall Corp.

48

4.3.2 Model Naphthenic Acid

A single-ring model naphthenic acid (trans-4-pentylcyclohexane carboxylic acid; MW = 198.30

g/mol; C12, z = -2) was used for majority of the kinetic and equilibrium study to understand the

fundamentals behind non-competitive adsorption process. Consequently, synthetic tailings pond

water (STPW) solution was prepared for the adsorption study by dissolving 25 mmol/L NaCl, 15

mmol/L NaHCO3, 2 mmol Na2SO4, 0.3 mmol/L CaCl2, and 0.3 mmol/L of MgCl2 in de-ionized

water. The pH of the solution was adjusted to 8.5 using NaOH (0.1M) to match the pH of real

tailings pond water. The chemical composition of STPW solution was directly adopted from

Kiran et al., (2011). An equilibrium adsorption study with real tailings pond water (provided by

Suncor Energy) was also performed to evaluate the adsorption capacity of activated carbon for

total acid-extractable organics (TAOs).

4.3.3 Quantification of Model Naphthenic Acid and Total Acid-extractable Organics

A Fourier-Transform infrared (FTIR) spectroscopy-based technique was utilized to quantify the

model naphthenic acid (mNA) in aqueous phase. The technique was developed by researchers at

Syncrude Canada Ltd., and the experimental details have been described in the open literature

[30, 39, 66, 81]. The two major steps – sample extraction from aqueous phase and FTIR

quantification – involving this method are described in the following sections.

4.3.3.1 Sample Extraction from Aqueous Phase

Prior to the FTIR analysis, the model naphthenic acid was extracted from its initial aqueous

phase to an organic phase. This was performed by liquid-liquid extraction with dichloromethane

(DCM) as the solvent. Prior to extraction, the mNA molecules were converted to their non-ionic

49

form by acidify the aqueous solution to a pH below 2 with concentrated H2SO4 solution. In

STPW solution, the molecules exist in their salt (ionic) form because the pH of STPW is higher

than the pKa of the mNA (~5.2). Therefore, extraction without acidification would lead to loss of

mNA molecules to the aqueous layer. Using a separatory funnel, each acidified sample was

extracted twice with DCM in a 2:1 by volume of sample to solvent ratio. The combined DCM

extracts containing the naphthenic acid was evaporated to dryness under air flow. The residue

was reconstituted with 6 mL of DCM. This concentrated solution of mNA in DCM was then

transferred into a 20 mL glass vial and stored at 0 oC until the FTIR analysis.

4.3.3.2 FTIR-based Quantification

The majority of the FTIR quantification of model naphthenic acid was performed using a Perkin

Elmer Spectrum-BX spectrometer with a sealed KBr liquid sample cell (pathlength = 5 mm). Due

to a malfunction of the BX spectrometer, a second Perkin Elmer spectrometer (Spectrum-One)

was utilized for the later segment of the study. The conditions and the software for the analyses

were kept the same for the both instruments for consistency.

Both instruments were operated in the absorbance mode. The spectra were acquired within a

frequency band range of 4000-400 cm-1

and processed using the Spectrum 5.0.1 software. The

analysis chamber was purged with dry air for ~ 4 minutes before each run. The background

spectrum of dichloromethane (DCM) was obtained prior to sample analysis. For each spectrum

(including the background), 64 co-added scans were performed. The resolution of the instrument

was set to 4 cm-1

.

50

Carboxylic acids have signature C=O absorbance peaks in the wavelength regions of 1740-1750

cm-1

and 1700-1715 cm-1

, representing monomer and dimer absorption respectively [10]. The

absorbance peak heights at 1743 cm-1

(monomer) and 1704 cm-1

(dimer) were measured and

summed. The baseline used for peak height measurement was from 1850 to 1660 cm-1

. The

concentration of an unknown mNA sample was determined by comparing its combined peak

heights (monomer and dimer) to a calibration curve. The calibration curve was prepared as

follows: a stock solution of model naphthenic acid was prepared in DCM. The aliquots of this

solution were diluted with DCM to yield calibration standards with concentrations ranging

between 0.5 to 300 ppm mNA to a final volume of 6 mL. The calibration standards were

analyzed using the FTIR method described above. Their combined peak heights (at 1743 and

1704 cm-1

) were then plotted against the known concentrations to construct the calibration curve.

Two calibration curves were used in this study, corresponding to the two spectrometers. The

detection limit of the spectrometers was experimentally determined to be approximately 1.0 ppm.

These calibration curves can be found in Appendix A.

Control tests were performed to determine the mNA recovery from aqueous phase by liquid-

liquid extraction. The mNA recovery was calculated using equation 4.6. During the adsorption

study, the aqueous mNA samples were syringe-filtered with membrane filters to remove spent

activated carbon. This step was accounted for during the recovery control tests.

The recovery efficiency was determined to be 84.3 ± 2.6 % based on a total of 5 samples. The

loss of the model naphthenic acid might be attributed to the following factors: losses at the

51

interfacial emulsion layer between the organic and the aqueous phase during extraction, losses

due to micro-filtration with membrane filters, and losses due to evaporation of mNA molecules

during the solvent-drying step. The recovery efficiency was plotted against the initial sample

concentrations (Appendix A). The result suggests that the recovery was likely independent of the

initial concentrations. To account for the losses, a correction factor (~15%) was incorporated into

all adsorption calculations.

Total acid-extractable organics or TAOs (complex naphthenic acid mixtures) were also

quantified using the FTIR method described above. The same calibration curve as above was

used to determine the concentrations of TAOs. This is justifiable because the FTIR quantification

of carboxylic acids is solely based on the infrared absorbance of the CO double bonds, which is

proportional to the concentration of total carboxylic acids in a sample, regardless of the type of

acids.

4.3.4 Batch-type Adsorption Kinetics Study

For the kinetic experiments, stock solution of the model naphthenic acid was prepared in

synthetic tailings pond water (STPW). 40 mL of the solution was introduced into a series of 50

mL polypropylene centrifuge vials, which were utilized as batch-type reactors. 10 mg of

activated carbon was added to each of these vials. The vials were placed in a temperature-

controlled water bath shaker at 200 rpm. This was experimentally determined to be the shaking

speed at which the mass transfer resistance due to external diffusion was minimized (See Section

5.2.1 for details). All of the kinetic experiments were conducted at 25 oC except during the

‘effect of temperature’ study. At pre-set time intervals, the sample vials were removed from the

shaker. The spent adsorbent particles were removed by syringe filtration using polysulphone

52

membrane filters (pore diameter = 0.45 µm). The additional time required to remove adsorbent

from the solutions was accounted for in the kinetic calculations. The concentrations of the

filtered sample solutions were determined by the FTIR technique described in the previous

section. For a given experiment, each sample would serve as a data point on a concentration

versus time graph. Figure 4-4 illustrates a schematic flow diagram of the experimental details of

the adsorption study.

4.3.5 Adsorption Equilibrium (Isotherm) Study

The equilibrium adsorption study had a similar experimental protocol to that of the kinetic study.

Stock solutions of desired mNA concentrations were prepared in STPW. The data points for

adsorption isotherms were obtained by fixing the initial mNA concentration of each sample

constant while varying the adsorbent dosage (ranging between 2-30 mg). 40 mL of stock solution

was placed into a series of 50 mL polypropylene centrifuge vials. Adsorbent dosage of

increasing mass (2-30 mg) was then introduced into each sample vial. The vials were placed in

the water bath shaker at 25 oC, and the adsorption process was allowed to reach equilibrium. The

equilibration time was determined from preliminary experiments: 18 hours for commercially

activated carbon (CAC) and 20 hours for activated delayed coke (ADC). Spent adsorbents were

removed by syringe filtration, and the concentration of the remaining solution was determined by

the FTIR method.

53

Figure 4-4: Experimental flow diagram of the adsorption study

An equilibrium study was also performed with real tailings pond water provided by Suncor

Energy to investigate the equilibrium adsorption of total acid-extractable organics (TAOs). The

pond water was syringe-filtered to remove larger solid particles prior to the adsorption. The

experiment was carried out in a similar manner as described above. The amount of solute

adsorbed onto the adsorbent surface at equilibrium (qe) was calculated using the following

equation-

54

where qe is the amount of the mNA or TAOs adsorbed at equilibrium (mg/g); Co and Ce are the

initial and equilibrium aqueous concentration of the adsorbate (mg/L), respectively; V is the

volume of the aqueous sample (L); and m is the mass of adsorbent dosage (g).

4.3.6 Adsorption Data Quality

To ensure good repeatability of the adsorption data, a few kinetic and equilibrium experiments

were repeated. The concentration values from the replicate runs, the standard deviation and the

% relative standard deviation are summarized in this section. Table 4-10 displays these values

for a kinetic study that was performed to investigate the effects of surface functionalities on the

mNA adsorption using CO2-ADC-9h.

Table 4-10: Adsorption kinetics data quality for “effect of surface functional groups” study

Aqueous Phase mNA Concentration, Ct (mg/L)

Time

(min) Run 1 Run 2 Run 3 Average

Standard

Deviation % R.S.D

9 27.0 24.1 25.4 25.5 1.2 4.56

24 21.4 21.9 22.1 21.8 0.32 1.47

48 19.3 18.9 18.7 18.9 0.26 1.36

*Conditions: Solution volume = 40 mL; Solution pH = 8.5; Temperature = 25oC; Shaking speed = 200

rpm; Initial mNA concentration = 33mg/L; CO2-ADC-9h dosage = 10 mg

To obtain thermodynamic parameters for adsorption, an equilibrium adsorption study was carried

out to investigate the effect of temperature on the model naphthenic acid adsorption. For each

temperature value, three runs were performed to determine the equilibrium solution

55

concentration of the model naphthenic acid. Table 4-11 lists the concentration values from the

replicate runs, the standard deviation and the % relative standard deviation.

Table 4-11: Equilibrium adsorption data quality

Equilibrium Solution Concentration of mNA

(mg/L)

Temp., oC

Run 1 Run 2 Run 3 Average Standard

Deviation % R.S.D

22 11.2 9.0 12.5 10.9 1.4 13.3

40 13.0 12.3 12.3 12.6 0.33 2.62

60 20.4 19.2 19.2 19.6 0.57 2.91

* Conditions: Solution volume = 40 mL; pH = 8.5; Shaking speed = 200 rpm.

Initial mNA concentration (Co) = 51 mg/L; CO2-ADC-6h dosage = 30 mg.

Triplicate samples were run for the first FTIR calibration curve and duplicate samples were run

for the second calibration curve. The FTIR absorbance data along with the standard deviation for

each run are summarized in Appendix A.

56

CHAPTER 5 RESULTS AND DISCUSSION

5.1 Characterization of Activated Carbon

5.1.1 Pore Structure Analysis

The physical properties of the activated carbon adsorbents were investigated by porous structure

analysis. The graphical representations of N2 adsorption isotherms for the adsorbents can be

found in Appendix B. The physical pore properties deduced from the isotherm data are

summarized in Table 5-1.

Table 5-1: Physical pore properties of activated carbon adsorbents

Adsorbent

Apparent

SSABET SSAMESO VMIC VMESO VTOT

VF, MIC

VF,MESO

(m2/g) (m

2/g) (cm

3/g) (cm

3/g) (cm

3/g)

Steam-ADC-3h 408 85 0.135 0.046 0.181 0.75 0.25

CO2-ADC-6h 276 92 0.080 0.083 0.163 0.49 0.51

CO2-ADC-9h 405 134 0.100 0.202 0.302 0.33 0.67

CO2-ADC-9h-pAir 295 100 0.073 0.185 0.258 0.28 0.72

CO2-ADC-9h-pO2 300 116 0.067 0.212 0.279 0.24 0.76

CO2-ADC-9h-pNH3 305 130 0.077 0.136 0.213 0.36 0.64

CAC 1120 213 0.308 0.230 0.538 0.57 0.43

* SSA = Specific surface area; SSAMESO = Mesopore SSA; VMIC = Micropore volume; VMESO =

Mesopore volume; VTOT = Total pore volume, VF = Volume fraction.

57

As listed in Table 5-1, upon increasing the activation time from 6 to 9 hours, most of the porous

properties of the CO2-activated delayed coke increased. Both the microporosity and

mesoporosity increased with activation time, resulting in higher total pore volume and SSA. Guo

et al., (2009) observed a similar trend for CO2 activation of coconut shell-based AC precursor.

The specific surface area generally increases with activation time as the porous structure is

developed further. However, at a given temperature, there exist an optimal activation time for a

maximum SSA, beyond which the SSA will start to decrease [27, 49]. This is because prolonged

activation time leads to the enlargement of pre-existing micropores and possible destruction of

mesopores. This would result in a lower SSA and total pore volume [27].

Although Steam-ADC-3h and CO2-ADC-9h have similar apparent BET SSA, their pore size

distributions are considerably different. The mesoporous volume fractions of Steam-ADC-3h and

CO2-ADC-9h were 0.25 and 0.67, respectively. The microporous volume fraction of Steam-

ADC-3h was 0.75 whereas for CO2-ADC-9h, it was only 0.33. This may be due to the fact that

water molecules have a smaller dimension than carbon dioxide molecules, and the reaction

between the water molecules and the carbon skeleton creates smaller (micro) pores comparing to

the reaction between CO2 and carbon [54].

All the post-modified CO2-ADC-9h samples exhibit lower specific surface area and lower total

pore volume than the parent adsorbent. According to Awoyemi (2011), surface oxidation of

activated carbon could have both positive and negative effects on the pore structure. The reaction

between oxygen and the carbon skeleton could result in the formation of new micropores,

thereby increasing the SSA. However, the same reaction could lead to the enlargement and

possible destruction of pre-existing pores, and consequently decreasing the total SSA and pore

58

volume [4]. Pre-existing pores can also be blocked by newly formed oxygen-based surface

functional groups. This would similarly reduce the SSA and the pore volume. For both the

surface-oxidized adsorbents (CO2-ADC-9h-pAir and CO2-ADC-9h-pO2), a decrease in the SSA

and the total pore volume is observed.

Similar trends are observed for ammonia post-treated activated carbon. The decrease in the SSA

and the total pore volume of the sample are presumably due to enlargement of pre-existing

micropores and mesopores. Upon the ammonia post-treatment, the micropore and the mesopore

volume of the CO2-ADC-9h decreased from 0.100 to 0.077 cm3/g and from 0.202 to 0.136

cm3/g, respectively. Nitrogen-based functional groups may contribute to this effect by blocking

the pores. The commercially activated carbon (CAC) displayed the highest SSA (1120 m2/g) and

the total pore volume (0.538 cm3/g) among all the adsorbents.

5.1.2 Surface Chemistry Analysis

5.1.2.1 FTIR Analysis

The surface functional groups of the petroleum coke-derived activated carbon samples were

qualitatively analyzed by solid-state FTIR spectroscopy. The FTIR spectra of all the petroleum

coke-derived activated carbons (with the exception of CO2-ADC-9h-pNH3) are depicted in

Figure 5-1. The FTIR spectrum of ammonia post-treated activated carbon (CO2-ADC-9h-pNH3)

is illustrated in Figure 5-2.

59

Figure 5-1: FTIR spectra of petroleum coke-derived activated carbon adsorbents

All the FTIR spectra in Figure 5-1 show transmittance peaks in very similar regions. The broad

peak between 3600-3150 cm-1

is attributed to the O−H stretching vibration, which suggests the

presence of hydroxyl, phenol, and carboxyl groups on the carbon surface [55]. The peak centred

around 1750-1600 cm-1

is most likely due to the oscillation of C=O bonds associated with

carbonyl, lactonic and carboxylic groups. The sharp peak between 1450-1370 cm-1

may be

caused by the bending vibrations of –CH3 bonds, which suggests the presence of methyl groups

on the adsorbent surface [4]. The peak around 1300-1100 cm-1

is associated with the stretching

vibration of C−O bonds of lactonic, carboxylic and anhydride groups. The weak bands between

900-550 cm-1

are due to the out-of-plane bending vibrations of =(C−H) bonds of aromatic rings

[55].

45

50

55

60

65

70

01000200030004000

Tra

nsm

itta

nce

, (%

)

Wavenumber, (cm-1)

Steam-ADC-3h

CO2-ADC-6h

CO2-ADC-9h

CO2-ADC-9h-pAir

CO2-ADC-9h-pO2

60

In summary, the spectral analysis of Figure 5-1 suggests the presence of oxygen-based surface

functionalities such as carboxylic, lactonic, and phenolic groups on the activated carbon surface.

The FTIR spectrum in Figure 5-2 illustrates the effect of gaseous ammonia post-treatment on

CO2-activated carbon. According to Przepiorski et al. (2004), ammonia decomposes at high

temperature to form radicals such as NH2, NH and H, which could react with the carbon surface

to form nitrogen-based functional groups such as –NH2, −CN, pyridinic, pyrrolic and quaternary

nitrogen [58].

Figure 5-2: FTIR spectrum of ammonia post-treated CO2-activated carbon

In Figure 5-2, the peak between 2400-2250 cm-1

suggests the infrared absorption of nitrile

(−C≡N) groups. The presence of C−N stretching vibrations (1350-1000 cm-1

) and the N−H

bending vibrations (~1550 cm-1

), coupled with C=O stretches (1615 cm-1

) indicate the formation

of amide functional groups on the adsorbent surface. Przepiorski et al., (2004) also reported the

presence of amide groups on the surface of gaseous ammonia-treated activated carbon. The

59.5

60

60.5

61

61.5

62

05001000150020002500300035004000

Tra

nsm

itta

nce

, (%

)

Wavenumber, (cm-1)

61

stretching vibration of N−H bonds (~3400 cm-1

) and the out-of-plane bending vibrations of N−H

bonds (~800 cm-1

) in combination with C−N stretching vibrations (1350-1000 cm-1

) suggest the

presence of amine groups [55]. The peak around 1690-1640 cm-1

implies the presence of

possible imine groups. In summary, the spectral analysis of the ammonia post-treated activated

carbon strongly suggests the development of nitrogen-based surface functionalities.

5.1.2.2 Boehm Titration

The results obtained from the Boehm titrations are graphically represented in Figure 5-2, and

numerically summarized in Table 5-2. Titrations were performed with the CO2-ADC-9h, CO2-

ADC-9h-pAir, and CO2-ADC-9h-pO2 samples to investigate the effects of post-oxidation.

Figure 5-3: Amount of surface oxygen groups on activated carbon adsorbents

The Boehm titration results were consistent with the findings from the FTIR analysis, verifying

the presence of phenolic, carboxylic, and lactonic (mmol/g of adsorbent) groups on activated

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

CO2-ADC-9h CO2-ADC-9h-pAir CO2-ADC-9h-pO2

Am

ou

nt

of

Su

rface

Oxygen

Gro

up

s, (

mm

ol/

g)

Activated Carbon Adsorbents

Phenolic

Carboxylic

Lactonic

Total

62

carbon surface. The results show that the effect of surface oxidation was most pronounced for

humidified-oxygen treated activated carbon. The total amount of surface oxygen groups was the

highest for CO2-ADC-9h-pO2, followed by CO2-ADC-9h-pAir and CO2-ADC-9h. This indicates

that humidified-oxygen is more effective in oxidizing adsorbent surface than humidified-air. This

was expected because the humidified-oxygen stream had a higher oxygen composition (97% O2

+ 3% H2O) than humidified-air (~21% O2). Figure 5-3 shows that the amount of carboxylic and

lactonic groups on the adsorbents increased with increasing degree of oxidation (CO2-ADC-9h <

CO2-ADC-9h-pAir < CO2-ADC-9h-pO2).

Table 5-2: Amount of surface oxygen groups on activated carbon adsorbents

Adsorbent

Acidic Surface Oxygen Groups (mmol/g)

Carboxylic Lactonic Phenolic Total

CO2-ADC-9h 0.020 0.472 0.136 0.62

CO2-ADC-9h-pAir 0.061 0.906 0.088 1.05

CO2-ADC-9h-pO2 1.416 1.433 0.226 3.07

5.1.2.3 Elemental Analysis

Elemental analysis was performed on CO2-ADC-9h and the two post-oxidized AC samples

(CO2-ADC-9h-pAir and CO2-ADC-9h-pO2) to determine their elemental oxygen content. The

results are graphically presented in Figure 5-4, and summarized in Table 5-3. The oxygen

content obtained from the Exeter Analytical CE-440 elemental analyzer is referred to as

‘organic’ oxygen because the method of detection does not account for inorganic oxygen species.

63

The nitrogen content of CO2-ADC-9h and CO2-ADC-9h-pNH3 were obtained to evaluate the

effects of gaseous ammonia post-treatment. These results are tabulated in Table 5-3.

Figure 5-4: Elemental ‘organic’ oxygen content of activated carbon adsorbents

Figure 5-4 displays the following trend in the ‘organic’ oxygen content of the adsorbents: CO2-

ADC-9h-pO2 > CO2-ADC-9h-pAir > CO2-ADC-9h. It is apparent that the oxygen content of the

adsorbents increased with increasing degree of surface oxidation. These results are consistent

with the Boehm titration data. This further supports the conclusion that the humidified-oxygen

treatment was more effective at oxidizing activated carbon than humidified-air.

As listed in Table 5-3, the nitrogen content of activated carbon increased by 10% (from 1.11 to

1.23 wt%) after post-treatment with gaseous ammonia at 850 oC for 2 hours. This observation is

in good accordance with the qualitative FTIR analysis of CO2-ADC-9h-pNH3, which indicated

the development of nitrogen-based surface functionalities.

1.5 3.5

10.1

0

2

4

6

8

10

12

CO2-ADC-9h CO2-ADC-9h-pAir CO2-ADC-9h-pO2

Ele

men

tal

'Org

an

ic'

Ox

ygen

Co

nte

nt,

(%

)

Activated Carbon Adsorbents

64

Table 5-3: Elemental analysis of activated carbon

Adsorbent Oxygen (wt%) Nitrogen (wt%)

CO2-ADC-9h 1.53 ± 0.26 1.11 ± 0.02

CO2-ADC-9h-pAir 3.50 ± 0.39 n/m

CO2-ADC-9h-pO2 10.1 ± 0.27 n/m

CO2-ADC-9h-pNH3 n/m 1.23 ± 0.23

* n/m = not measured.

65

5.2 ADSORPTION STUDY

5.2.1 Adsorption Kinetics

One of the main objectives of the kinetic study was to investigate the effects of adsorbent and

solution properties on adsorption rate parameters. Kinetic models that best describe the

experimental mNA adsorption data were identified. Experiments were designed and performed

to answer some of the following research questions-

What is the effect of shaking speed? At what shaking speed would the mass transfer

resistance due to external diffusion be minimized?

Which kinetic model best fits the experimental data? What are the implications?

How do the adsorbent pore structure and surface functionalities affect rate parameters?

What is the rate-limiting step in the adsorption process?

What are the effects of initial adsorbate concentration and temperature?

The first step of an adsorption process in a solid-liquid system is the external diffusion of

adsorbate molecules from bulk liquid phase through a stagnant film layer to the external surface

of the adsorbent (Figure 3-1). The shaking or the mixing speed within a system has a

considerable impact on the external diffusion of adsorbate species [67, 84]. Figure 5-5 illustrates

the effect of shaking speed on the model naphthenic acid adsorption (% adsorbed) from aqueous

phase. After 40 minutes, without shaking, the mNA adsorption was quite low for CO2-ADC-6h

(8%). However, the figure shows an increase in adsorption with increasing shaking speed until a

plateau is reached around 175-200 rpm. Increasing the shaking speed reduces the film

66

(boundary) layer thickness. In other words, the mass transfer resistance due to external diffusion

is reduced by shaking. Consequently, the adsorbate molecules can diffuse through the film layer

easily and their adsorbed quantity (%) increases.

Figure 5-5: The effect of shaking speed on the model naphthenic acid adsorption

The mNA adsorption does not increase any further beyond the shaking speed of 200 rpm. At this

point, the adsorption process is most likely controlled by either internal pore diffusion or surface

adsorption, and the mNA adsorption will not increase any further as the shaking speed rises.

Therefore, it is reasonable to assume that the mass transfer resistance due to external diffusion is

negligible at this shaking speed. The subsequent kinetic and equilibrium experiments were all

performed at 200 rpm.

0

20

40

60

80

100

0 50 100 150 200 250

Per

cen

tage

Ad

sorb

ed, (%

)

Shaking Speed, (rpm)

CO2-ADC-6h

Conditions: Agitation duration = 40 mins; Solution volume = 40 mL

Initial mNA concentration ~48 mg/L; AC dosage ~20 mg

67

5.2.1.1 Effect of Pore Structure

To investigate the effect of adsorbent pore structure on the model naphthenic acid adsorption

kinetics, three adsorbents (CO2-ADC-6h, Steam-ADC-3h, and CAC) with contrastingly different

porosity were used. Figure 5-6 illustrates the time-dependent adsorption of the mNA onto the

three adsorbents. The raw adsorption data for this study can be found in Appendix C, along with

kinetic model fitting graphs.

Figure 5-6: Time-dependent adsorption of mNA onto adsorbents of different porosity

As illustrated in Figure 5-6, CAC had the highest amount of mNA adsorbed after ~ 95 minutes,

followed by CO2-ADC-6h and Steam-ADC-3h (~270, 200, and 160 mg/g, respectively). To fully

understand the effect of pore structure on mNA adsorption kinetics, the kinetic data was

0

50

100

150

200

250

300

0 20 40 60 80 100

Am

ou

nt

Ad

sorb

ed, q

t (m

g/g

)

Time, (min)

CO2-ADC-6h CO2-ADC-6h pseudo-2nd order

Steam-ADC-3h Steam-ADC-3h pseudo-2nd oder

CAC CAC pseudo-2nd order

Conditions: Solution volume = 40 mL; pH = 8.5; Temperature = 25oC.

Shaking speed = 200 rpm.

Initial mNA concentration ~92mg/L; Adsorbent dosage ~10 mg.

68

formulated into several kinetic models, and the model parameters such as the rate constant and

the equilibrium adsorbed quantity were obtained. These parameters were then used as a basis for

comparing the performance of the different adsorbents. More specifically, the kinetic parameters

were correlated with adsorbent pore properties to understand the effect of pore structure. Table

5-4 summarizes the model parameters for the pseudo-first order, pseudo-second order, and

Elovich models and the correlation coefficients.

From the correlation coefficients of the models in Table 5-4, it is evident that the pseudo-second

order rate model exhibits the best fit for the experimental data for all three adsorbents. The

model-predicted adsorbed quantities are depicted by the dashed lines in Figure 5-6. The dashed-

fits indicate a good correlation between the model and the experimental data. Previously, Zubot

(2010) demonstrated that the adsorption kinetics of naphthenic acid mixture onto Syncrude

petroleum coke also follow the pseudo-second order model.

Thereby, the pseudo-second order rate parameters were used to investigate the effect of pore

structure on adsorption kinetics. From here on, this will be the only empirical model used to

report kinetic data.

69

Table 5-4: Empirical kinetic model parameters from the ‘Effect of Pore Structure’ study

Activated Carbon Adsorbents

Kinetic Models CAC CO2-ADC-6h Steam-ADC-3h

Pseudo-2nd Order

k2 (g/mg min) 5.62E-04 9.68E-04 8.14E-05

qe,2 (mg/g) 286 208 238

r2 0.999 0.998 0.975

Pseudo-1st Order

k1 (min-1

) 9.67E-03 2.05E-02 2.46E-02

qe,1 (mg/g) 165 113 180

qe,exp (mg/g) 339 221 174

r2 0.624 0.765 0.975

Elovich Model

β (g/mg) 2.40E-02 4.36E-02 1.97E-02

α (mg/g min) 290 1.5E+03 10

r2 0.855 0.931 0.974

Table 5-5 compares the adsorbent pore properties with the pseudo-2nd

order rate parameters. No

distinct trend is observed between the apparent BET SSA and the pseudo-second order rate

constant (k2). This is also true for the case of the total pore volume and the rate constant. CO2-

ADC-6h exhibits the highest rate constant even though it has the smallest apparent BET SSA.

However, among the three adsorbents, CO2-ADC-6h has the highest mesoporous volume

fraction, followed by CAC and Steam-ADC-3h. A careful examination of the mesoporous

volume fraction and the rate constant reveals a positive correlation between the two.

70

Table 5-5: Correlations between adsorbent pore properties and pseudo-2nd

order rate parameters

Adsorbent Porous Properties

Pseudo-second Order

Rate Parameters

Adsorbent

Apparent

SSABET,

(m2/g)

Total

Pore

Volume

(cm3/g)

Mesopore

Volume

Fraction

Micropore

Volume

Fraction

Rate

constant,

k2 (g/mg

min)

Equilibrium

Adsorbed

Amount, qe

(mg/g)

CO2-ADC-6h 276 0.163 0.51 0.49 9.68E-04 208

Steam-ADC-3h 408 0.181 0.25 0.75 8.14E-05 238

CAC 1120 0.538 0.47 0.53 5.62E-04 286

Figure 5-7 further illustrates the positive correlation between the rate constant and the

mesoporous volume fraction of the adsorbents.

Figure 5-7: Correlation between mesoporous volume fraction and pseudo-2nd

order rate constant

This observation may be rationalized by the following set of arguments. Mesopore diameters are

between 2 and 50 nm while micropore diameters are below 2 nm. The model naphthenic acid

(trans-4-pentylcyclohexane carboxylic acid) has the following approximate dimensions (Figure

5-8).

0.0E+00

3.0E-04

6.0E-04

9.0E-04

1.2E-03

0.20 0.30 0.40 0.50 0.60

Pse

ud

o-2

nd O

rder

Rate

Con

stan

t, k

2 (

g/m

g m

in)

Mesoporous Volume Fraction

CO2-ADC-6h

Steam-ADC-3h

CAC

71

Figure 5-8: Molecular dimensions of the model naphthenic acid [9]

It is apparent that the mNA molecules will not be able to diffuse into micropores with diameter

smaller than ~0.25 nm due to physical constraint. Even for micropores with diameters greater

than 0.25 nm, it would be rather difficult for the molecules to diffuse into these pores because

their steric orientation has to be perfectly aligned with the pore geometry. In contrast, the

diffusion and subsequent adsorption into the mesopores would be conceivably easier due to the

wider pore diameter. For this reason, the adsorbent with highest mesoporous volume fraction

(CO2-ADC-6h) exhibits the highest rate constant, and the adsorbent with the highest

microporous volume fraction (Steam-ADC-3h) exhibits the lowest rate constant.

0.22 nm

1.14 nm

0.25 nm

72

5.2.1.2 Effect of Initial Adsorbate Concentration

The effect of initial adsorbate concentration on the adsorption kinetics was investigated using

different initial concentrations of model naphthenic acid solutions (31, 51, and 175 mg/L). CO2-

ADC-6h was used as the adsorbent for this study. The time-dependent adsorption plots and the

pseudo-2nd

order fits to the experimental data are depicted in Figure 5-9.

Figure 5-9: Time-dependent adsorption on mNA on CO2-ADC-6h with different initial mNA

concentrations

The above figure shows that the mNA adsorption onto CO2-ADC-6h increases with initial mNA

concentration. The initial rate of adsorption, ho, is defined as the adsorption rate as time

0

20

40

60

80

100

120

0 15 30 45 60 75 90 105 120 135Am

ou

nt

Ad

sorb

ed, q

t (m

g/g

)

Time, (min)

31 mg/L 31 mg/L pseudo-2nd order

51 mg/L 51 mg/L pseudo-2nd order

175 mg/L 175 mg/L pseudo-2nd order

Conditions: Solution volume = 40 mL; pH = 8.5; Temperature = 25oC.

Shaking speed = 200 rpm; CO2-ADC-6h dosage ~10 mg.

73

approaches zero. The correlation between the initial rate of adsorption and the initial mNA

concentration is listed in Table 5-6 and depicted in Figure 5-10.

Table 5-6: Effect of initial mNA concentration on initial rate of adsorption

Initial mNA Concentration

(mg/L)

Initial Rate of Adsorption,

ho (mg/g min) r

2

31 1.84 0.946

51 4.30 0.989

175 8.98 0.988

The positive correlation between the initial rate of adsorption and the initial mNA concentration

has a certain implication in regards to the rate-controlling step of the adsorption process.

Figure 5-10: Effect of initial mNA concentration on initial rate of adsorption

The diffusional mass transfer of a solute species between two boundaries (A and B) can be

represented by the following form of Fick’s Law of diffusion [46]-

1.84

4.30

8.98

0

2

4

6

8

10

31 51 175Init

ial

Rate

of

Ad

sorp

tion

,

(m

g/g

min

)

Initial mNA Concentration, (mg/L)

74

where N is the mass flux of the solute; km is mass transfer coefficient related to the diffusivity

coefficient; CA and CB are the concentrations of the solute at boundaries A and B, respectively.

This concept could be readily applied for the case of the model naphthenic acid (mNA) diffusion

from the exterior surface of an activated carbon particle to the active sites through the internal

pores. Since the effects of external diffusion were minimized by appropriate shaking, the

concentration gradient between the bulk liquid and the exterior surface of the adsorbent is

practically zero. Hence, CA, in this case, is Cbulk, which is the initial mNA concentration in the

bulk liquid phase. In this scenario, CB, is the adsorbate concentration on the active sites.

At time = 0, CB would essentially be zero. Thereby, the rate of mass transfer through the internal

pores becomes the initial rate of mass transfer (adsorption). At this point, equation 5.1 takes the

following form-

where ho is the previously defined initial rate of adsorption (mg/g min), and Cbulk (mg/L) is the

initial bulk phase mNA concentration. A diffusion-controlled process should display a similar

positive correlation between ho and Cbulk.

The results summarized in Table 5-6 and Figure 5-10 display a positive correlation between the

initial rate of adsorption and the initial mNA concentration. This suggests that the adsorption of

the model naphthenic acid is presumably an intraparticle diffusion-controlled process. The

positive correlation between the initial rate of adsorption and the initial mNA concentration is

not exactly linear. This is because the interfacial surface area that the adsorbate molecules have

75

to diffuse through is not constant; rather it varies with pore diameter. Consequently, the positive

correlation is not perfectly linear.

5.2.1.3 Effect of Temperature

The effect of temperature on the mNA adsorption kinetics was investigated using CO2-ADC-9h.

The graphical representation of the experimental data (Figure 5-14) shows a slight increase in

mNA adsorption with temperature, but the results are not very conclusive. However, Table 5-9

shows that the pseudo-2nd

order rate constant increases with increasing temperature.

Figure 5-11: Effect of temperature on time-dependent mNA adsorption

An Arrhenius plot was generated using the linearized form of equation 5-3 [46]. Ea is the

activation energy of the adsorption process; R is the universal gas constant (8.3145 J/mol K) and

T is the temperature. The pseudo-second order rate constant was used for k value of equation.

0

15

30

45

60

75

0 10 20 30 40 50 60

Ad

sorb

ed A

mou

nt,

qt (m

g/g

)

Time, (min)

25 degrees C 50 degrees C 75 degrees C

Conditions: Solution volume = 40 mL; pH = 8.5; Shaking speed = 200 rpm.

CO2-ADC-9h dosage ~10 mg; Initial mNA concentration ~ 51 mg/L

76

Table 5-7: Effect of temperature on pseudo-second order rate constant

Temperature

(oC)

Pseudo-2nd

Order Rate

Constant, k2 (g/mg min) r

2

25 5.51E-04 0.996

50 7.12E-04 0.995

75 8.63E-04 0.991

Figure 5-12 displays the Arrhenius plot of the model naphthenic acid adsorption. The apparent

activation energy of the adsorption process (Ea,app) was obtained from the slope of the Arrhenius

plot, and the pre-exponential factor (A) was obtained from the y-intercept. The values of Ea,app

and A were calculated to be ~7.7 kJ/mol and 0.013, respectively. The magnitude of activation

energy often gives an idea about an adsorption process. Diffusion-controlled kinetic processes

are typically known to have relatively small activation energy [64]. In contrast, high values of

activation energy generally suggest surface adsorption-controlled adsorption process [12, 64].

Figure 5-12: Arrhenius plot of the mNA adsorption

y = -930.11x - 4.38

R² = 0.998

-7.6

-7.4

-7.2

-7

0.0028 0.003 0.0032 0.0034

ln (

k2)

1/T, (K-1)

77

The relatively small magnitude of the apparent activation energy further supports the argument

that the adsorption process of the model naphthenic acid is most likely intraparticle diffusion-

controlled.

5.2.1.4 Effect of Surface Functional Groups

Three adsorbents – CO2-ADC-9h, CO2-ADC-9h-pAir and CO2-ADC-9h-pO2 – were used to

investigate the effect of surface oxygen groups on the adsorption of the model naphthenic acid

from synthetic tailings pond water. Figure 5-9 displays the time-dependent adsorption of mNA

onto the surface-modified activated carbons.

Figure 5-13: Time-dependent adsorption of mNA onto adsorbents of different surface

functionalities

0

20

40

60

80

100

0 20 40 60 80 100 120 140

Am

ou

nt

Ad

sorb

ed, q

t (m

g/g

)

Time, (min)

CO2-ADC-9h CO2-ADC-9h-pAir CO2-ADC-9h-pO2

Conditions: Solution volume = 40 mL; pH = 8.5; Temperature = 25oC.

Shaking speed = 200 rpm; Initial mNA concentration ~33 mg/L.

Adsorbent dosage ~10 mg.

78

Figure 5-13 shows that after about 125 minutes, CO2-ADC-9h had the highest mNA adsorption,

followed by CO2-ADC-9h-pAir and CO2-ADC-9h-pO2, respectively. This trend suggests that

activated carbon’s affinity for the model naphthenic acid decreases with increasing degree of

surface oxidation.

The Weber-Morris (W-M) intraparticle diffusion model was used to interpret the experimental

data from this study. The W-M intraparticle diffusion plots are depicted in Figure 5-14.

Figure 5-14: Weber-Morris intraparticle diffusion plots of mNA adsorption onto surface-

oxidized activated carbons

The intraparticle diffusion rate constant, kid, is the slope of the linear portion of the Weber-

Morris plot. If the extrapolation of the straight line passes through the origin (i.e., y-intercept =

0), then the adsorption process is said to be solely intraparticle (internal) diffusion-controlled

[79]. However, many researchers have reported double-natured W-M plots, displaying a curved

initial portion followed by a linear portion [59, 79]. The initial portion represents the film layer

0

20

40

60

80

100

0 2 4 6 8 10 12Qu

an

tity

Ad

sorb

ed, q

t (m

g/g

)

Time0.5, (min0.5)

CO2-ADC-9h CO2-ADC-9h-pAir CO2-ADC-9h-pO2

y = 3.06 x + 18

y = 1.23 x + 18

y = 7.43 x + 8.9

79

effects due to external diffusion, while the linear portion represents the intraparticle or internal

pore diffusion. The y-intercept of the linear portion, C, reflects the magnitude of the film layer

effect. When C is zero, there is essentially no film layer effect. However, the film layer effect

becomes more significant as C increases. In Figure 5-14, the initial data points reflecting external

diffusion were omitted from the W-M plots, because they were not used for the straight-line

fittings of the linear portions. The intraparticle diffusion rate constants and the C values obtained

from the W-M plots are summarized in Table 5-8.

Table 5-8: Weber-Morris intraparticle diffusion rate parameters

Adsorbent

Intraparticle Diffusion

Rate Constant,

kid (mg/g min0.5

)

Weber-Morris

Constant, C

(mg/g)

r2

CO2-ADC-9h 7.43 8.9 0.929

CO2-ADC-9h-pAir 3.06 18 0.865

CO2-ADC-9h-pO2 1.23 18 0.844

The intraparticle diffusion rate constants in Table 5-8 decreased in the following in the order:

CO2-ADC-9h-pO2 < CO2-ADC-9h-pAir < CO2-ADC-9h. This elucidates that the mass transfer

resistance through the internal pores increased with surface oxidation. The values of the W-M

constant (C) for the adsorbents were above zero, suggesting that the external film diffusion still

has some effects on the initial diffusion process. However, the relatively small values of C imply

that the mass transfer resistance due to external diffusion was not substantial. This was expected

because the effect of external film diffusion was previously minimized by adjusting the shaking

speed. The higher C values of CO2-ADC-9h-pO2 and CO2-ADC-9h-pAir imply a greater

boundary layer effect for surface-oxidized adsorbents. The surface chemistry data of the

80

adsorbents were compared to the intraparticle rate parameters to fully understand the effect of

surface oxygen groups on adsorption kinetics (Table 5-9).

Table 5-9: Correlations between adsorbent surface chemistry and Weber-Morris rate parameters

Adsorbent

Total Surface

Oxygen Group

(Boehm Titration),

(mmol/g)

‘Organic’

Elemental

Oxygen

(wt%)

Intraparticle

Diffusion Rate

Constant,

kid (mg/g min0.5

)

W-M

Constant,

C, (mg/g)

CO2-ADC-9h 0.63 1.53 7.43 8.9

CO2-ADC-9h-pAir 1.05 3.50 3.06 18

CO2-ADC-9h-pO2 3.07 10.1 1.23 18

Table 5-9 exhibits a decreasing trend in the intraparticle diffusion (ID) rate constant with the

increasing oxygen surface groups of the adsorbents. The same trend is also observed between the

ID rate constant and the ‘organic’ oxygen content of the adsorbents. This suggests that the

adsorption kinetics of model naphthenic acid is adversely affected by the introduction of surface

oxygen groups.

Surface oxidation introduces oxygen functionalities such as carboxylic, lactonic and phenolic

groups on activated carbon surface as demonstrated by Boehm titration. The presence of these

functionalities presumably induces steric hindrance, limiting the adsorbate diffusion into the

internal pores. Consequently, higher surface oxygen content results in lower intraparticle

diffusion rate constant. Secondly, the oxygen heteroatoms on the activated carbon surface are

electron-rich. The same is also true for the oxygen atoms of the model naphthenic acid.

Therefore, electrostatic repulsion may occur between the oxygen groups on the carbon surface

81

and the model naphthenic acid during the adsorption process. This would also reduce the rate

constant. Wu and Pendleton (2001) studied the adsorption of dodecanoic and octanoic acid onto

oxygen-rich activated carbon. They suggested that the water molecules from the solution phase

may form clusters around the hydrophilic oxygen groups on the carbon surface. These clusters

prohibit the hydrophobic interactions between the carbon chain of the adsorbate and the activated

carbon surface [78]. This phenomenon may also contribute to the decreasing rate constant

observed in this study.

82

5.2.2 Equilibrium Adsorption Study

The equilibrium adsorption experiments were performed by fixing the initial adsorbate

concentration constant while varying the adsorbent dosage. The equilibration time for the

experiments was determined from preliminary experiments: 18 hours for the commercially

activated carbon (CAC) and 20 hours for the petroleum coke-derived activated carbons. The

experimental data was formulated into several isotherm models to determine the best model fit.

The effect of surface functional groups on the equilibrium adsorption of the model naphthenic

acid was investigated. A thermodynamic analysis was carried out to understand the nature of the

mNA adsorption. The equilibrium adsorption of NA-type total acid-extractable organics (TAOs)

from real tailings pond water was also investigated.

5.2.2.1 Equilibrium Adsorption Isotherms

CAC, Steam-ADC-3h, and CO2-ADC-6h were used to determine the isotherm model that best

represented the equilibrium adsorption data. The adsorption isotherms are depicted in Figure 5-

15. The experimental data was formulated into the Langmuir, Freundlich and Temkin isotherm

models. The model parameters are summarized in Table 5-10. The raw data can be found in

Appendix D along with isotherm model fitting graphs.

83

Figure 5-15: Model naphthenic acid adsorption isotherms

The Langmuir model generated negative values of the maximum adsorption capacity (qm) and

the Langmuir constant (kL) for CO2-ADC-6h, indicating an inadequate fit. The correlation

coefficients of both the Langmuir and the Temkin model fittings for Steam-ADC-3h were quite

low (r2 = 0.813 and 0.821, respectively). In contrast, the Freundlich model fittings of the

adsorption data displayed high correlation coefficients for all three adsorbents. The Freundlich

fits are represented by the dashed lines in Figure 5-15.

Although the adsorption capacity constant of an isotherm model is generally used to compare the

performances of different adsorbents, the Freundlich adsorption capacity constant kF was not

used for this purpose in this study. The unit of kF [(mg/g)(L/mg)1/n

] is dependent on the ‘n’

0

70

140

210

280

350

0 10 20 30 40 50

Eq

uil

ibri

um

Ad

sorb

ed

Qu

an

tiry

, q

e (m

g/g

)

Equilibrium Solution Concentration, Ce (mg/L)

CO2-ADC-6h CO2-ADC-6h Freundlich fit

CAC CAC Freundlich Fit

Steam-ADC-3h Steam-ADC-3h Freundlich fit

Conditions: Solution volume = 40 mL; pH = 8.5; Temperature = 25oC.

Shaking speed = 200 rpm.

Initial mNA concentration ~44 mg/L; Adsorbent dosage range ~2-40 mg.

84

values. Since the ‘n’ values of the adsorbents are quite different (Table 5-10), the units of kF will

vary for the adsorbents. Hence, kF is not an adequate comparative basis. Instead the equilibrium

adsorption quantity (qe) at a fixed equilibrium solution concentration was used for comparing the

performances of the adsorbents.

Table 5-10: Adsorption isotherm model parameters for mNA adsorption

Activated Carbon Adsorbents

Isotherm Models CAC CO2-ADC-6h Steam-ADC-3h

Langmuir

qm, (mg/g) 313 -53.5 152

kL, (L/mg) 0.84 -0.02 0.02

r2 0.998 0.926 0.813

Freundlich

kF, (mg/g)(L/mg)1/n

167 0.026 2.15

n 5.10 0.40 1.04

r2 0.925 0.967 0.916

Temkin

B, (J/mol) 43 205 49

AT, (L/g) 50 0.06 0.11

r2 0.957 0.925 0.821

The adsorption isotherms in Figure 5-15 show that the commercially activated carbon (CAC) had

the highest equilibrium adsorbed quantity for any given solution concentration, followed by

CO2-ADC-6h and Steam-ADC-3h, respectively. In Table 5-11, a positive correlation is observed

between the mesoporous SSA of the adsorbents and the equilibrium adsorbed quantity. This

seems to suggest that the pore size distribution may play a role in equilibrium adsorption of the

85

model naphthenic acid. However, the lesser adsorption onto petroleum coke-derived activated

carbon may also be caused by poor wettability of these adsorbents.

Table 5-11: Relationship between mesoporous SSA and equilibrium adsorbed quantity

Adsorbents Mesoporous SSA,

(m2/g)

Equilibrium Adsorbed Quantity,

qe (mg/g) [at Ce=20 mg/L]

CAC 213 300

CO2-ADC-6h 92 45

Steam-ADC-3h 85 25

5.2.2.2 Effect of Surface Functional Groups

CO2-ADC-9h and the three post-treated activated carbons were used to determine the effect of

surface functional groups on the model naphthenic acid adsorption. The adsorption isotherms are

depicted in Figure 5-16.

Among the four adsorbents, the surface-oxidized activated carbons (CO2-ADC-9h-pAir and

CO2-ADC-9h-pO2) exhibited the lowest equilibrium adsorption of the model NA. Between CO2-

ADC-9h and CO2-ADC-9h-pNH3, the latter displayed a higher mNA adsorption at equilibrium.

Figure 5-16 also displays the Freundlich model fits for the adsorbents.

86

Figure 5-16: Adsorption isotherms reflecting the effect of adsorbent surface functional groups

Table 5-12: Effect of surface functional groups on the equilibrium adsorption of mNA

Adsorbents

Acidic Surface

Oxygen Group,

(mmol/g)

Elemental

Oxygen Content,

(wt%)

Equilibrium Adsorbed

Quantity (mg/g), [at Ce = 20 mg/L]

CO2-ADC-9h 0.62 1.53 63

CO2-ADC-9h-pAir 1.05 3.50 52

CO2-ADC-9h-pO2 3.07 10.1 34

CO2-ADC-9h-pNH3 n/m n/m 102

*n/m= not measured

0

20

40

60

80

100

120

0 5 10 15 20 25 30 35

Eq

uil

ibri

um

Ad

sorb

ed

Qu

an

tity

, q

e (

mg/g

)

Equilibrium Solution Concentration, Ce (mg/L)

CO2-ADC-9h CO2-ADC-9h Freundlich fit

CO2-ADC-9h-pAir CO2-ADC-9h-pAir Freundlich fit

CO2-ADC-9h-pO2 CO2-ADC-9h-pO2 Freundlich fit

CO2-ADC-9h-pNH3 CO2-ADC-9h-pNH3 Freundlich fit

R2 = 0.912

Conditions: Solution volume = 40 mL; pH = 8.5; Temperature = 25oC.

Shaking speed = 200 rpm.

Initial mNA concentration ~33 mg/L; Adsorbent dosage range ~3-22 mg.

R2

= 0.905 R

2

= 0.948

R2

= 0.992

87

As listed in Table 5-12, both of the surface-oxidized activated carbons exhibits lower equilibrium

adsorbed quantity of mNA than their parent adsorbent (CO2-ADC-9h). This implies that surface

oxidation of the adsorbents has an adverse effect on the equilibrium adsorption of the model NA.

As the amount of acidic surface oxygen group and the elemental oxygen content of the

adsorbents increased, the equilibrium adsorbed mNA quantity decreased (Table 5-12).

As suggested in Section 5.2.1.4, the development of oxygen functionalities on activated carbon

surface may induce steric hindrance, which limits the diffusion of the adsorbate molecules into

the adsorbent pores. This may be one of the reasons why lower equilibrium adsorption was

observed for the surface-oxidized adsorbents. Secondly, the oxygen-rich adsorbent surfaces will

most likely be negatively charged. Thereby, electrostatic repulsion between the oxygen

functionalities of the adsorbents and the oxygen groups of the adsorbate molecules is expected.

This effect will also prohibit the diffusion and the subsequent adsorption process [78]. Finally,

the polar water molecules in the solution phase typically have a high affinity for hydrophilic

oxygen functionalities. This could lead to the formation of ‘water clusters’ surrounding the

oxygen functionalities of the adsorbents, which would reduce the hydrophobic interactions

between the activated carbon surface and the carbon chain of the adsorbate [78]. This could also

contribute to the low equilibrium adsorption onto the surface-oxidized activated carbons.

For any given equilibrium solution concentration (Ce), the equilibrium adsorbed mNA quantity

(qt) was the highest for the gaseous ammonia post-treated activated carbon (CO2-ADC-9h-

pNH3). At Ce = 20 mg/L, the equilibrium adsorbed mNA quantity was 102 mg/g for CO2-ADC-

9h-pNH3 and 63 mg/g for CO2-ADC-9h. The adsorption capacity of the adsorbent was evidently

enhanced upon gaseous ammonia treatment. FTIR analysis of CO2-ADC-9h-pNH3 surface

revealed the development nitrogen-based functional groups (amine, amide and nitrile). This

88

finding was consistent with the elemental analysis data, which showed a 10% increase in the

elemental nitrogen content after post-treatment. Due to the presence of basic surface groups, the

surface of the adsorbent is expected to be positively charged when submerged in the STPW

solution. The interaction between the positively charged adsorbent surface and the negatively

charged naphthenate molecules is suspected to enhance the equilibrium adsorption. This suggests

that chemisorption most likely plays a role in the adsorption process between basic adsorbents

and the model naphthenic acid.

From the Freundlich fit of the adsorption isotherm, it was calculated that 2.7 g/L of CO2-ADC-

9h-pNH3 would be needed to reduce mNA concentration in STPW from 120 mg/L to 2.5 mg/L

(~98 % removal). Lowering the pH of the tailings pond water solution could increase the

adsorption of mNA onto basic adsorbents. At low pH values, the solution will have more protons

to donate to the adsorbent surface, making it more positively charged. This would enhance the

adsorption of the model naphthenic acid.

5.2.2.3 Thermodynamic Analysis of the mNA Adsorption

Thermodynamic parameters such as the change in Gibbs free energy, enthalpy, entropy and

equilibrium constant often provide a good understanding about the nature of an adsorption

process. The thermodynamic equilibrium constant (Kc) for an adsorption process is defined as

follows [84]-

89

where CAE is the adsorbed amount of adsorbate at equilibrium, Co is the initial concentration of

the adsorbate (mg/L), and Ce is the equilibrium solution concentration of the adsorbate (mg/L).

Using CO2-ADC-6h, the equilibrium constant (Kc) for the model naphthenic acid adsorption was

determined experimentally at three different temperatures (22oC, 40

oC and 60

oC) to investigate

the temperature effect on Kc. Triplicate samples were run for each temperature. The same initial

model naphthenic acid concentration was used for all the samples. The results are summarized in

the Table 5-13.

Table 5-13: Effect of temperature on equilibrium constant for mNA adsorption

Temperature

(K)

Equilibrium mNA

concentration, Ce* (mg/L)

Equilibrium

Constant, Kc

295 10.9 3.68 ± 0.82

313 12.6 3.06 ± 0.13

333 19.6 1.60 ± 0.09

* Ce is the average value from triplicate runs

The higher equilibrium mNA concentrations (Ce) in Table 5-13 infer to a lesser adsorption of the

adsorbate. The equilibrium constant exhibits a decreasing trend with increasing solution

temperature. The solubility of a solute is generally known to increase with increasing solution

temperature. Consequently, the intermolecular forces between the solvent and the solute

(adsorbate) molecules become stronger [64]. As a result, the adsorbate is more difficult to

adsorb. This may explain why the equilibrium constant decreases with increasing solution

temperature. This is generally a nature of a physisorption-dominated adsorption process.

Conditions: Solution volume = 40 mL; pH = 8.5; Shaking speed = 200 rpm.

Initial mNA concentration (Co) = 51 mg/L; CO2-ADC-6h dosage = 30 mg.

Equilibration time = 18 hours

90

The change in Gibbs free energy is defined as follows [73]-

where ΔGo is the change in Gibbs free energy at standard conditions (J/mol), R is the universal

gas constant (8.3145 J/mol K), T is the temperature (K) and QR is the reaction quotient. At

equilibrium, ΔG = 0 and the reaction quotient becomes the equilibrium constant (QR = Kc). By

applying these conditions to equation 5.5, the following expression can be obtained-

The change in Gibbs free energy can also be defined in terms the change in enthalpy and entropy

[73]-

The following expression can be obtained by substituting equation 5.6 into equation 5.7-

where ΔHo is the change in enthalpy (kJ/mol) and ΔS

o is the change in entropy (J/mol K) at

standard conditions. Figure 5-17 displays a plot of ln Kc versus 1/T. The slope and the y-intercept

were used to determine the values of ΔHo and ΔS

o. The ΔG

o values for corresponding

temperatures were then calculated using equation 5.7. The thermodynamic parameters are

summarized in Table 5-14.

91

Figure 5-17: Plot of ln Kc versus 1/T for the estimation of thermodynamic parameters

Table 5-14: Thermodynamic parameters for the mNA adsorption onto activated carbon

Temperature (K)

295 313 333

ΔGo (kJ/mol)

-3.19 -2.91 -1.29

ΔHo (kJ/mol) -17.9

ΔSo (J/mol K) -49.2

According to Zhu et al. (2011), a more negative value of ΔGo typically indicates a greater

driving force for adsorption. Table 5-14 shows that ΔGo values become more negative with

decreasing temperature, suggesting that the driving force for adsorption is greater at lower

temperature. This is consistent with the experimental observation.

The negative value of ΔHo suggests that the adsorption process is exothermic. According to

Atkins (2002), a relatively low magnitude of ΔHo

(<|10-15| kJ/mol) generally suggests a

y = 2153.2x - 5.92

R² = 0.904

0

0.3

0.6

0.9

1.2

1.5

0.0028 0.003 0.0032 0.0034 0.0036

ln (

Kc)

1/T, (1/K)

92

physisorption-dominated process while chemisorption is reflected by relatively high magnitudes

of enthalpy of adsorption (>|20|kJ/mol) [3]. Based on the enthalpy value obtained for the model

naphthenic acid adsorption (-17.9 kJ/mol), a physisorption-dominated process may be suggested

for adsorption onto CO2-activated petroleum coke.

Due to the mNA adsorption, the disorder or the randomness at the solid-liquid interface

decreases from the initial state. This explains why a negative ΔSo

value is observed at the solid-

liquid interface upon adsorption.

5.2.2.4 Equilibrium Adsorption of TAOs from Real Tailings Pond Water

The equilibrium adsorption of NA-type total acid-extractable organics (TAOs) from ‘real’

tailings pond water (RTPW) was investigated. TAOs are complex naphthenic acid mixtures

quantified by FTIR spectroscopy. The tailings pond water was provided by Suncor Energy.

Figure 5-18 displays the isotherms for the TAOs adsorption from RTPW and the mNA

adsorption from synthetic tailings pond water (STPW). CO2-ADC-9h-pNH3 was used as the

adsorbent for both experiments.

93

Figure 5-18: Adsorption isotherms of TAOs and mNA onto CO2-ADC-9h-pNH3

From the above figure, it is observed that the adsorption of the model naphthenic acid from

STPW is much higher than the adsorption of the TAOs from RTPW. Table 5-15 compares the

equilibrium adsorbed quantity of TAOs and mNA at equilibrium solution concentration of 20

mg/L.

Table 5-15: Equilibrium adsorption of TAOs and mNA onto CO2-ADC-9h-pNH3

TAOs from RTPW mNA from STPW

Equilibrium Adsorbed

Quantity (mg/g) [at Ce = 20 mg/L]

27 102

0

20

40

60

80

100

120

0 5 10 15 20 25

Eq

uil

ibri

um

Ad

sorb

ed

Qu

an

tity

, q

e (

mg/g

)

Equilibrium Solution Concentration, Ce (mg/L)

Model NA from STPW TAOs from RTPW

Model NA Freundlich fit TAOs Freundlich fit

Conditions: Initial mNA conc. ~33 mg/L; Initial TAOs conc. ~ 25 mg/L.

Solution volume = 40 mL; Temperature = 25oC.

Shaking speed = 200 rpm; Adsorbent dosage range ~3-22 mg.

R2

= 0.912

R2

= 0.887

94

The above results show that under the same experimental conditions, the equilibrium adsorbed

quantity of NA-type total acid-extractable organics (TAOs) from real TPW was only 26% of that

of the model NA from synthetic TPW. There may be two reasons for this observation.

Firstly, the real tailing pond water is known to contain a wide range of chemical compounds.

Zubot (2010) reported the presence of a wide variety of polycyclic aromatic hydrocarbons such

as naphthalene, fluorene, anthracene, and pyrene in Syncrude tailings pond water. In a separate

study, Small (2011) reported the presence of the following trace metals in tailings pond water:

Al, Ni, Mo, Mn, Cu, Pb and V. It is highly probable that these other chemicals present in RTPW

will compete with TAOs (complex naphthenic acids mixture) for the active sites of activated

carbon. This would consequently reduce the amount of active site available for the TAOs

adsorption, and thereby reducing the overall adsorption of TAOs. However, this is not the case

for the model naphthenic acid adsorption since the synthetic tailings pond water did not contain

those other chemicals. This may explain why the model naphthenic adsorption from STPW is

greater than the adsorption of the TAOs from RTPW using the same adsorbent.

Secondly, the naphthenic acids present in the real tailing ponds water may be larger in

dimensions in comparison to the model acid. Zubot (2010) characterized the structures of the

naphthenic acids found in real tailings pond water and concluded that most of the NA molecules

in real tailings pond water have carbon numbers between 12 to16 and z values between -2 to -6

(1 to 3 ring structures). Hence, it is conceivable that the larger naphthenic acid molecules in the

real tailing pond water may not be able to diffuse into the smaller pores that are accessible to the

model naphthenic acid. This could also contribute to the lesser adsorption of the TAOs observed

in this study.

95

CHAPTER 6 CONCLUSIONS AND RECOMMENDATIONS

6.1 Conclusions

In this study, the adsorptive removal of a model naphthenic acid and total acid-extractable

organics from oil sands tailings pond water was investigated using petroleum coke-derived

activated carbon. The adsorbents were produced from Suncor ‘delayed’ coke by several

activation methods. A significant portion of this research study was dedicated to understanding

the effects of the adsorbent pore structure and surface functionalities on the model naphthenic

acid adsorption by means of a kinetic and an equilibrium study. The conclusions drawn from

these studies are summarized below-

Activated Carbon Synthesis and Characterization:

1. The steam-activated carbon exhibited a high microporous volume (75%), whereas the

CO2-activated carbons (CO2-ADC-6h and CO2-ADC-9h) displayed high mesoporous

volume (51% and 67%, respectively). This trend in the pore size distribution was most

likely influenced by the difference in the size of the reagent molecules between H2O and

CO2.

2. The specific surface area and the total pore volume of CO2-activated carbon increased

with activation time. CO2-ADC-6h had a SSA of 276 m2/g, while CO2-ADC-9h had a

SSA of 405 m2/g.

3. The post-oxidative treatment of activated carbon introduced acidic oxygen

functionalities on the adsorbent surface. Carboxylic, lactonic, and phenolic groups were

detected on the post-oxidized adsorbent and quantified by Boehm titration. The results

were consistent with the findings from FTIR surface analysis of the adsorbents. As

96

expected, oxidation with an oxygen-rich stream was more effective than oxidation with

air.

4. Post-treatment with gaseous ammonia introduced basic nitrogen groups on the activated

carbon surface. FTIR analysis indicated the presence of nitrile, amine and amide groups

on the ammonia-treated activated carbon. Elemental analysis of the adsorbent displayed a

10% increase in the elemental nitrogen content after post-treatment with ammonia.

Adsorption Kinetics:

A batch-type kinetic study was performed with only the model naphthenic acid. The key findings

are highlighted below-

1. The experimental kinetic data was best represented by the empirical pseudo-2nd

order rate

model. Positive correlation was observed between the pseudo-2nd

order rate constant and

the mesoporous volume fraction of the adsorbents. Given the molecular dimensions of

the model acid (1.2x0.25x0.25 nm), it was expected that the diffusion of the adsorbate

molecules through mesopores would be conceivably easier than the diffusion through

micropores.

2. The rate-controlling step is most likely the intraparticle diffusive transport of naphthenic

acid. Key evidences include the relatively low value of the apparent activation energy

(7.7 kJ/mol) and the observed positive correlation between the initial rate of adsorption

and the initial bulk concentration of the adsorbate.

3. The ‘effect of surface functional groups’ study revealed a negative correlation between

the surface oxygen content of the adsorbents and the model naphthenic acid adsorption

rates. The Weber-Morris intraparticle diffusion rate model was used to interpret the

97

experimental data. The intraparticle diffusion rate constants decreased with increasing

adsorbent surface oxygen content. The steric hindrance and electrostatic repulsion caused

by the oxygen functionalities presumably limits diffusive mass transfer.

Equilibrium Adsorption Study:

1. The model naphthenic acid equilibrium adsorption data was best represented by the

Freundlich isotherm model, which generally indicates heterogeneity of adsorbent

surface and multilayer adsorption, suggesting a major role of physisorption. For given

equilibrium concentration of the adsorbate in water, a positive correlation between the

equilibrium adsorbed quantity of mNA and the adsorbent mesoporous SSA was

observed.

2. A negative correlation between the equilibrium adsorbed quantity and the adsorbent

surface oxygen content suggested that acidic oxygen functional groups had

diminishing effects on the equilibrium adsorption of the model naphthenic acid. This

was most likely due to the steric hindrance and electrostatic repulsion caused by the

oxygen-based surface functional groups.

3. Among the four adsorbents used in the ‘effect of surface functionalities’ study, the

ammonia post-treated activated carbon displayed the highest equilibrium adsorbed

mNA quantity. The expected charge-based interaction between the model naphthenic

acid and the basic nitrogen functional groups on the adsorbent surface may have

caused the high equilibrium adsorption. Consequently, there is a potential for

chemisorption in naphthenic acid adsorption with chemically-modified activated

carbons.

98

4. There was a negative temperature dependence of adsorption equilibrium constant of

the model naphthenic acid - a character of physisorption. The thermodynamic

analysis on CO2-activated coke revealed negative ΔHo and ΔS

o, which indicates an

exothermic process and a decrease in system randomness. Consequently, the removal

of naphthenic acid with CO2-activated coke is likely dominated by physisorption.

5. Using the ammonia post-treated activated carbon, the adsorption of NA-type total

acid-extractable organics (TAOs) from real tailings pond water was compared to that

from the synthetic tailings pond water. The adsorption amount from the latter was

greater. This observation may suggest the competitive adsorption between NA-type

organics and other organics in tailings pond water. Another possibility that needs

verification is that in tailings pond water the NA-type TAOs consist of mainly

naphthenic acids larger than the model naphthenic acid. It is anticipated that the

intraparticle diffusion and the subsequent adsorption of these larger naphthenic acids

might have been hindered due to physical constraints resulting from their larger size.

99

6.2 Recommendations

1. The FTIR technique for naphthenic acid detection cannot differentiate between different

naphthenic acid species. This is not an issue when working with a single naphthenic acid

species (i.e., model naphthenic acid). However, when the studying the adsorption of

complex naphthenic acid mixtures from tailings pond water, it would be beneficial to

employ an instrument that can both characterize and quantify different naphthenic acids

species (i.e., GC-MS or HPLC-MS). Information about the adsorption of different

naphthenic acid species would allow one to better design and tailor the activated carbon

adsorbents.

2. The kinetic study demonstrated that mesopores are favourable for the model naphthenic

acid adsorption. Zubot’s work (2011) suggested that majority of the naphthenic acid

species in tailing pond water have carbon numbers between 12 and 16 and z values

between -2 and -6. This implies that most of the naphthenic acid species in tailing pond

water are either larger than or just as large as the model acid used in this study. Hence,

the use of highly microporous activated carbon for adsorption should be avoided.

Techniques that produce mostly mesoporous adsorbents should be investigated.

3. The point of zero charge (pzc) of the surface-modified activated carbons could be

obtained to better understand the interactions between the surface functionalities of

activated carbon and naphthenic acid. Increasing the adsorbent basicity had a positive

effect on the equilibrium adsorption of naphthenic acid. It would be worth awhile to

investigate more efficient ways to introduce basic functionalities on activated carbon

surface.

100

4. Adsorption of surfactants is influenced by the surface charge of adsorbents, and the

surface charge of an adsorbent is pH-dependent. Hence, the effect of pH on adsorption

should be investigated.

5. Tailings pond water is known to contain a large quantity of fine clay minerals (i.e.,

kaolinite). The effects of these minerals on naphthenic acid adsorption should also be

investigated.

6. The tailings pond water should be analyzed for other chemical groups (i.e., heavy metals,

total organic carbons) before and after the adsorption process. With this information, it

would easier to understand the competitive adsorption between naphthenic acids and

other chemical species. Chemicals leached from petroleum coke-derived activated carbon

could also be detected from such analysis.

7. The toxicity of naphthenic acids is known to vary with their chemical structure. It would

be beneficial to target the naphthenic acid species that are most suspected for causing the

toxicity.

101

REFERENCES

[1] Amuda, O. S., Giwa, A. A., & Bello, I. A. (2007). Removal of heavy metal from industrial

wastewater using modified activated coconut shell carbon. Biochemical Engineering

Journal, 36(2), 174-181. doi: 10.1016/j.bej.2007.02.013

[2] Anirudhan, T. S., & Sreekumari, S. S. (2011). Adsorptive removal of heavy metal ions from

industrial effluents using activated carbon derived from waste coconut buttons. Journal of

Environmental Sciences, 23(12), 1989-1998. doi: 10.1016/S1001-0742(10)60515-3

[3] Atkins, P. W. (2002). In DePaula J. (Ed.), Physical chemistry (7th ed. ed.). New York:

Freeman.

[4] Awoyemi, A. (2011). Understanding the adsorption of polycyclic aromatic hydrocarbons

from aqueous phase onto activated carbon. Master of Applied Science Thesis, University of

Toronto

[5] Bandosz, T. J. (1999). Effect of pore structure and surface chemistry of virgin activated

carbons on removal of hydrogen sulfide. Carbon, 37(3), 483-491. doi: 10.1016/S0008-

6223(98)00217-6

[6] Bandosz, T. J. (2006). In edited by Teresa J. Bandosz. (Ed.), Activated carbon surfaces in

environmental remediation / Amsterdam ;London : Elsevier, 2006.

[7] Bansal, R. C., Goyal, M. (2005), Activated carbon adsorption. New York: Taylor and

Francis, 2005.

[8] Bataineh, M., Scott, A. C., Fedorak, P. M., & Martin, J. W. (2006). Capillary HPLC/QTOF-

MS for characterizing complex naphthenic acid mixtures and their microbial transformation.

Analytical Chemistry, 78(24), 8354-8361. doi: 10.1021/ac061562p

102

[9] Baumeister, U., Hartung, H., & Jaskólski, M. (1982). The crystal and molecular structure of

the nematogenic compound 4′-cyanophenyl-4-n-pentylcyclohexanoate. Crystal Research

and Technology, 17(2), 153-160.

[10] Biryukova, O. V., Fedorak, P. M., & Quideau, S. A. (2007). Biodegradation of naphthenic

acids by rhizosphere microorganisms. Chemosphere, 67(10), 2058-2064. doi:

10.1016/j.chemosphere.2006.11.063

[11 Boehm, H. P. (1994). Some aspects of the surface chemistry of carbon blacks and other

carbons. Carbon, 32(5), 759-769. doi: 10.1016/0008-6223(94)90031-0

[12] Boparai, H. K., Joseph, M., & O’Carroll, D. M. (2011). Kinetics and thermodynamics of

cadmium ion removal by adsorption onto nano zerovalent iron particles. Journal of

Hazardous Materials, 186(1), 458-465. doi: 10.1016/j.jhazmat.2010.11.029

[13] Bratu, M. (2008). Activation of fluid coke for mercury removal from gases. (M.Sc.,

University of Alberta (Canada)). ProQuest Dissertations and Theses, . (304409210).

[14] Cansado, I. P. P., Mourão, P. A. M., Falcão, A. I., Carrott, M. M. L. R., & Carrott, P. J. M.

(2012). The influence of the activated carbon post-treatment on the phenolic compounds

removal. Fuel Processing Technology, 103(0), 64-70. doi: 10.1016/j.fuproc.2011.10.015

[15] Chakraborty, A., Deva, D., Sharma, A., & Verma, N. (2011). Adsorbents based on carbon

microfibers and carbon nanofibers for the removal of phenol and lead from water. Journal

of Colloid and Interface Science, 359(1), 228-239. doi: 10.1016/j.jcis.2011.03.057

[16] Clemente, J. S., & Fedorak, P. M. (2004). Evaluation of the analyses of tert-

butyldimethylsilyl derivatives of naphthenic acids by gas chromatography-electron impact

mass spectrometry. Journal of Chromatography A, 1047(1), 117-128. doi:

10.1016/j.chroma.2004.06.065

[17] Clemente, J. S., & Fedorak, P. M. (2005). A review of the occurrence, analyses, toxicity,

and biodegradation of naphthenic acids. Chemosphere, 60(5), 585-600. doi:

10.1016/j.chemosphere.2005.02.065

103

[18] Crank, J. (1964). The mathematics of diffusion. Oxford, Clarendon Press [1964].

[19] DiPanfilo, R., & Egiebor, N. O. (1996). Activated carbon production from synthetic crude

coke. Fuel Processing Technology, 46(3), 157-169. doi: 10.1016/0378-3820(95)00054-2

[20] Drzewicz, P., Afzal, A., El-Din, M. G., & Martin, J. W. (2010). Degradation of a model

naphthenic acid, cyclohexanoic acid, by vacuum UV (172 nm) and UV (254 nm)/H2O2. The

Journal of Physical Chemistry A, 114(45), 12067-12074. doi: 10.1021/jp105727s

[21] Everett, D. H. (1971). Manual of symbols and terminology for physicochemical quantities

and units. Butterworths and Pure and Applied Chemistry, 21(1)

[22] Exeter Analytical Inc. (2005). CE-440 elemental analyzer manual;

http://www.eai1.com/theory.htm.

[23] Fuertes, A. B., & Alvarez, S. (2004). Graphitic mesoporous carbons synthesised through

mesostructured silica templates. Carbon, 42(15), 3049-3055. doi:

10.1016/j.carbon.2004.06.020

[24] Gamal El-Din, M., Fu, H., Wang, N., Chelme-Ayala, P., Pérez-Estrada, L., Drzewicz, P., . . .

Smith, D. W. (2011). Naphthenic acids speciation and removal during petroleum-coke

adsorption and ozonation of oil sands process-affected water. Science of the Total

Environment, 409(23), 5119-5125. doi: 10.1016/j.scitotenv.2011.08.033

[25] Goertzen, S. L., Thériault, K. D., Oickle, A. M., Tarasuk, A. C., & Andreas, H. A. (2010).

Standardization of the boehm titration. part I. CO2 expulsion and endpoint determination.

Carbon, 48(4), 1252-1261. doi: 10.1016/j.carbon.2009.11.050

[26] Grewer, D. M., Young, R. F., Whittal, R. M., & Fedorak, P. M. (2010). Naphthenic acids

and other acid-extractables in water samples from alberta: What is being measured? Science

of the Total Environment, 408(23), 5997-6010. doi: 10.1016/j.scitotenv.2010.08.013

104

[27] Guo, S., Peng, J., Li, W., Yang, K., Zhang, L., Zhang, S., & Xia, H. (2009). Effects of CO2

activation on porous structures of coconut shell-based activated carbons. Applied Surface

Science, 255(20), 8443-8449. doi: 10.1016/j.apsusc.2009.05.150

[28] Han, X., MacKinnon, M. D., & Martin, J. W. (2009). Estimating the in situ biodegradation

of naphthenic acids in oil sands process waters by HPLC/HRMS. Chemosphere, 76(1), 63-

70. doi: 10.1016/j.chemosphere.2009.02.026

[29] Ho, Y. S., & McKay, G. (1999). Pseudo-second order model for sorption processes. Process

Biochemistry, 34(5), 451-465. doi: 10.1016/S0032-9592(98)00112-5

[30] Holowenko, F. M., MacKinnon, M. D., & Fedorak, P. M. (2001). Naphthenic acids and

surrogate naphthenic acids in methanogenic microcosms. Water Research, 35(11), 2595-

2606. doi: 10.1016/S0043-1354(00)00558-3

[31] Holowenko, F. M., MacKinnon, M. D., & Fedorak, P. M. (2002). Characterization of

naphthenic acids in oil sands wastewaters by gas chromatography-mass spectrometry. Water

Research, 36(11), 2843-2855. doi: 10.1016/S0043-1354(01)00492-4

[32] Hsu, L., & Teng, H. (2000). Influence of different chemical reagents on the preparation of

activated carbons from bituminous coal. Fuel Processing Technology, 64(1-3), 155-166.

doi: 10.1016/S0378-3820(00)00071-0

[33] Huynh, D. (2012). The fate of sulphur and vanadium in oil-sands petroleum coke during

physical activation and while in contact with tailings pond water. Masters of Engineering

Thesis, University of Toronto

[34] Ihara, Y. (1992). Adsorption of anionic surfactants and related compounds from aqueous

solution onto activated carbon and synthetic adsorbent. Journal of Applied Polymer Science,

44(10), 1837-1840. doi: 10.1002/app.1992.070441017

[35] İnci, İ., Bayazit, Ş. S., & Uslu, H. (2011). Investigation of adsorption equilibrium and

kinetics of propionic acid and glyoxylic acid from aqueous solution by alumina. Journal of

Chemical & Engineering Data, 56(8), 3301-3308. doi: 10.1021/je2000765

105

[36] Janfada, A., Headley, J. V., Peru, K. M., & Barbour, S. L. (2006). A laboratory evaluation

of the sorption of oil sands naphthenic acids on organic rich soils. Journal of Environmental

Science and Health, Part A, 41(6), 985-997. doi: 10.1080/10934520600620105

[37] Jia, C. Q. (2009). Activated carbon from Suncor delayed coke. Reporting Period #1,

September 1st to 15th, 2009

[38] Jiang, T., Hirasaki, G. J., Miller, C. A., & Ng, S. (2011). Wettability alteration of clay in

solid-stabilized emulsions. Energy & Fuels, 25(6), 2551-2558. doi: 10.1021/ef2000079

[39] Jivraj, M., MacKinnon, M. D., & Fung, B. (1996). Naphthenic acid extraction and

quantitative analysis with FT-IR spectroscopy. Syncrude Canada Ltd. Report,

[40] Kannel, P. R., & Gan, T. Y. (2012). Naphthenic acids degradation and toxicity mitigation in

tailings wastewater systems and aquatic environments: A review. Journal of Environmental

Science and Health, Part A, 47(1), 1-21. doi: 10.1080/10934529.2012.629574

[41] Kavanagh, R. J., Burnison, B. K., Frank, R. A., Solomon, K. R., & Van Der Kraak, G.

(2009). Detecting oil sands process-affected waters in the alberta oil sands region using

synchronous fluorescence spectroscopy. Chemosphere, 76(1), 120-126. doi:

10.1016/j.chemosphere.2009.02.007

[42] Kiran, S. K., Ng, S., & Acosta, E. J. (2011). Impact of asphaltenes and naphthenic

amphiphiles on the phase behavior of Solvent−Bitumen−Water systems. Energy Fuels,

25(5), 2223-2231. doi: 10.1021/ef1016285

[43] Machida, M., Amano, Y., & Aikawa, M. (2011). Adsorptive removal of heavy metal ions by

activated carbons. Carbon, 49(10), 3393-3393. doi: 10.1016/j.carbon.2011.04.011

[44] Mangun, C. L., DeBarr, J. A., & Economy, J. (2001). Adsorption of sulfur dioxide on

ammonia-treated activated carbon fibers. Carbon, 39(11), 1689-1696. doi: 10.1016/S0008-

6223(00)00300-6

106

[45] Merlin, M., Guigard, S. E., & Fedorak, P. M. (2007). Detecting naphthenic acids in waters

by gas chromatography-mass spectrometry. Journal of Chromatography A, 1140(1-2), 225-

229. doi: 10.1016/j.chroma.2006.11.089

[46] Missen, R. W. (1999). In Mims C. A., Saville B. A. (Eds.), Introduction to chemical

reaction engineering and kinetics. New York: Wiley.

[47] Mohamed, M. H., Wilson, L. D., Headley, J. V., & Peru, K. M. (2008). Screening of oil

sands naphthenic acids by UV-vis absorption and fluorescence emission spectrophotometry.

Journal of Environmental Science and Health, Part A, 43(14), 1700-1705. doi:

10.1080/10934520802330255

[48] Mohamed, M. H., Wilson, L. D., Headley, J. V., & Peru, K. M. (2010). Sequestration of

naphthenic acids from aqueous solution using β-cyclodextrin-based polyurethanes. Physical

Chemistry Chemical Physics, 13(3), 1112-1122. doi: 10.1039/c0cp00421a

[49] Morris, E. A. (2012). Modification of carbonaceous materials with sulfer and its impact on

mercury capture and sorbent regeneration. Doctor of Philosophy Thesis, University of

Toronto

[50] Morris, E. A., Choi, R., & Jia, C. Q. (2012). Sulfur dioxide as an activating agent for sulfur-

impregnated activated carbon produced from dense petroleum coke. Journal of Sulfur

Chemistry, , 1-12. doi: 10.1080/17415993.2012.739622

[51] Neimark, A. V., Lin, Y., Ravikovitch, P. I., & Thommes, M. (2009). Quenched solid density

functional theory and pore size analysis of micro-mesoporous carbons. Carbon, 47(7), 1617-

1628. doi: 10.1016/j.carbon.2009.01.050

[52] Oickle, A. M., Goertzen, S. L., Hopper, K. R., Abdalla, Y. O., & Andreas, H. A. (2010).

Standardization of the boehm titration: Part II. method of agitation, effect of filtering and

dilute titrant. Carbon, 48(12), 3313-3322. doi: 10.1016/j.carbon.2010.05.004

107

[53] Pastor-Villegas, J., & Durán-Valle, C. J. (2002). Pore structure of activated carbons

prepared by carbon dioxide and steam activation at different temperatures from extracted

rockrose. Carbon, 40(3), 397-402. doi: 10.1016/S0008-6223(01)00118-X

[54] Patrick, J. W. (1994). In Patrick J. W. (Ed.), Porosity in carbons characterization and

applications. London: E. Arnold.

[55] Pavia, D., Lampman, G., Kriz, G., & Vyvyan, J. (2009). Introduction to spectroscopy (4th

Edition ed.) Brooks/Cole.

[56] Pendleton, P., Wu, S., & Badalyan, A. (2002). Activated carbon oxygen content influence

on water and surfactant adsorption RID B-2148-2010. Journal of Colloid and Interface

Science, 246(2), 235-240. doi: 10.1006/jcis.2001.8052

[57] Peng, J., Headley, J. V., & Barbour, S. L. (2002). Adsorption of single-ring model

naphthenic acids on soils. Canadian Geotechnical Journal, 39(6), 1419-1426.

[58] Przepiórski, J., Skrodzewicz, M., & Morawski, A. W. (2004). High temperature ammonia

treatment of activated carbon for enhancement of CO2 adsorption. Applied Surface Science,

225(1-4), 235-242. doi: 10.1016/j.apsusc.2003.10.006

[59] Qu, Y., Zhang, C., Li, F., Bo, X., Liu, G., & Zhou, Q. (2009). Equilibrium and kinetics

study on the adsorption of perfluorooctanoic acid from aqueous solution onto powdered

activated carbon. Journal of Hazardous Materials, 169(1-3), 146-152. doi:

10.1016/j.jhazmat.2009.03.063

[60] Quantachrome Instruments. (2008). AUTOSORB-1 AS1Win version 1.5X operating

manual., 101.

[61] Ramana, D. K. V., Jamuna, K., Satyanarayana, B., Venkateswarlu, B., Rao, M. M., &

Seshaiah, K. (2010). Removal of heavy metals from aqueous solutions using activated

carbon prepared from cicer arietinum. Toxicological and Environ Chemistry, 92(8), 1447-

1460. doi: 10.1080/02772241003614312

108

[62] Rogers, V. V., Liber, K., & MacKinnon, M. D. (2002). Isolation and characterization of

naphthenic acids from athabasca oil sands tailings pond water. Chemosphere, 48(5), 519-

527. doi: 10.1016/S0045-6535(02)00133-9

[63] Rouquerol, J., Llewllyn, P., & Rouquerol, F. (2007). Is the BET equation applicable to

microporous adsorbents. Studies in Surface Science and Catalysis, 160, 49-56.

[64] Saha, P., & Chowdhury, S. (2011). Insight into adsorption thermodynamics. In M. Tadashi

(Ed.), Thermodynamics (pp. Chapter 16) InTech.

[65] Scott, A. C., Mackinnon, M. D., & Fedorak, P. M. (2005). Naphthenic acids in athabasca oil

sands tailings waters are less biodegradable than commercial naphthenic acids.

Environmental Science & Technology, 39(21), 8388-8394. doi: 10.1021/es051003k

[66] Scott, A. C., Young, R. F., & Fedorak, P. M. (2008). Comparison of GC–MS and FTIR

methods for quantifying naphthenic acids in water samples. Chemosphere, 73(8), 1258-

1264. doi: 10.1016/j.chemosphere.2008.07.024

[67] Seader, J. D., Henley, E. J. (2005), Separation process principles (2nd ed. ed.). Chichester:

John Wiley & Sons.

[68] Shawwa, A. R., Smith, D. W., & Sego, D. C. (2001). Color and chlorinated organics

removal from pulp mills wastewater using activated petroleum coke. Water Research, 35(3),

745-749. doi: 10.1016/S0043-1354(00)00322-5

[69] Sing, K., Everett, D. H., Haul, R., Moscou, L., Pierotti, R. A., Rouquerol, J., &

Siemieniewska, T. (1985). Reporting physisorption data for gas/solid systems with special

reference to the determination of surface area and porosity (recommendations 1984). Pure

and Applied Chemistry, 57(4), 603-619. doi: doi: 10.1351/pac198557040603

[70] Small, C., Ulrich, A., & Hashisho, Z. (2012). Adsorption of acid extractable oil sands

tailings organics onto raw and activated oil sands coke. Journal of Environmental

Engineering, 138(8), 833-840. doi: 10.1061/(ASCE)EE.1943-7870.0000543

109

[71] Small, C. C. (2011). Activation of delayed and fluid petroleum coke for adsorption and

removal of naphthenic acids from oil sands tailings pond water. Master of Science Thesis,

University of Alberta

[72] Small, C. C., Hashisho, Z., & Ulrich, A. C. (2012). Preparation and characterization of

activated carbon from oil sands coke. Fuel, 92(1), 69-76. doi: 10.1016/j.fuel.2011.07.017

[73] Smith, J. M., & Abbott, M. (2001). In Van Ness H. C. (Ed.), Introduction to chemical

engineering thermodynamics (7th ed. ed.). New York: McGraw-Hill.

[74] Strausz, O. P. (1977). The chemistry of alberta oil sand bitumen. Chemical Institute of

Canada and American Chemical Society Joint Confernece, Montreal,

[75] Valderrama, C., Gamisans, X., de las Heras, X., Farran, A., & Cortina, J. L. (2008).

Sorption kinetics of polycyclic aromatic hydrocarbons removal using granular activated

carbon: Intraparticle diffusion coefficients. Journal of Hazardous Materials, 157(2-3), 386-

396. doi: 10.1016/j.jhazmat.2007.12.119

[76] Wang, X., & Kasperski, K. L. (2010). Analysis of naphthenic acids in aqueous solution

using HPLC-MS/MS. Analytical Methods, 2(11), 1715-1722.

[77] Whitby, C. (2010). Chapter 3 - Microbial naphthenic acid degradation. Advances in applied

microbiology (pp. 93-125) Academic Press. doi: 10.1016/S0065-2164(10)70003-4

[78] Wu, S. H., & Pendleton, P. (2001). Adsorption of anionic surfactant by activated carbon:

Effect of surface chemistry, ionic strength, and hydrophobicity. Journal of Colloid and

Interface Science, 243(2), 306-315. doi: 10.1006/jcis.2001.7905

[79] Yalçin, M., Gürses, A., Doğar, Ç., & Sözbilir, M. (2005). The adsorption kinetics of

cethyltrimethylammonium bromide (CTAB) onto powdered active carbon. Adsorption,

10(4), 339-348. doi: 10.1007/s10450-005-4819-9

[80] Yang, R. T. (2003). Adsorbents : Fundamentals and applications. Hoboken, NJ: Wiley-

Interscience.

110

[81] Yen, T., Marsh, W. P., MacKinnon, M. D., & Fedorak, P. M. (2004). Measuring naphthenic

acids concentrations in aqueous environmental samples by liquid chromatography. Journal

of Chromatography A, 1033(1), 83-90. doi: 10.1016/j.chroma.2004.01.030

[82] Yuan, M., Tong, S., Zhao, S., & Jia, C. Q. (2010). Adsorption of polycyclic aromatic

hydrocarbons from water using petroleum coke-derived porous carbon. Journal of

Hazardous Materials, 181(1-3), 1115-1120. doi: 10.1016/j.jhazmat.2010.05.130

[83] Zhang, Y., Shi, W., Zhou, H., Fu, X., & Chen, X. (2010). Kinetic and thermodynamic

studies on the adsorption of anionic surfactant on quaternary ammonium cationic cellulose.

Water Environment Research, 82(6), 567-573.

[84] Zhu, H., Yang, X., Mao, Y., Chen, Y., Long, X., & Yuan, W. (2011). Adsorption of EDTA

on activated carbon from aqueous solutions. Journal of Hazardous Materials, 185(2-3),

951-957. doi: 10.1016/j.jhazmat.2010.09.112

[85] Zubot, W. A. (2010). Removal of naphthenic acids from oil sands process water using

petroleum coke. Master of Science Thesis, University of Alberta

[86] Zubot, W., MacKinnon, M. D., Chelme-Ayala, P., Smith, D. W., & Gamal El-Din, M.

(2012). Petroleum coke adsorption as a water management option for oil sands process-

affected water. Science of the Total Environment, 427–428(0), 364-372. doi:

10.1016/j.scitotenv.2012.04.024

111

APPENDIX A: Naphthenic acid quantification data

Table A1: mNA calibration data for Spectrum-BX Perkin Elmer FTIR spectrometer

Combined Monomer and Dimer Peak Height, (a.u.)

Concentration,

(mg/L) Run 1 Run 2 Run 3 Mean

Standard

Deviation

1 0.0019 0.0016 0.0027 0.0021 0.0005

5 0.0040 0.0040 - 0.0040 0.0000

10 0.0092 0.0099 0.0102 0.0098 0.0004

15 0.0186 0.0179 0.0215 0.0193 0.0016

20 0.0237 0.0249 0.0247 0.0244 0.0005

25 0.0335 0.0339 0.0321 0.0332 0.0008

50 0.0691 0.0703 0.0672 0.0689 0.0013

100 0.1258 0.1275 0.1277 0.1270 0.0009

150 0.2070 0.2075 0.2109 0.2085 0.0017

200 0.2823 0.2822 0.2794 0.2813 0.0013

250 0.3627 0.3624 0.3695 0.3649 0.0033

300 0.4402 0.4397 0.4384 0.4394 0.0008

Figure A1: FTIR calibration curve of mNA with Spectrum-BX spectrometer

y = 0.0015x - 0.0051

R² = 0.997

0.00

0.10

0.20

0.30

0.40

0.50

0 50 100 150 200 250 300 350

Co

mb

ind

Mon

om

er a

nd

Dim

er

Pea

k H

eigh

t, (

a.u

)

mNA Concentration in dichloromethane, (mg/L)

112

Table A2: mNA calibration data for Spectrum-One Perkin Elmer FTIR spectrometer

Combined Monomer and

Dimer Peak Height, (a.u.)

Concentration,

(mg/L) Run 1 Run 2 Standard Deviation

50 0.0490 0.0463 0.0013

100 0.1523 0.1484 0.0019

150 0.2060 0.2101 0.0021

200 0.2700 0.2697 0.0002

250 0.3424 0.3416 0.0004

Figure A2: FTIR calibration curve of mNA with Spectrum-One spectrometer

y = 0.0015x - 0.0087

R² = 0.989

0

0.1

0.2

0.3

0.4

0 50 100 150 200 250 300

Com

bin

d M

on

om

er a

nd

Dim

er

Pea

k H

eigh

t, (

a.u

)

mNA Concentration in dichloromethane, (mg/L)

113

Model naphthenic acid recovery by liquid-liquid extraction

Figure A3: Model naphthenic acid recovery efficiency by liquid-liquid extraction

0

20

40

60

80

100

6.0 8.0 10.0 12.0 14.0

Rec

ov

ery E

ffic

ien

cy, %

Initial mNA Concentration, (mg/L)

Recovery Efficiency = 84

2%

114

APPENDIX B: Nitrogen isotherms for adsorbent pore structure analysis

Figure B-1: 40-point N2 adsorption isotherms of (A) CAC, Steam-ADC-3h, CO2-ADC-6h and

CO2-ADC-9h; and (B) post-treated activated carbon samples (CO2-ADC-9h, CO2-ADC-9h-pAir,

CO2-ADC-9h-pO2 and CO2-ADC-pNH3)

0

75

150

225

300

375

450

0 0.2 0.4 0.6 0.8 1

Volu

me,

(cm

3/g

)

Relative Pressure, P/Po

CAC

Steam-ADC-3h

CO2-ADC-6h

CO2-ADC-9h

A

0

50

100

150

200

250

0 0.2 0.4 0.6 0.8 1

Volu

me, (c

m3/g

)

Relative Pressure, P/Po

CO2-ADC-9h

CO2-ADC-9h-pAIR

CO2-ADC-9h-pO2

CO2-ADC-9h-pNH3

B

115

APPENDIX C: Adsorption Kinetics Data

Table C-1: Effect of shaking speed on model NA adsorption on two different adsorbents

Using CAC

Using CO2-ADC-9h

Shaking Speed

(rpm)

mNA Concentration

after 40 mins (mg/L)

mNA Adsorbed

(%)

mNA Concentration

after 40 mins (mg/L)

mNA Adsorbed

(%)

0 44.87 31.5 44.71 8.3

50 32.15 50.9 40.89 16.1

75 24.90 62.0 n/m n/m

100 16.87 74.3 30.68 37.1

125 12.95 80.2 28.68 41.2

150 7.79 88.1 20.47 58.0

175 1.51 97.7 12.84 73.7

200 1.26 98.1 12.43 74.5

*n/m = not measured.

Conditions: Agitation duration = 40 mins; Solution volume = 40 mL; Solution pH = 8.5; Temperature = 25oC

For CAC: Initial mNA concentration ~66 mg/L; Adsorbent dosage ~20 mg

For CO2-ADC-9h: Initial mNA concentration ~48 mg/L; Adsorbent dosage ~20 mg

116

Table C-2: Time-dependent model NA adsorption data for the ‘Effect of Pore Structure’ study

Using CO2-ADC-6h Using Steam-ADC-3h

Using CAC

Time,

(min)

Aqueous Phase

Concentration,

Ct (mg/L)

Amount

Adsorbed, qt

(mg/g)

Time,

(min)

Aqueous

Phase

Concentration,

Ct (mg/L)

Amount

Adsorbed, qt

(mg/g)

Time,

(min)

Aqueous Phase

Concentration,

Ct (mg/L)

Amount

Adsorbed, qt

(mg/g)

0 91.8 0.0 0 91.8 0.0 0 91.8 0.0

10 56.0 143.4 9 82.1 38.9 9 54.5 149.3

19 48.7 172.3 19 76.9 59.8 19 36.8 220.2

40 47.8 176.2 40 65.0 107.3 40 30.5 245.3

60 43.9 191.9 59 62.1 119.0 59 25.1 266.8

95 41.9 199.6 95 52.1 158.7 98 24.7 268.7

Conditions: Solution volume = 40 mL; Solution pH = 8.5; Temperature = 25oC; Shaking speed = 200 rpm

Initial mNA concentration ~92mg/L; Adsorbent dosage ~10 mg

117

Sample kinetic model fits for ‘Effect of Pore Structure’ study using CO2-ADC-6h data

Figure C-1: Kinetic model fits for mNA adsorption on CO2-ADC-6h: (A) pseudo-2nd

order fit, (B) pseudo-1st order fit, and (C) Elvoich

model fit

y = 0.0048x + 0.0238

R² = 0.9984

0

0.1

0.2

0.3

0.4

0.5

0.6

0 20 40 60 80 100

t/q

t, (

min

g/m

g)

Time, (min)

CO2-ADC-6h pseudo-2nd order fit

Linear (CO2-ADC-6h pseudo-2nd order fit)

y = -0.0089x + 2.0536

R² = 0.7652

0

0.5

1

1.5

2

2.5

0 20 40 60 80 100

log

(qe-q

t)

Time, (min)

CO2-ADC-6h pseudo-1st order fit

Linear (CO2-ADC-6h pseudo-1st order fit)

y = 22.948x + 96.346

R² = 0.9311

0

50

100

150

200

250

0 1 2 3 4 5

qt,

(m

g/g

)

Ln(t)

CO2-ADC-6h Elovich fit Linear (CO2-ADC-6h Elovich fit)

B

C

A

118

Table C-3: Time-dependent model NA adsorption data for the ‘Effect of Surface Functional Groups’ study

Using CO2-ADC-9h

Using CO2-ADC-9h-pAir

Using CO2-ADC-9h-pO2

Time,

(min)

Aqueous Phase

Concentration,

Ct (mg/L)

Amount

Adsorbed, qt

(mg/g)

Time,

(min)

Aqueous Phase

Concentration,

Ct (mg/L)

Amount

Adsorbed, qt

(mg/g)

Time,

(min)

Aqueous Phase

Concentration,

Ct (mg/L)

Amount

Adsorbed, qt

(mg/g)

0 33.2 0.0 0 33.2 0.0 0 33.2 0.0

9 27.0 24.9 10 27.9 21.1 10 29.5 14.7

24 21.4 46.0 25 25.8 30.7 25 26.2 22.7

48 19.3 55.2 50 22.1 44.3 49 27.4 28.0

79 12.6 82.4 81 22.0 44.3 79 25.7 29.7

125 10.4 88.9 124 20.5 51.3 125 25.6 30.6

Conditions: Solution volume = 40 mL; Solution pH = 8.5; Temperature = 25oC; Shaking speed = 200 rpm

Initial mNA concentration ~33mg/L; Adsorbent dosage ~10 mg

119

Table C-4: Time-dependent Adsorption data for the ‘Effect of Initial Adsorbate Concentration’ study with CO2-ADC-6h

Initial mNA Concentrations

31 mg/L 51 mg/L 175 mg/L

Time,

(min)

Amount Adsorbed,

qt (mg/g)

Time,

(min)

Amount Adsorbed,

qt (mg/g)

Time,

(min)

Amount Adsorbed,

qt (mg/g)

0.0 0.00 0.0 0.00 0.0 0.00

9.5 19.81 9.0 24.51 9.5 48.09

23.0 26.38 24.5 63.30 28.0 80.22

57.0 46.54 52.0 70.87 52.0 81.47

81.0 58.74 80.0 82.07 81.0 96.21

127.0 72.46 127.0 89.90 127.0 103.37

Conditions: Solution volume = 40 mL; pH = 8.5; Temperature = 25oC.

Shaking speed = 200 rpm; CO2-ADC-6h dosage ~10 mg.

120

Table C-5: Adsorption data for the ‘Effect of Temperature’ on adsorption kinetics

Temperature, oC

25 oC 50

oC 75

oC

Time,

(min)

Amount Adsorbed,

qt (mg/g)

Time,

(min)

Amount Adsorbed,

qt (mg/g)

Time,

(min)

Amount Adsorbed,

qt (mg/g)

0 0.0 0 0.0 0 0.0

9 21.7 9 29.3 9 27.1

25 44.4 25 45.1 25 51.2

56 57.4 55 60.6 55 62.8

Conditions: Solution volume = 40 mL; pH = 8.5; Shaking speed = 200 rpm.

CO2-ADC-9h dosage ~10 mg; Initial mNA concentration ~ 51 mg/L

121

APPENDIX D: Adsorption Isotherm Data

Table D-1: Equilibrium adsorption data for the ‘Effect of Pore Structure’ study

Activated Carbon Adsorbents

CO2-ADC-6h CAC Steam-ADC-3h

Equilibrium

Solution Conc.,

Ce (mg/L)

Equilibrium

Adsorbed Quantity,

qe (mg/g)

Equilibrium

Solution Conc.,

Ce (mg/L)

Equilibrium

Adsorbed Quantity,

qe (mg/g)

Equilibrium

Solution Conc.,

Ce (mg/L)

Equilibrium

Adsorbed Quantity,

qe (mg/g)

29.8 126.3 25.4 301.1 41.9 84.0

27.8 100.1 17.3 291.2 39.2 81.8

26.9 85.9 11.1 280.1 34.9 57.1

24.9 73.4 7.3 245.7 30.4 48.5

23.0 67.6 1.1 190.2 22.8 43.1

21.3 51.6 0.9 143.5 14.6 29.7

Conditions: Solution volume = 40 mL; pH = 8.5; Temperature = 25oC.

Shaking speed = 200 rpm.

Initial mNA concentration ~44 mg/L; Adsorbent dosage range ~2-40 mg.

122

Sample isotherm model fits for ‘Effect of Pore Structure’ study using CO2-ADC-6h data

Figure D-1: Isotherm model fits for mNA adsorption on CO2-ADC-6h: (A) Langmuir fit, (B) Freundlich fit, and (C) Temkin fit

y = -0.0187x + 0.799

R² = 0.926

0.1

0.2

0.3

0.4

0.5

20 22 24 26 28 30 32

Ce/

qe,

(g

/L)

Ce, (mg/L)

CO2-ADC-6h Langmuir fit

Linear (CO2-ADC-6h Langmuir fit)

y = 2.488x - 3.660

R² = 0.967

3.0

3.5

4.0

4.5

5.0

3.00 3.10 3.20 3.30 3.40 3.50

Ln

(q

e)

Ln (Ce)

CO2-ADC-6h Freundlich fit

Linear (CO2-ADC-6h Freundlich fit)

y = 204.7x - 5789

R² = 0.925

0

40

80

120

160

3.00 3.10 3.20 3.30 3.40 3.50

qe,

(m

g/g

)

Ln (Ce)

CO2-ADC-6h Temkin fit Linear (CO2-ADC-6h Temkin fit)

A B

C

123

Table D-2: Equilibrium adsorption data for the ‘Effect of Surface Functional Groups’ study

Activated Carbon Adsorbents

CO2-ADC-9h CO2-ADC-9h-pAir CO2-ADC-9h-pO2 CO2-ADC-9h-pNH3

Equilibrium

Solution

Conc., Ce

(mg/L)

Equilibrium

Adsorbed

Quantity,

qe (mg/g)

Equilibrium

Solution

Conc., Ce

(mg/L)

Equilibrium

Adsorbed

Quantity,

qe (mg/g)

Equilibrium

Solution

Conc., Ce

(mg/L)

Equilibrium

Adsorbed

Quantity,

qe (mg/g)

Equilibrium

Solution

Conc., Ce

(mg/L)

Equilibrium

Adsorbed

Quantity,

qe (mg/g)

20.9 66.7 27.6 60.4 29.2 41.1 23.0 111.7

17.8 61.6 23.6 53.1 27.1 39.1 18.9 102.0

14.5 57.6 19.5 51.8 23.3 37.0 13.5 78.1

11.1 56.7 16.5 47.6 20.9 35.0 8.8 70.0

5.7 51.4 14.2 40.2 18.6 32.8 6.4 67.8

Conditions: Solution volume = 40 mL; pH = 8.5; Temperature = 25oC.

Shaking speed = 200 rpm.

Initial mNA concentration ~33 mg/L; Adsorbent dosage range ~3-22 mg.

124

Table D-3: Equilibrium adsorption data for TAOs adsorption from real tailings pond water

Equilibrium Solution Conc., Ce (mg/L) Equilibrium Adsorbed Quantity, qe (mg/g)

22.8 28.8

19.0 28.0

18.2 25.2

16.6 24.1

14.4 22.7

Conditions: TAOs conc. ~ 25 mg/L. pH = 7.6

Solution volume = 40 mL; Temperature = 25oC.

Shaking speed = 200 rpm; Adsorbent dosage range ~3-22 mg.