255
DEVELOPMENT OF NOVEL OIL RECOVERY METHODS FOR PETROLEUM REFINERY OILY SLUDGE TREATMENT by Guangji Hu B.Eng., South Central University for Nationalities, 2007 M.Sc ., Wuhan University, 2011 THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY IN NATURAL RESOURCES AND ENVIRONMENTAL STUDIES UNIVERSITY OF NORTHERN BRITISH COLUMBIA January 2016 © Guangji Hu, 2016 1

DEVELOPMENT OF NOVEL OIL RECOVERY METHODS FOR … · DEVELOPMENT OF NOVEL OIL RECOVERY METHODS FOR PETROLEUM REFINERY OILY SLUDGE TREATMENT by Guangji Hu B.Eng., South Central University

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

DEVELOPMENT OF NOVEL OIL RECOVERY METHODS FOR PETROLEUM REFINERY OILY SLUDGE TREATMENT

by

Guangji Hu

B.Eng. , South Central University for Nationalities, 2007 M.Sc., Wuhan University, 2011

THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY IN

NATURAL RESOURCES AND ENVIRONMENT AL STUDIES

UNIVERSITY OF NORTHERN BRITISH COLUMBIA

January 2016

© Guangji Hu, 2016

1

ABSTRACT

Oily sludge is one of the most significant wastes generated in the petroleum industry.

It is a complex emulsion of various petroleum hydrocarbons (PHCs), water, metals, and

fine solids. Due to its hazardous nature and increased generation quantities around the

world, the effective treatment of oily sludge has attracted widespread attention. The

complexity of its composition and diversity of its origin sources make oily sludge

management a difficult and costly undertaking. Many methods have been developed for

the treatment of oily sludge through oil recovery or sludge disposal approaches, but no

single specific process can be considered as a panacea since each method is associated

with different advantages and limitations. Efforts should focus on the improvement of

current technologies and the combination of oil recovery with sludge disposal in order to

comply with both resource reuse recommendations and environmental regulations. The

object of this study was to develop novel combined methods for oil recovery treatment on

different refinery oily sludges. The investigation focused on the oil recovery performance

of combined methods based on four individual treatment processes including ultrasonic

irradiation, solvent extraction, freeze/thaw, and pyrolysis in oily sludge treatment.

Firstly, the oil recovery and salt removal effects of ultrasonic irradiation on oil

refinery tank bottom sludge were investigated, together with those of direct heating.

Ultrasonic power, treatment duration, sludge-to-water ratio, and initial sludge-water slurry

temperature were examined for their impacts on sludge treatment. It was found that the

increased initial slurry temperature could enhance the ultrasonic irradiation performance,

especially at lower ultrasonic power level (i.e., 21 W), but the application of higher-power

ultrasonic irradiation could rapidly increase the bulk temperature of slurry. Ultrasonic

2

irradiation had a better oil recovery and salt removal performance than direct heating

treatment. More than 60% of PH Cs in the sludge was recovered at an ultrasonic power of

75 W, a treatment duration of 6 min, an initial slurry temperature of 25 °C, and a sludge-

to-water ratio of 1 :4, while salt content in the recovered oil was reduced to <5 mg/L,

thereby satisfying the salt requirement in refinery feedstock oil. In general, ultrasonic

irradiation could be an effective method in terms of oil recovery and salt removal from

refinery oily sludge, but the separated wastewater still contains relatively high

concentrations of PH Cs and salt which requires proper treatment.

Two types of ultrasonic assisted extraction (UAE) treatment including UAE probe

(UAEP) system and UAE bath (UAEB) system were investigated for oil recovery from

refinery oily sludge. Their oil recovery efficiencies were compared to that of mechanical

shaking extraction (MSE). Three solvents including cyclohexane (CHX), ethyl acetate

(EA), and methyl ethyl ketone (MEK) were examined as the extraction solvents. The

influence of experimental factors on oil and solvent recovery was investigated using an

orthogonal experimental design. Results indicated that solvent type, solvent-to-sludge (S/S)

ratio, and treatment duration could have significant effects on oil recovery in UAE

treatment. Under the optimum conditions, UAEP treatment can obtain an oil recovery of

68.8% within 20 s, which was higher than that (i.e., 63.7%) by MSE treatments after 60

mins ' extraction. UAEB treatment can also obtain a promising oil recovery within a shorter

extraction duration (i.e. , 15 min) than MSE. The experimental results indicated that two

solvent extraction cycles on oily sludge were sufficient to obtain a satisfactory oil recovery

for all three extraction treatments. The recovered oil by CHX contained the lowest total

petroleum hydrocarbon (TPH) content (i.e., about 50%), while the recovered oil by EA

3

and MEK had relatively higher TPH content (i.e., >80%). UAE was thus illustrated as an

effective and improved approach for oily sludge waste recycling.

A combination of solvent extraction and freeze thaw was examined for recovering oil

from the high-moisture petroleum refinery wastewater treatment pond sludge. Five

solvents including cyclohexane (CHX), dichloromethane (DCM), methyl ethyl ketone

(MEK), ethyl acetate (EA), and 2-propanol (2-Pro) were examined. It was found that these

solvents except for 2-Pro showed a promising oil recovery rate of about 40%, but the

recycling of DCM solvent after oil extraction was quite low. As a result, three solvents

(CHX, MEK and EA) were selected for further combination with freeze/thaw treatment to

improve the quality of recovered oil in terms of its total petroleum hydrocarbon (TPH)

content. The results indicated that the freeze/thaw treatment after solvent extraction could

increase the TPH content in recovered oil from about 40% to 60% for both MEK and EA

extractions, but little effect was observed for CHX extraction. Although the solid residue

after oil recovery had a significantly decreased TPH content, a high concentration of heavy

metals was observed, indicating that this residue may require proper management. In

general, the combination of solvent extraction with freeze/thaw is effective for high-

moisture oily hazardous waste treatment.

The treatment ofrefinery oily sludge through co-pyrolysis with wood waste (sawdust)

was carried out in a fixed-bed reactor. Response surface method (RSM) was applied to

evaluate the main and interaction effect of three experimental factors (i.e., sawdust

percentage in feedstock, temperature, and heating rate) on pyrolysis oil and char yields.

The oil quality in terms of elemental analysis, moisture content, and higher heating value

(HHV) was also investigated. A synergistic effect of co-pyrolysis was found, and the oil

and char yields increased with the sawdust percentage in feedstock. The interaction

4

between heating rate and sawdust percentage as well as between heating rate and

temperature was significant on the yield of pyrolysis oil. The HHV of oil originated from

sawdust increased by 5 MJ/kg due to co-pyrolysis at a sawdust/oily sludge ratio of 3: 1 as

compared to that of sawdust pyrolysis alone. The results indicated that the carbon content

of char increased as increasing sawdust percentage in feedstock. In general, refinery oily

sludge can be used as an additive in the pyrolysis of sawdust for improving the yield and

quality of oil products.

The results of this research indicate that the combined oil recovery methods have the

potential to be applied for the treatment of different complex oily wastes in petroleum

refining industries to meet sustainable development principles.

5

TABLE OF CONTENTS

ABSTRACT .. .. ... ... ....... ..... ... ...... ....... ... ... .. .... ... ...... .. .... ... .... ... .... ... ............. ............. .... ....... 2

LIST OF TABLES ........................ .... .... .. ............ .. ...................... .... ..................... ......... .... 10

LIST OF FIGURES ...................... .... .. .... ......................................................... ... ... .... ....... 12

GLOSSARY ..... .... .. ......... ..... .. .............. .... .. .. ....... ... .... ......... .. ........ .... ......... .... ........ .... ...... 18

ACKNOWLEDGEMENT .... ... .. .. ... ....... .......... .......................... ...... ..... ....... ..... .......... ..... . 21

Chapter 1 General Introduction ........... ...................................... ................................. ...... 22

1.1 Background .. .................... ......... .. .......... .... ................. .......... .............. .... .. ............ ... 22

1.2 Statement of the problem and research objectives ............................... .... .......... .... . 23

Chapter 2 Literature Review * ......................... .... ........... ....................... ...... ...... ............... 29

2.1 . Introduction ..... .......................................... ......... ....... ..................... ....... ... .... ..... .. ... 29

2.2 Source, characteristics and toxicity of oily sludge ..... ... ... .. ........ .... ........ ... ........ ... ... 30

2.2.1 Oily sludge source ........... ..... ............... .... ......... .. .... ... ........... .......................... .. 30

2.2.2 Characteristics of oily sludge ... ........ ............. ......... .. ..... ... .......... .. .......... .. ...... .. 32

2.2.3 Toxicity and impact of oily sludge ....... .... ............... ..... ....... ..... .. .......... ... ..... .... 35

2.3 Overview of sludge treatment methods .. ... ................... .. ...... ............ ..... ................. 37

2.4 Oil recovery methods for oily sludge treatment.. ......... ...... ................ ..... ..... ........... 38

2.4.1 Solvent extraction .. .... .... .............. .. ... .. ... ..... .. ..... ........... ...... ...... ..... ...... ....... ... .. . 38

2.4.2 Centrifugation treatment. ... ....... ............ ...... ..... ... ......... ..... .......... ................ ..... . 41

2.4.3 Surfactant enhanced oil recovery (EOR) .. ..... .... ...... ... ........ ...... ........... ......... ... . 44

2.4.4 Freeze/thaw treatment ............... ..... .. ...... .. ... ........ .... ....... ..................... ............. 47

2.4.5 Pyrolysis treatment ..... .. .... ....... .... ......... ..... ....... ......... ........ .... ... ................. ....... 50

2.4.6 Microwave irradiation ... ......... ....... ..... ................... .... ................ ..... .. ..... ..... ...... 53

2.4. 7 Electrokinetic method ..... ... ....... ..... .. .. ... ... ... ... .. ..... ..... ...... ........... .. .. ...... ............ 56

2.4.8 Ultrasonic irradiation ...... ............... ................ ...... ... ................. ......... ... ... .......... 58

2.4.9 Froth flotation ... .................. ........... ..... ............... .......... ..... .. ... ....... .. ........... .... ... 61

2.5 Oily Sludge Disposal Methods ... .......... ... .. ........ ...... ........... ............. ... .... ......... ... .... 63

2.5.1 Incineration ........... ...................... .. ... .. ... ............ .. ... ................. ....... .. ....... .. ... ..... 63

2.5.2 Stabilization/solidification .................... ..... ... ... .... ........ ......................... ............ 65

2.5.3 Oxidation treatment ....... ................... ..... ... ....... ........ ......... ..... .. .................. ....... 67

2.5.4 Bioremediation ........... ................ .. .... ... ............ ................ ... ........ ....... ... .......... .. 70

6

2.6 Discussion and conclusion ......... ... ............................. .. .................................. .. ... .... 78

Chapter 3 Ultrasonic oil recovery and salt removal from refinery tank bottom sludge * . 84

3 .1 Introduction ............................................................................................................. 84

3.2 Experimental materials and methods ...................................................................... 87

3.2.1 Materials .................... ....... .... .. ........ ... .......... ...... ... .................. .......................... 87

3.2.2 Methods .............. .......... .... .......... ... ... ....... ...... ...... ........... ....... ... ... .. .. .. .. .... ......... 88

3.2.3 Sample analysis ................. ..... .......... ....... .... ....... ......... ..... ............ ... ... ... .... ...... . 90

3.3 Results and discussion .......... .......... ....... ... .................. ....... .. .................. ................. 93

3. 3 .1 Temperature change during ultrasonic treatment.. ........ ........................... ...... .. 93

3.3.2 Effect of temperature and ultrasonic irradiation ............ ................... ............ .... 95

3.3.3 Effects of ultrasonic power and treatment duration ......................................... 99

3 .3 .4 Effect of sludge to water ratio .. ... ......... .. ...................... .. ................................ 106

3 .3 .5 Characteristics of recovered oil ............................................................... ....... 109

3.4 Conclusions .............. .. ... ....... ..... .. ..... .. ........... ..... ..... .............. ......... ........... ............ 112

Chapter 4 Oil recovery from refinery oily sludge through ultrasonic assisted extraction * .. .. ..... ... .............................................. .... ......................... ............................... .......... ...... .. 113

4.1 Introduction .......................... ........... .................................. ........ ............................ 113

4.2 Experimental materials and methods ...... .................. ..... ..... ... ....... ............... .. ....... 115

4.2.1 Materials and reagents ...... .. ..... .......... ...... ...... ........... ........................ ..... ......... 115

4.2.2 Methods .. .... ....... .. ........ ....... ............................ ..... ............ .......... .......... ........... 117

4.3 Results and discussion ................ ..... .... ......... ......... .............. .... ... .. ............. ...... ..... 123

4.3.1 The effect of mechanical shaking extraction (MSE) ... .... .. ... ......... .. ....... .. .. .... 123

4.3.2 Ultrasonic assisted extraction using a probe system (UAEP) ........... ............. 128

4.3.3 Ultrasonic assisted extraction using a bath system (UAEB) ....................... .. . 133

4.4.4 Impact of extraction cycles .... ... ...................................................................... 138

4.3.5 Characteristics ofrecovered oil.. ........ ....................... ........................... ... ....... 141

4.4 Conclusions ........................................................................................................... 148

Chapter 5 A combination of solvent extraction and freeze thaw for oil recovery from refinery wastewater treatment pond sludge * .................................................................. 150

5 .1 Introduction ........................................................................................................... 15 0

5.2 Experimental materials and methods .. .... ... ....... .................................................... 153

5.2.1 Materials .... .... ......................... ... ... .. ......... ... ............. .. .... ... ................ ..... .... .... . 153

5.2.2 Solvent extraction ........................................................................................... 155

7

5.2.3 Freeze/thaw treatment ........ .............. .. ... ....... ...... ... ... .... .... .... ...... .................. .. 157

5.2.4 Sample analysis ..... ............. ... ..... ............ ..... ... ... ..... .... .. .... ... ........ .................. . 159

5.3 Results and discussion ... ......... ...... ................. ........................................ ....... ........ 160

5.3.1 Impact of solvent-to-sludge (S/S) ratio .. ......................... ... ........ .... ............... . 160

5 .3 .2 Impact of extraction duration ......... ............. ... ... .. ......... ........................ ....... ... 165

5.3.3 Extraction cycle .... ..... .... .... ..... ... ................................ .... ...... .. ...... ................... 168

5.3.4 Freeze/thaw enhancement ... .... ... ... .... ...... ........ .......... ...................... ...... ... ...... 169

5.3.5 Characteristics ofrecovered oil and solid residue ........................ ............ ...... 174

5.4 Conclusions ........... ................... .. .. ................ ... .......... ............ ...... ... ........... ............ 179

Chapter 6 Treatment of refinery oily sludge through co-pyrolysis with wood waste ... . 180

6.1 Introduction ............ ........................ ... .......... .... .......... ..... .... ... ....... ...... ..... ... ... ..... ... 180

6.2 Experiment ....................... ..... ...... ................ .... ..... .... ............... .............. ....... ......... 183

6.2.1 Materials .. ..... ......... ... ... ......... .. ...... .............. ....... ... ............ ...... ... ..................... 183

6.2.2 Experimental design ........ ... ... ................ .. .... .. ..... ..................... .... ....... ... ..... ... . 185

6.2.3 Pyrolysis procedure ........... .. ........ .. ....... .. ... .... ......... .. ........ .................. ............ 186

6.2.4 Sample analysis ..... ... ... ........ ............ ... ......... ... ................. .... .... ... .................... 188

6.3 Results and discussion ....... ....... ..... .... ........................................ .......... ... ... ......... .. 189

6.3.1 Thermal gravimetric analysis (TGA) .......... ....................... .. ..... ..... ... ....... ...... 189

6.3.2 Pyrolysis oil yield .... ......... .................. .. .......... .. ............................. ................ . 191

6.3.3 Pyrolysis char yield ........... .. ................... ........... ... ................. .......... .. ... .... ... ... 198

6.3.4 Synergistic effect of co-pyrolysis on oil yield ........ ....................... ... .. ......... ... 201

6.3.5 Pyrolysis product characterization .......................................... ......... .. .. .......... 203

6.4 Conclusions ..... ..... ....... .... .......... ... .... ..... .................................. ..... ..... ......... ...... ... .. 206

Chapter 7 Conclusions and future research ........................... ... .......... .. ........ .............. .... 208

7.1 Thesis conclusions .... ..... ...... ................. ... ........ ................................. .................... 208

7.1.1 The oil recovery and desalting effect of ultrasonic irradiation on oily sludge208

7 .1.2 The combined effect of ultrasonic irradiation and solvent extraction on oil recovery ............. .................. .. .. .. ....... ... .... .. ... ... .................................... .......... .......... 210

7 .1.3 The combined effect of solvent extraction and freeze/thaw on oil recovery . 212

7.1.4 The co-pyrolysis of sawdust with oily sludge for pyrolysis oil production ... 214

7.2 Research achievements ......... ..... .... ......... ....... ......... .... ................ ... ..... .... .............. 217

7.3 Future research ........... ..... .... ...... .... ............. ... ... ... ...... .... .......................... ..... ......... 218

References ........... ... ....... ...... ..... .... .. ...... .. .................. ... ....... ...... ................. ... ....... ............ 220

8

APPENDIX .. .. ... .. .. ... ....... ..... ..... ...... ...... .................. ... .......... ................. .................... ..... 255

Appendix I ............ ........ .. ...... ............ ..... ...... .. .......... .......... ... .................. ...... ... ... .... ... . 255

Appendix II ... ... ... ... .. .... ............... ........ ..................... ... .. Error! Bookmark not defined.

9

LIST OF TABLES

Chapter 2

Table 2.1 Summary and comparison of oil recovery methods ... ......... .................. ......... . 81

Table 2.2 Summary and comparison of oily sludge disposal methods .... ...................... .. 82

Chapter 3

Table 3.1 Characteristics of tank bottom oily sludge sample for ultrasonic treatment ... . 87

Table 3.2 Experimental factors and values of ultrasonic irradiation treatment.. .............. 90

Chapter 4

Table 4.1 Properties of tank bottom oily sludge for extraction treatment.. ......... .... ...... . 116

Table 4.2 Impact factors and levels of each factor in three extraction processes for the

orthogonal experiments .................... .... .... .......... .. ...... .. ....... ............ .... ....... .... ... ... ... 11 7

Table 4.3 A L9 (34) orthogonal array of factors and levels for each extraction process . 118

Table 4.4 Single factor experimental conditions of the three extraction processes .... .... 119

Table 4.5 MSE oil recovery results and statistical analysis .... ..... .... ..... ... ......... ...... ........ 124

Table 4.6 ANOV A for oil and solvent recovery results from MSE ... ..... ... ..... .... ....... .. .. 125

Table 4.7 UAEP oil recovery results and statistical analysis ............ .. ....... .. .. ...... .. ... ... .. 129

Table 4.8 ANOVA for oil and solvent recovery results from UAEP ........... ......... ....... .. 130

Table 4.9 UAEB oil recovery results and statistical analysis ........... ... ..... ... ......... ..... ..... 134

Table 4.10 ANOVA of oil and solvent recovery results from UAEB ... ................ ......... 135

Table 4.11 Properties of crude oil and recovered oil (RO) by different solvents in UAEP

.. .. .... ....................... ....... ...... ...... ...... ...... ... .... .... ... .......... ................. .... ... ... ... .. .... .. ..... 147

10

Chapter 5

Table 5.1 Properties of petroleum refinery wastewater treatment pond sludge ............ . 154

Table 5.2 Experimental factors and values of solvent extraction ............ .. ... .............. ... . 156

Table 5.3 Properties of the original oily sludge and the solid residues after combined

solvent extraction and freeze/thaw treatment ..... ..... ............. ... .. .. ........ .... ....... 178

Chapter 6

Table 6.1 Properties of sawdust and oily sludge .... ........ ..... .............. ................... .. .... .. .. 184

Table 6.2 Experimental array of CCC design for co-pyrolysis experiments and product

yield results ..... .... ... .. .. ..... ........... .. ....... ...... .................. .. .. .. .. .. ... ... .. ... ... ...... ... .. 193

Table 6.3 Statistical parameters obtained from the analysis of variance for the model . 194

Table 6.4 Observed product yields from co-pyrolysis experiments as compared with the

predicted values (temperature: 500 °C, heating rate: 20 °C/min) .. ...... .......... . 203

Table 6.5 Properties of pyrolysis products from co-pyrolysis experiments (heating rate:

20 °C /min, temperature: 500 °C) ....... ......... .... ............................................... 205

11

LIST OF FIGURES

Chapter 2

Figure 2.1 Worldwide daily refining throughputs in recent years .............. .... .......... .. ...... 32

Figure 2.2 Overview of oily sludge treatment methods .. .. ............ .............. ...... .. .. .... ...... .. 37

Figure 2.3 Flow diagram of solvent extraction process (1 : reactor column; 2: distillation

system; 3: solvent recycling tank; 4: compressor and cooling system) ........... 41

Figure 2.4 Schematic view of centrifugation used in oily sludge treatment.. ................... 44

Figure 2.5 Schematic diagram of the mechanism of freeze/thaw-induced demulsification

for W/0 emulsion: (a) original emulsion, (b) water droplets freezing,

expansion and coalescing, ( c) oil phase freezing to form a solid cage, ( d) water

droplets freezing and expanding to break the cage, ( e) emulsion thawing and

water droplets coalescing, and (f) gravitational delamination .... .. ................... 48

Figure 2.6 Typical schematic view of fluidized bed systems used in sludge pyrolysis

treatment. ................................................ .. ........................................................ 52

Figure 2. 7 Schematic view of froth flotation in oil/water separation process (A: Air

control valve; B: Motor; C: Stirrer; D: Skimmed oil collector) ....................... 62

Chapter 3

Figure 3.1 Tank bottom sludge sample ... .. .... .. .......... .. ............................ .. .... .. .. .. .. ....... .... . 88

Figure 3.2 Schematic diagram of the ultrasonic irradiation set-up for oily sludge

treatment. ... ........................ ................ ...... ........... ....................... .. ....... .. ...... ... ... 90

Figure 3.3 Temperature change in oily sludge-water slurry mixture treated by ultrasonic

irradiation ( experimental condition: sludge to water ratio of 1 :4 and initial

slurry temperature of 25°C) ... ... ........ .. .... ................. ...... .. ............ ..... .. .... .... ... .. 94

12

Figure 3 .4 Three layers separated from ultrasonic irradiation treatment on oily sludge .. 94

Figure 3.5 Effect of direct heating on oily sludge treatment at different temperatures, (a)

oil recovery rate and salt content in recovered oil, (b) TPH and salt

concentration in wastewater ( error bar represents standard deviation) ............ 96

Figure 3.6 Effect of ultrasonic irradiation on oily sludge treatment at different initial

slurry temperatures, (a) oil recovery rate, (b) salt content in recovered oil, (c)

TPH concentration in wastewater, and ( d) salt content in wastewater

( experimental condition: treatment duration of 6 min and sludge to water ratio

of 1 :4; error bar represents standard deviation) ... ..... .. ............ ......................... 98

Figure 3. 7 Effect of ultrasonic irradiation duration on oily sludge treatment, (a) oil

recovery rate, (b) salt content in recovered oil, ( c) TPH concentration in

wastewater, and ( d) salt content in wastewater ( experimental condition: sludge

to water ratio of 1 :4 and initial slurry temperature of 25°C; error bar represents

standard deviation) ................. ..... ........ ....... ..... ...................... .... ... ..... .. ........... 100

Figure 3.8 Mixtures of oily sludge and water, (a) before ultrasonic treatment, (b) after

treatment under ultrasonic power of 21 W and duration of 2 min, ( c) after

treatment under ultrasonic power of 75 W and duration of 6 min ( experimental

condition: sludge to wat er ratio of 1 :4 and initial slurry temperature of 25°C)

..... .... ...................... ... .. ..... ....... ....... ............. ...... .. .... .. ....... ........ .... .. .. .. ... .......... 101

Figure 3.9 Wastewater generated from oil recovery treatment using different ultrasonic

irradiation powers and durations ......... .. .......... .... ....... ...... ... ...................... ..... 104

Figure 3.10 Effect of sludge to water ratio on the performance of oily sludge treatment by

ultrasonic irradiation, ( a) oil recovery rate, (b) salt content in recovered oil, ( c)

TPH concentration in wastewater, and ( d) salt content in wastewater

13

( experimental condition: treatment duration of 6 min and initial slurry

temperature of 25°C; error bar represents standard deviation) ........... ........... 108

Figure 3.11 Distribution of PHCs fractions and TPH in different hydrocarbon samples,

(a) PHCs fraction percentage, (b) TPH percentage (error bar represents

standard deviation) ............................... .... ........ ........ ... ... .... ............. .. .. ........... 111

Chapter 4

Figure 4 .1 Experimental setups of the three different extraction processes ........... ... ..... 121

Figure 4.2 Influence of experimental factors on the (a) oil and (b) solvent recovery in

MSE process ...... ....... ... .. .. ... ............... ... .. ...... .. ............ ....... ...... .. ...... ...... ..... .. . 124

Figure 4.3 Results of single-factor experiments of MSE ....... ...... ............. ... ... ............... 127

Figure 4.4 Influence of experimental factors on the (a) oil and (b) solvent recovery in

UAEP process .... .. .. ... .................... ............ .. .... .............. ....... ..... ... .. ........ ...... .. 129

Figure 4.5 Results of single-factor experiments of UAEP .. .......... ... .... ... ... ... ... .... ........ .. 132

Figure 4.6 Influence of experimental factors on the (a) oil and (b) solvent recovery in

UAEB process ..... .... .. ...................................... ..... .. .... ................. ......... ......... . 134

Figure 4.7 Results of single-factor experiments ofUAEB ............ ..... .... .. ... ................... 138

Figure 4.8 Effect of extraction cycles on oil recovery in (a) MSE, (b) UAEP, and (c)

UAEB treatment ...... ..... ... .. .......... .. ... .. .......... .................... ...... ...... ................ .. 140

Figure 4.9 Recovered oils by different solvents from UAEP treatment ..... .... .. .. ............ 143

Figure 4.10 TPH recovery (a) and TPH content in recovered oil (b) of three extraction

processes (MSE: SIS ratio of 4: 1, shaking speed of 250 rpm, and duration of

60 mins; UAEP: SIS ratio of 4:1, ultrasonic power of21 W, and duration of20

14

s; UAEB: S/S ratio of 4: 1, duration of 15 min, and bath temperature of 20 °C)

........... ... ....... ... ....... ..... .... ................................................................................ 144

Figure 4.11 Distribution of PH Cs fractions in the recovered oil by using three solvents in

(a) MSE, (b) UAEP, and (c) UAEB (MSE: S/S ratio of 4:1, shaking speed of

250 rpm, and duration of 60 mins; UAEP: S/S ratio of 4: 1, ultrasonic power of

21 W, and duration of 20 s; UAEB: S/S ratio of 4: 1, duration of 15 min, and

bath temperature of 20 °C) ............. .................. ........ ... ...... .................. .. .... ..... 146

Chapter 5

Figure 5.1 Dredged oily sludge from a petroleum refinery wastewater treatment pond 154

Figure 5.2 Freeze treatment on different extractants at -20 °Cina fridge .... ............. .... 159

Figure 5.3 Flow chart of combined solvent extraction and freeze/thaw treatment ........ 159

Figure 5.4 Effect of solvent to sludge ratio on solvent extraction performance, (a) oil

recovery rate, (b) solvent recovery rate, ( c) waste reduction rate ( experimental

condition: extraction duration of 120 min at 25 °C; error bar represents

standard deviation, n=3) ............ .. .. ............... ........... ... ......................... .... ....... 163

Figure 5.5 Extractant of extraction treatments using different solvents ....... .................. 164

Figure 5.6 Effect of extraction duration on solvent extraction performance, (a) oil

recovery rate, (b) solvent recovery rate, ( c) waste reduction rate ( experimental

condition: solvent-to-sludge ratio (v/m) of 4: 1 at 25 °C; error bar represents

standard deviation, n=3) ... ... ....... .................. ..... ....... .. ... ...... ......... .......... ........ 168

Figure 5. 7 Effect of solvent extraction cycles on oil recovery rate ........... .......... ........... 169

Figure 5.8 Effect of freeze/thaw enhanced solvent extraction and the comparison of

separated water level from the extractant with freeze/thaw treatment (left

15

yellow bar) and without freeze/thaw treatment (right red bar), (a) ice lattice

formation in the extractant from MEK extraction, (b) CHX extraction, ( c)

MEK extraction, and ( d) EA extraction ........................................... .... .... ...... 171

Figure 5.9 Effect of freeze/thaw (FIT) treatment on (a) TPH content in recovered oil by

solvent extraction and (b) TPH content in recovered oil as a function ofF/T

cycle ........ ..... .... .. .. ... ... .. .................. ..... ......................... .. ...... .... .. .................... 174

Figure 5.10 Distribution of PH Cs fractions in fresh crude oil and the recovered oil by

using the combination of solvent extraction with freeze/thaw ....................... 175

Figure 5 .11 Solid residues from refinery wastewater pond sludge after combined solvent

extraction and freeze/thaw treatment ............................................................. 1 77

Chapter 6

Figure 6.1 The feedstock of sawdust and oily sludge (left: before mixed; right after

mixed) ............................... ................... ........ ... .. ....... ... ........ .............. ........ ... 184

Figure 6.2 The fixed-bed reactor designed for the co-pyrolysis of sawdust with oily

sludge ................................. ... .. ..................................................................... 187

Figure 6.3 Schematic diagram of fixed-bed pyrolysis reactor.. ...................................... 188

Figure 6.4 TG and DTG curves of (a) sawdust and (b) oily sludge (solid line: TG; dashed

line: DTG) .................................................................................................... 190

Figure 6.5 Liquid oil product from the pyrolysis of sawdust alone (left), mixture of

sawdust (75 wt.%) and oily sludge (25 wt.%), and oily sludge alone (right)

(Heating rate: 20 °C/min; Temperature: 500 °C) .......... ................ ..... .......... 192

Figure 6.6 The main effects of single factors on pyrolysis oil yield: (a) sawdust

percentage (temperature: 500 °C, heating rate: 12.5 °C/min), (b) temperature

16

(sawdust percentage: 50 wt.%, heating rate: 12.5 °C/min), and (c) heating

rate (sawdust percentage: 50 wt.%, temperature: 500 °C) ........................... 197

Figure 6. 7 The interaction effect of experimental factors on pyrolysis oil yield ........... 198

Figure 6.8 The main effects of single factors on char yield: (a) sawdust percentage

(temperature: 500 °C, heating rate: 12.5 °C/min), (b) temperature (sawdust

percentage: 50 wt.%, heating rate: 12.5 °C/min), and (c) heating rate

(sawdust percentage: 50 wt.%, temperature: 500 °C) ............. ........ ... ... .... ... 200

Figure 6.9 The interaction effect of experimental factors on char yield .................... .... 201

Figure 6.10 The solid product from co-pyrolysis of sawdust with oily sludge at a mass

ratio of 3: 1 ( 500 °C, 20 °C/min) .... ... .... ..... ............. .. ..... ....... ..... .. ............. ... 205

17

GLOSSARY

Letters

2-Pro

ANOVA

API

ASTM

CCC

CCME

CHX

DCM

DTG

EA

EOR

EPA

FIT

GC

GC-FID

2-Propane

Analysis of variance

American Petroleum Institute

American Society for Testing and Materials

Decane

Hexadecane

Tetratriacontane

Pentacontane

Central composite circumscribed

Canadian Council of Ministers of Environment

Cyclohexane

TPH concentrations (mg/g) in the recovered oil layer

TPH concentrations (mg/g) in the original sludge

Dichloromethane

Differential thermal-gravimetric analysis

Ethyl acetate

Enhanced oil recovery

Environmental Protection Agency

Freeze thaw

Gas chromatograph

Gas chromatograph with flame ionization

18

ICP

K

MAE

MEK

MSE

ORi

PAHs

PH Cs

PLE

R

RCRA

RO

Ro

RSM

S/S

SDS

SFE

B

SW

TGA

Inductively coupled plasma

The sum of percentage of the effect of impact factor at each level

Microwave assisted extraction

Methyl ethyl ketone

Mass of recovered oil (g)

Mass of residues (g)

Mass of oily sludge (g)

Mechanical shaking extraction

The oil recovery rate of the ith extraction

Poly aromatic hydrocarbons

Petroleum hydrocarbons

Pressurized liquid extraction

Extreme difference

Resource Conservation and Recovery Act

Recovered oil

Oil recovery rate

Response surface methodology

Solvent-to-sludge

Sodium dodecyl sulphate

Supercritical fluid extraction

The mass percentage of oily sludge in mixture (wt.%)

Sludge-to-water

Thermos-gravimetric analysis

19

TPH

UAE

UAEB

UAEP

W/0

Xi, Xj

y

Yproduct

Yproduct I

Yproduct2

a

/Jo

/Ji

/Ju

/Ju

Total petroleum hydrocarbon

Ultrasonic assisted extraction

Ultrasonic assisted extraction bath

Ultrasonic assisted extraction probe

water-in-oil

The independent variables

The response variable

The theoretical yield of co-pyrolysis product (wt.%)

The yield of sawdust during co-pyrolysis (wt.%)

The yield of oily sludge during co-pyrolysis (wt.%)

The mass percentage of sawdust in mixture (wt.%)

The intercept coefficient of the model

The linear coefficient of the model

The quadratic coefficient of the model

The interaction coefficient of the model

20

ACKNOWLEDGEMENT

This thesis reflects the final part of my Ph.D. study at the University of Northern

British Columbia. First of all, I am grateful to my supervisor, Dr. Jianbing Li, for

support and direction throughout my Ph.D. study. His academic advising,

professional knowledge, and insightful thoughts not only benefit my Ph.D. study,

but also will help me greatly in my future research career.

I would like to thank Dr. Joselito M. Aracena, Dr. Ron Turing, Dr. Jueyi Sui, and Dr.

Liang Chen for being my supervisory committee members. I greatly appreciate their

insightful comments, guidance, and suggestions.

Thanks goes to fellow Ju Zhang, Siddhartho Shekhar Paul, and Gopal Saha for their

kind sharing their knowledge, skills, and thoughts during the entire research

process. Thanks to visiting scholars: Xinying Zhang and Shuhui Huang for their help

with designing experimental apparatus, securing equipment, and conducting

experiments.

I would also like to thank Quanji Wu, Dominic Reiffarth, Bill McGill, Heath de la

Giroday, and Conan Ma for their support with sample analysis at Northern

Analytical Laboratory and sample delivery at ChemStore. Furthermore, I wish to

thank Alida Hall for her helpful instructions as the supervisor of my teaching

assistant position at UNBC.

Special thanks go to my wife, parents, and friends for their encouragement and

support throughout my entire Ph.D. study. I wish to thank Dr. Xinghui Xia from

Beijing Normal University and Dr. Haobo Hou from Wuhan University for

supporting me to study at UNBC.

Appreciation is extended to Prince George Husky Refinery for providing various oily

sludge samples to my research.

Enormous thanks is extended to the Natural Sciences and Engineering Research

Council of Canada, Environment Canada, and UNBC for providing financial support.

21

Chapter 1 General Introduction

1.1 Background

Petroleum industries generate considerable quantities of oily sludge during various

oil production processes including crude oil exploration, conveyance, storage, and

downstream refining. It is estimated that each refinery generates an annual average of

30,000 tons of oily sludge (EPA, 1991). The oily sludge production volume is expected to

increase as a result of the ascending demand on refined petroleum products worldwide (BP,

2012). The quantities and properties of oily sludge depend on the nature of crude oil,

storage conditions, down-stream processing schemes, and the design of refining apparatus.

Generally, oily sludge physically exists as a stable water-in-oil (W/0) emulsion, which

consists of water, solids, various petroleum hydrocarbons (PHCs), and metals. The

stability ofW/0 emulsions depends mainly on a protective film that inhibits water droplets

from coalescing with each other. This interfacial film is composed of many natural

emulsifiers such as some PHCs constituents (e.g., asphaltenes and resins), fine solids, oil

soluble organic acids, and other finely divided materials (Hu et al., 2013). The chemical

composition of oily sludge can vary over a wide range, depending on sources, processing

scheme, and equipment and reagents used in refining process. Typically, oily sludge

contains 15-50% of total petroleum hydrocarbon (TPH, or oil), 30-85% of water, and 5-

46% of solids. Due to the existence of carcinogenic and mutagenic poly aromatic

hydrocarbons (PAHs) and toxic heavy metals (e.g., chromium, cadmium, and lead) in oily

sludge, it has been categorized as a hazardous waste by many environmental regulations

world widely ( da Rocha et al., 201 O; Liu et al., 2009). Any improper disposal of oily sludge

could cause serious environmental contamination and pose threats to the health of

22

surrounding receptors. Due to its hazardous nature and increasing quantities around the

world, there is a pressing need to develop effective, economically feasible, and

environmentally friendly techniques to address the oily sludge problem in petroleum

industries.

1.2 Statement of the problem and research objectives

In the past decades, oily sludge is sent to sludge pit or landfills for natural attenuation

(da Silva et al., 2012). However, this approach is associated with various disadvantages

such as low degradation efficiency, too time-consuming, high risks of environmental

contamination, and occupying large amount of valuable land resources (da Silva et al.,

2012; Hu et al., 2013). It is also not suitable in cold regions such as the vast area of Canada

because the activity of PHCs degradation microbes could be compromised in cold

environment (Yang et al., 2009). Moreover, the becoming more restrictive environmental

regulations (e.g., Ontario Environmental Protection Act, Regulation 347 and Resource

Conservation and Recovery Act) have been enacted, by which the direct land treatments

for hazardous wastes disposal are limited.

Considering oily sludge contains relatively high amount of oil, oil recovery could be

the most desirable approach for its treatment (Elektorowicz and Habibi, 2005; da Silva et

al., 2012). The oil recovery technology should be capable of recovering oil in a form that

can be sent to a refinery for processing to produce high value petrochemical products,

while the residues of oily sludge can be easily cleaned up. This approach not only can reuse

the valuable energy content, but also can significantly reduce the volume and toxic level

of waste, alleviating its negative impact on the environment. In recent years, a number of

physical, chemical, and biological processing methods have been proposed for the oil

23

recovery from oily sludge (Hu et al., 2013). These novel methods including ultrasonic

irradiation, microwave irradiation, pyrolysis, electrokinetic processing, biosurfactant

demusification, freeze/thaw demusification, and solvent extraction (Hu et al., 2013). These

methods are associated with various advantages and limitations (da Silva et al., 2012; Hu

et al., 2013). For example, microwave and pyrolysis treatment can greatly reduce the

volume of oily sludge, but the energy consumption of these approaches is high; solvent

extraction can recover most useable PH Cs from oily sludge but this method requires large

amount of organic solvent to achieve promise oil recovery rate (Hu et al., 2013). Moreover,

none of these methods is universally applicable because the properties of oily sludges from

different sources and/or petrochemical production schemes vary significantly. Single

treatment might not be effective in treating complex oily sludge, and thus the combined

treatment methods are needed to address the limitations of each single method.

Among various oil recovery methods, ultrasonic irradiation, solvent extraction,

freeze/thaw, and pyrolysis represent promising techniques due to their inherent merits such

as short treatment duration compared to biodegradation, environmentally friendly, and

promising oil recovery performance. Ultrasonic irradiation has been used for the removal

of adsorbed materials from solid particles and the demulsification of stable water/oil (W/0)

emulsions (Xu et al., 2009; Zhang et al., 2012). The microbubbles cavitation spawned

during ultrasonic irradiation does not only generate strong shear force in the bulk system

but also generate large amount of heat in a few microseconds, and could thus improve the

separation of oil from oily sludge (Li et al., 2013). It also has been reported effective in

reducing the salt and water amount in crude oils (Ye et al., 2008; Gholam and Dariush,

2013), however, its desalting effect on the recovered oil from oily sludge is yet to known.

Solvent extraction is a simple and effective process that can separate PHCs from various

24

matrix, has been proven successful for oily wastes treatment (Zubaidy and Abouelnasr

2010; Taiwo and Otolorin 2009). In this process, oily wastes and solvent are mixed in an

appropriate proportion to ensure adequate miscibility of oil in solvent, while most water

and solids are rejected as unwanted impurities which can be removed by gravitational

settling or centrifugation. The oil and solvent mixture can then be separated by distillation

for the purpose of both oil and solvent recycling (Al-Zahrani and Putra, 2013). Moreover,

freeze/thaw treatment has been proven as a cost-effective dewatering process for the break

ofW/0 emulsion. The volume expansion of water droplets when turning to ice could cause

the coalescence of emulsified water droplets and the change of interfacial tension between

water and oil phases, and these were the main driving forces of dewatering (Lin et al.,

2007). Pyrolysis is an effective thermo-chemical conversion process during which oily

wastes is heated in a closed oxygen-free reactor system at moderate operating temperatures

(i.e., usually 200 to 500 °C) (Isahak et al., 2012). This process can convert organic wastes

into combustible gases, pyrolysis oil, and char. Combustible gas and pyrolysis oil can be

used for energy supply and char after proper modification can be used as an adsorbent for

pollution control (Bernardo et al., 2012).

In order to effectively handle complex oily sludge, it is important to improve these

oil recovery techniques. The combinational utilization of these techniques could be an

alternative solution for the improvement of single method. Ultrasonic assisted extraction

(UAE) uses the turbulence and heat generated by ultrasonic irradiation to facilitate the

mixing of solvents and PHCs in various matrixes, which could significantly reduce the

extraction time and increase the extraction efficiency (Bossio et al., 2008). In treating high

moisture oily sludge, solvent extraction treatment alone is not sufficient to remove the

highly emulsified water in the extraction matrix. The combined usage of solvent extraction

25

and freeze/thaw could be a solution to remove the undesirable emulsified water. Moreover,

the co-pyrolysis of organic wastes (i.e., waste tyres, used lubricating oils, and municipal

sewage sludge) with various biomass has been reported effective in improving the quantity

and/or the quality of recovered oil without any change in the system process (Kar, 2011;

Onal et al., 2014). Therefore, investigating the combined effect of different techniques on

oil recovery from oily sludge is of great importance and could provide more advanced

solutions for the oily wastes treatment in petrochemical industries.

The main objective of this thesis is to develop novel combined physical-chemical

techniques based on ultrasonic irradiation, freeze/thaw, solvent extraction, and pyrolysis

treatments for the oil recovery from various refinery oily sludge. The specific research

objectives include:

(1) Investigating the oil recovery and desalting effect of ultrasonic irradiation on

refinery tank bottom sludge. The effect of influential factors including ultrasonic power,

treatment duration, sludge-to-water (S/W) ratio, and sludge-water slurry temperature were

studied to examine their impacts on the treatment performance. In addition, the

concentration of PH Cs and salt in wastewater generated as a by-product from this treatment

were quantified.

(2) Examining the effect of ultrasonic assisted extraction on oil recovery from tank

bottom sludge. The oil recovery performance of two types of UAE system including

ultrasonic assisted extraction probe (UAEP) system and ultrasonic assisted extraction bath

(UAEB) system were studied compared to that of mechanical shaking extraction (MSE)

treatment. A number of factors including solvent type, solvent-to-sludge (S/S) ratio,

ultrasonic irradiation duration, ultrasonic power, ultrasonic bath temperature, and the

26

number of extraction cycles, were investigated for their individual effect on the treatment

using orthogonal experimental design.

(3) Developing a combinational solution of solvent extraction and freeze/thaw for the

oil recovery from high moisture dredged sludge from refinery wastewater pond. The oil

recovery rate, solvent recovery rate, and the waste reduction rate of sludge were examined.

The performance of freeze thaw treatment on improving the quality of recovered oil in

terms of its TPH content. Three groups of experiments were conducted, including

freeze/thaw treatment alone, solvent extraction alone, and combined solvent extraction

with freeze/thaw. The TPH content and PHCs fraction distribution in the recovered oil

were reported, and the properties of solid residue as the by-product of the treatment process

were analyzed.

( 4) Evaluating the co-pyrolysis process of oily sludge with biomass to improve the

generation ofbio-oil and bio-char. The synergistic effect of the oily sludge addition on the

bio-oil derived from pyrolysis of sawdust was investigated. Co-pyrolysis of sawdust with

oily sludge was carried out in a fixed-bed reactor under different pyrolysis conditions. The

effect and interaction of different influential factors including sawdust addition amount,

temperature, and heating rate on the yield of bio-oil was investigated by the response

surface methodology (RSM). The characteristics of products from the co-pyrolysis of

sawdust with oily sludge were determined to evaluate their possibility of being a potential

energy source and petrochemical feedstock.

It should be noted that two different oil recovery calculation methods were used in

this thesis. In Chapter 3, the oil recovery and desalting effect of ultrasonic irradiation on

oily sludge was investigated. This research project was a successive study to the previous

research project "Oil recovery from refinery oily sludge via ultrasound and freeze/thaw",

27

so the oil recovery was calculated using the same equation reported by Zhang et al. (2012)

as the ratio of TPH mass in the recovered oil to the TPH mass in the original oily sludge;

In Chapter 4-6, the oil recovery/yield was calculated as the mass percentage ratio of

recovered oil to that of original oily sludge, which is a widely reported oil recovery

calculation method in literatures.

28

Chapter 2 Literature Review *

* This review has been published as: Hu, G.J., Li, J.B., and Zeng, G.M., 2013. Recent

development in the treatment of oily sludge from petroleum industry: A review. Journal

o_f Hazardous Materials, 261 , 470-490.

2.1. Introduction

Oily sludge, generated from petroleum production processes in petrochemical

industry at a large quantity, is one of the major wastes that has received increasing attention

in recent years. It contains a high concentration of petroleum hydrocarbons (PH Cs) and

other recalcitrant components. As being recognized as a hazardous waste in many countries,

the improper disposal or insufficient treatment of oily sludge can pose serious threats to

the environment and human health (da Silva et al., 2012; Xu et al., 2009). The effective

management of oily sludge has become a worldwide problem due to its hazardous nature

and increasing production quantity around the world. During the past years, a variety of

oily sludge treatment methods have been developed, such as landfarming, incineration,

solidification/stabilization, solvent extraction, ultrasonic treatment, pyrolysis,

photocatalysis, chemical treatment, and biodegradation (Xu et al., 2009; Mrayyan and

Battikhi, 2005; Liu et al. , 2009; Mater et al. , 2006; da Rocha et al. , 2010; Roldan-Carrillo

et al. , 2010; Zubaidy and Abouelnasr, 2010; Yan et al., 2012) . By employing these

technologies, the contents of hazardous constituents can be reduced or eliminated, and its

deleterious environmental and health impacts can thus be mitigated. However, due to the

recalcitrant nature of oily sludge, few technologies can reach a compromised balance

between satisfying strict environmental regulations and reducing treatment costs. As a

result, there is a need for a comprehensive discussion of current oily sludge treatment

29

methods to identify their advantages and limitations. The main objectives of this chapter

are (a) to introduce the source, characteristics, and environmental impact of oily sludge in

the petroleum industry, (b) to summarize current treatment methods available for dealing

with oily sludge, (c) to discuss the advantages and limitations of these methods, and (d) to

discuss future development needs to meet resource recycling and waste disposal standards.

2.2 Source, characteristics and toxicity of oily sludge

2.2.1 Oily sludge source

Both the upstream and downstream operations in petroleum industry can generate a

large amount of oily wastes. The upstream operation includes the processes of extracting,

transporting and storing crude oil, while the downstream operation refers to crude oil

refining processes. The oily waste generated in petroleum industry can be categorized as

either simple oil or sludge depending on the ratio of water and solids within the oily matrix

(Al-Futaisi et al. , 2007). Simple waste oil generally contains less water than sludge that is

highly viscous and contains a high percentage of solids. Stable water-in-oil (W/0)

emulsion is a typical physical form of petroleum sludge waste (Elektorowicz and Habibi,

2005). In the hazardous wastes list by US EPA, oily sludges have been coded as wastes

F037-F038 and K048-K052 depending on the sources they produced (EPA, 2008). In the

upstream operation, the related oily sludge sources include slop oil at oil wells, crude oil

tank bottom sediments, and drilling mud residues (Dara and Sarah, 2003). A variety of oily

sludge sources exist in downstream operation, including (a) slop oil emulsion solids

(K049); (b) heat exchange bundle cleaning sludge (KOSO); ( c) residues (K05 l) from

oil/water separator, such as the American Petroleum Institute (API) separator, parallel plate

30

interceptor, and corrugated plate interceptor (CPI); (d) sediments (K052, K169, and Kl 70)

at the bottom ofrail, truck, or storage tanks; (e) sludge (K048) from flocculation-flotation

(FFU), dissolved air flotation (DAF), or induced air flotation (IAF) units, and (f) sludges

(F037 and F038) from the primary and secondary separation of solids/water/oil during the

storage or treatment of process refining wastewaters (van Oudenhoven et al., 1995). In

particular, the bottom sediments in crude oil storage tanks represent the most intensively

studied oily sludge in literatures. Prior to being refined to petroleum products, crude oil is

temporarily housed in storage tanks where it has a propensity to separate into heavier and

lighter petroleum hydrocarbons (PHCs). The heavier PHCs often settle along with solid

particles and water (Ayotamuno et al. , 2007). This mixture of oil, solids, and water

deposited at the storage tank bottom is known as oily sludge (Greg et al., 2004). It is

removed during tank cleaning operations and sent for further treatment or disposal.

The sludge quantity generated from petroleum refining processes depends on several

factors such as crude oil properties ( e.g., density and viscosity), refinery processing scheme,

oil storage method, and most importantly, the refining capacity. According to an

investigation conducted by US EPA, each refinery in the United States produces an annual

average of30,000 tons of oily sludge (EPA, 1991). In China, the annual production of oily

sludge from petrochemical industry is estimated to be 3 million tons (Wang et al. , 2012).

Generally, a higher refining capacity is associated with a larger amount of oily sludge

production. It has been estimated that one ton of oily sludge waste is generated for every

500 tons of crude oil processed (van Oudenhoven et al. , 1995). Figure 2.1 shows the global

refining throughputs in recent years, and it is estimated that that more than 60 million tons

of oily sludge can be produced every year and more than 1 billion tons of oily sludge has

been accumulated worldwide (da Silva et al., 2012; BP, 2012). It is also expected that the

31

total oily sludge production amount is still increasing as a result of the ascending demand

on refined petroleum products worldwide (Bhattacharyya and Shekdar, 2003; BP, 2012).

25000

-;:: ISi 2009 ra

"C ~2010 ....... "' 20000 a:i ... ... ra

..c 0 0 0 ~ 15000 "' ... ::, C. .c t>.O ::, 0 ... 10000 .c ... >-... Q) C

I+: Q) a: 5000

0

Countries & regions

Figure 2.1 Worldwide daily refining throughputs in recent years

2.2.2 Characteristics of oily sludge

In general, oily sludge is a recalcitrant residue characterized as a stable W /0 emulsion

of water, solids, PHCs, and metals (Mazlova and Meshcheryakov, 1999). An emulsion is

the mixture of a liquid dispersed in another immiscible liquid as fine colloidal droplets.

The stability of W /0 emulsions depends on many factors such as the nature of oil, the

water-to-oil ratio in the mixture, and emulsifiers. Emulsifiers, such as some PHCs

constituents (e.g., asphaltenes and resins), fine solids, oil soluble organic acids, and other

32

finely divided materials, can form a protective film around the surface of water droplets

that inhibits them from coalescing with each other (Yang et al., 2009; Kralova et al., 2011 ).

The pH value of oily sludge is usually in a range between 6.5 and 7.5 and its chemical

composition varies over a wide range, depending on crude oil source, processing scheme,

and equipment and reagents used in refining process (da Silva et al., 2012). For example,

the total petroleum hydrocarbon (TPH) contents in oily sludge can range from 5% to 86.2%

by mass, but more frequently in the range of 15-50% (Tahhan et al., 2011; Biswal et al.,

2009; Mohan and Chandrasekhar, 2011; Liu et al., 2010; Liu et al., 2012), whereas the

contents of water and solids are in the range of 30-85% and 5-46%, respectively

(Ramaswamy et al., 2007). The PHCs and other organic compounds in oily sludge can be

generally classified into four fractions, including aliphatics, aromatics, nitrogen sulphur

oxygen (NSO) containing compounds, and asphaltenes (Mrayyan and Battikhi, 2005;

Reddy et al., 2011). The aliphatics and aromatic hydrocarbons usually account for up to

75% of PHCs in oily sludge (Ward et al., 2003), and their most common compounds

include alkanes, cycloalkanes, benzene, toluene, xylenes, naphthalene, phenols and

various polycyclic aromatic hydrocarbons (PAHs) (e.g., methylated derivatives of fluorine,

phenanthrene, anthracene, chrysene, benzofluorene, and pyrene) (Kriipsalu et al., 2008).

The NSO fraction contains polar compounds such as naphthenic acids, mercaptans,

thiophenes and pyridines (Kriipsalu et al., 2008). The nitrogen (N) content accounts for

less than 3% in oily sludge, and most of them are contained in the distillate residue as part

of asphalt and resin fraction. The sulphur (S) content can be in the range of 0.3-10%

whereas the oxygen (0) content is usually less than 4.8% (Kriipsalu et al., 2008).

Asphaltenes are mixtures of pentane-insoluble and colloidal compounds including poly

aromatic and alicyclic molecules with alkyl substitutes (usually methyl groups), and they

33

vary in molecular weight between 500 and several thousand (Tavassoli et al. , 2012).

Asphaltenes and resins can be responsible for the stability of oily sludge emulsion since

these constituents contain hydrophilic functional groups and consequently can act as

lipophilic emulsifiers (Rondon et al., 2006). Usually, oily sludge is composed of 40-52%

alkanes, 28-31 % aromatics, 8-10% asphaltenes, and 7-22.4% resins by mass (Mishra et al.,

1999; van Hamme et al., 2000).

As a result of diverse chemical compositions in oily sludge, its physical properties

such as density, viscosity, and heat value can vary significantly. The property

measurements obtained from one oily sludge source cannot be applied to another source

or to another sludge sample of the same source but collected on a different day or different

location (API, 2010). However, a key factor affecting the physical properties of oily sludge

is the polarity and molecular weight of chemical constituents in sludge, and it is possible

to make an empirical estimation of physical properties based on the chemical compositions

of sludge (API, 1992; API, 2010). In addition to organic chemical components, oily sludge

also contains a variety of heavy metals resulted from different sources. The species and

concentrations of these heavy metals could vary over a wide range as similar to organic

compounds. According to a report from American Petroleum Institute (API) (API, 1989),

metal concentrations in oily sludge obtained from petroleum refineries are generally 7-80

mg/kg for zinc (Zn), 0.001-0.12 mg/kg for lead (Pb), 32-120 mg/kg for copper (Cu), 17-

25 mg/kg for nickel (Ni), and 27-80 mg/kg for chromium (Cr). It is possible that a very

high concentration of heavy metals could be found in oily sludge. For example, the metal

concentration in oily sludge from refineries was reported in recent literatures as 1299

mg/kg for Zn, 60200 mg/kg for iron (Fe), 500 mg/kg for Cu, 480 mg/kg for Cr, 175 mg/kg

34

for Ni, and 565 mg/kg for Pb, respectively (Reddy et al., 2011; Marin et al., 2006; Smadar

et al., 2001; Otidene et al., 2010).

2.2.3 Toxicity and impact of oily sludge

Due to the existence of high-concentration toxic substances, the improper disposal of

oily sludge can pose serious threats to the receiving environment. After entering the

terrestrial environment, oily sludge can disturb the physical and chemical properties of

receiving soils, leading to soil morphological change (Robertson et al., 2007). The oily

sludge contaminated soils may create nutrient deficiency, inhibit seed germination, and

cause restricted growth or demises of plants on contact (Al-Mutairi et al. , 2008). Due to its

high viscosity, oily sludge components can be fixed in soil pores, adsorbed onto the surface

of soil mineral constituents, or form a continuous cover on soil surface (Trofimov and

Rozanova, 2003). These would lead to reduced hygroscopic moisture, hydraulic

conductivity, and water retention capacity (i.e. wettability) of soils (Suleimanov et al. ,

2005; Trofimov and Rozanova, 2003). In particular, the components with higher molecular

weight in sludge and their degradation products could remain near soil surface and form

hydrophobic crusts that decrease water availability and limit water/air exchange (Tang et

al., 2012). A long-term (i.e. several years) hydrophobicity of oily wastes contaminated

agricultural soils has been reported in western Canada although many PHCs-contaminated

soils eventually take up water (Robertson et al. , 2007).

The disposal of oily sludge to the environment could lead to various toxic effects

caused by PH Cs and heavy metals. Most of the heavy metals have a cumulative effect and

are of particular hazard. In terms of PH Cs, the polycyclic aromatic hydrocarbons (PAHs)

are of major concerns since they are genotoxic to humans and other ecological receptors

35

(Robertson et al., 2007). The PHCs in oily sludge could migrate down through the soil

profile and enter groundwater that is linked with other aquatic systems, causing serious

adverse consequences such as reduced diversity and abundance of fish in the aquatic

system (Trofimov and Rozanova, 2003). PHCs in oily sludge could also decrease the

activity of soil enzymes (i.e. hydrogenase and invertase) and pose toxic effects on the soil

microorganisms (Suleimanov et al., 2005). Moreover, after remaining in the terrestrial

environment for an extended period of time, the weathered ( or aged) chemical residues

may appear to resist de-sorption and degradation, and they have considerable time to

interact with soil components (Tang et al. , 2012). Covalent bonding between organic

compounds in sludge residues and humic polymers (e.g., humin, fulvic acid and humic

acid) in soil could form stable dialkylphthalates, long-chain alkanes and fatty acids that are

resistant to microbial degradation (Certini, 2005; Alexander, 2000). Due to the hazardous

nature of oily sludge, many regulations in the world such as the Resource Conservation

and Recovery Act (RCRA) in USA have established strict standards for its handling,

storage and disposal (EPA, 1980). For example, it was regulated that all surface

impoundments which treat or store hazardous wastes must either be double lined or taken

out of service (API, 2010). Even if oily sludge is disposed of in lagoon which is lined with

cement and bricks, problems of odour and fire hazard would still be created (Bhattacharyya

and Shekdar, 2003). Refinery oily sludge deposited in lagoons or landfills has also been

identified as a stationary source of atmospheric volatile organic compounds (VOCs)

pollution (Cheremisinoff and Rosenfeld, 2009). Such air pollutant emissions can create

health risks to facility workers and surrounding communities (Santiago et al., 2002).

36

2.3 Overview of sludge treatment methods

Generally, a three-tiered oily sludge waste management strategy should be applied

(Al-Futaisi et al. , 2007). This includes (1) employing technologies to reduce the quantity

of oily sludge production from petroleum industry, (2) recovering and reclaiming valuable

fuel from existing oily sludge, and (3) disposing of the unrecoverable residues or oily

sludge itself if neither of the first two tiers is not applicable (Pinheiro and Rolanda, 2009).

The first tier is to prevent the generation of oily sludge and reduce its volume of generation,

while the next two tiers are more concerned about the effective treatment of existing oily

sludge which is the focus of this chapter. A variety of methods have been developed for

the treatment of oily sludge as discussed below (Figure 2.2).

Oily sludge

Fuel Oil recovery Residuals

Disposal

; t!:! 5l fr.

.-,J - S' 0 n ~ .~ s ct ,::' t:::f :a- E; ~ ~ B l!!j ; .. "' S" ;· I ;, "' ·1 • - 0 IC 0 =-::;' It·

~ S' • !'ir Q. ct lo( .... & ~ ;;· Q rr! 1:1; = e!.. ...

= I!. 0 19 19 r:; = 11 =-= a '< ... ::t. ~ .. l'I I;' ~ !'ir .. .. 0 = Q ::t. =-... .. !':I l ;: = .. ::t. t:!l!j Iii) -· Iii) ::. 5 0 .. a .. a ::t. 0 = ::t. • 0 $1 =- •• = • = -· & Iii) • " • ,= .. ::. = ! = = ::t. - • 0 • = § = =-

Figure 2.2 Overview of oily sludge treatment methods

37

2.4 Oil recovery methods for oily sludge treatment

Recycling is the most desirable environmental option for handling oily sludge since

it enables petroleum industry to reuse valuable oil for reprocessing and reformulating or

energy recovery. Moreover, recycling of oily sludge can reduce the disposal volume of

hazardous waste outside the industrial zone, prevent the extent of contamination, and

decrease the use of non-renewable energy resources. According to API (1989), the primary

environmental consideration in handling oily sludge should be the maximization of

hydrocarbon recovery. It was reported that in the USA, more than 80% of PHCs wastes

generated within a refinery is recycled, with the remaining 20% managed by an acceptable

disposal method (API, 2010). In general, oily sludge with a high concentration(> 50%) of

oil and a relatively low concentration of solids ( < 30%) are preferable to be recycled (Hahn,

1994). Other studies suggested that oily sludge even containing a relatively low oil content

(> 10%) still merits a treatment of oil recovery (Ramswamy et al., 2007). A number of

methods have been available for recycling hydrocarbons from oily sludge.

2.4.1 Solvent extraction

Solvent extraction has been widely used for removing semi-volatile and non-volatile

organic compounds from soiVwater matrices. It mixes oily wastes with solvent at desired

proportions to ensure complete miscibility, while the water, solid particles and

carbonaceous impurities are rejected by extraction solvent. The solvent/oil mixture is then

sent for distillation to separate oil from solvent (Al-Zahrani and Putra, 2013). Various

solvents have been reported for oily sludge treatment. Gazineu et al. (2005) used turpentine

as a solvent for oil extraction, and they found that the extracted oil accounted for 13-53%

of the original sludge mass. Zubaidy and Abouelnasr (2010) compared the effects of

38

several organic solvents such as methyl ethyl ketone (MEK) and liquefied petroleum gas

condensate (LPGC), and they found that at a solvent-to-sludge ratio of 4: 1, the highest oil

recovery rate of 39% and 32% was obtained by MEK and LPGC extraction, respectively.

Their results indicated that the ash, carbon residue, and asphaltene levels in the recovered

oil were mostly improved when using MEK as the solvent, but the recovered oil still

contained high levels of sulfur and carbon residue, thus the recovered oil would require

further purification prior to be used as a fuel. El Naggar et al. (2010) used several solvents

such as naphtha cut, kerosene cut, n-heptane, toluene, methylene dichloride, ethylene

dichloride and diethyl ether to recover oil from dry and semi dry petroleum sludge, while

toluene gave the highest PHCs recovery rate of 75.94%. Meyer et al. (2006) found that

petroleum solvent oil with a high percentage of ring compounds (e.g., naphthenics and

aromatics) such as catalytic cracking heating oil was highly effective in dissolving

asphaltenic components in oily sludge, and the solvent oil with paraffinic character like

virgin paraffinic diesel was effective for sludge with more paraffinic (waxy) components.

Hexane and xylene have also been used as solvents to recover hydrocarbons from

petroleum sludge, and it was reported that 67.5% of PH Cs in sludge could be recovered,

with most of them in the range of C9 to C2s (Taiwo and Otolorin, 2009).

Figure 2.3 presents a simplified conceptual diagram for a field-scale solvent

extraction process. Oily sludge waste is firstly mixed in the reactor column with a solvent

which selectively dissolves the oil fraction of sludge and leaves the less soluble impurities

at the column bottom. The oil-solvent solution is then transferred to a solvent distillation

system where the solvent is separated from oil. The separated oil is considered as oil

recovery, while the separated solvent vapour can be liquefied through a compressor and

cooling system and sent to a solvent recycling tank. The solvent can be used for repeating

39

the extraction cycle. The bottom impurities from reactor column are pumped to a second

distillation system, and the solvent contained in the impurities is separated and then sent

to the solvent recycling tank, while the waste residues after separation may need further

treatment. In general, the performance of solvent extraction is affected by a number of

factors such as temperature and pressure, solvent-to-sludge ratio, mixing, and the property

of solvent. Mixing and heating are usually required to improve the dissolution of sludge

organic components in solvent (Meyer, 2006). High temperature can accelerate the

extraction process but it can result in the loss of volatile PHCs and solvent through

evaporation, while low temperature would decrease the cost of extraction process but it

can lead to lower oil recovery efficiency (Fisher, et al., 1997). Lower pressure is favoured

during distillation since solvent evaporation could occur at a relatively lower distillation

temperature. A lower distillation temperature can not only save heating cost, but also

prevent thermos-degradation of solvents. Moreover, the quantity and quality of recovered

oil can be improved with increasing solvent-to-sludge ratio. For example, it was observed

that the amount of ash and high-molecular-weight hydrocarbons in the recovered oil

decreased with the increasing amount of solvents (Zubaidy and Abouelnasr, 2010).

Generally, solvent extraction represents a simple but efficient method to separate oily

sludge into valuable hydrocarbon and a solid or semi-solid residue with reduced volume.

This approach has the potential to treat a large volume of oily sludge depending on the

extraction column design. In order to prevent the emission of solvent vapour, a closed and

continuous process capable of retaining the evaporated solvent is usually desired. Heating

is also required for solvent recycling, and this could increase the energy cost of application.

One major obstacle of applying solvent extraction to field-scale oily sludge treatment is

that a large volume of organic solvents are required. This could result in significant

40

economic and environmental concerns. Some alternative methods have been developed to

improve the performance of solvent extraction. For example, supercritical fluid extraction

(SFE) can extract PH Cs in soil matrices more rapidly than conventional solvent extraction,

and more importantly, it can eliminate the use of organic solvents (Avila-Chavez et al. ,

2007) . However, when this method is used for extracting oil from a large volume of oily

sludge, it may be subject to low efficiency and high variability (Schwab et al., 1999).

1

Oily sludge Oil and solvent mixture

Recycled solvent

2 Solvent vapor

4

3

----Bottom impurities

2

Solvent vapor

.... _ .. wast.e residues

Figure 2.3 Flow diagram of solvent extraction process (1: reactor column; 2: distillation

system; 3: solvent recycling tank; 4: compressor and cooling system)

2.4.2 Centrifugation treatment

Centrifugation has been widely applied to field-scale oily sludge treatment although

few scholarly literatures have been reported in recent years. It utilizes a special high-speed

rotation equipment to generate a strong centrifugal force which can separate components

with different densities ( such as water, solids, oil, and pasty mixtures in oily sludge) within

41

a short time. In order to enhance the centrifugation performance and reduce energy

consumption, the viscosity of oily sludge needs to be reduced through sludge pre-treatment,

such as the addition of organic solvents, demulsifying agents and tensioactive chemicals,

the injection of steam, and direct heating (Zubaidy and Abouelnasr, 201 O; Conaway, 1999;

Cambiella et al. , 2006; Nahmad, 2012). For example, Conaway (1999) reported that after

viscosity reduction using heat pre-treatment, the less viscous petroleum sludge could be

effectively treated by a disc/bowl centrifuge, with more than 80% of the waste volume

being obtained as liquid effluent from the first centrifugation, and the residue from

centrifugation was then mixed with hot water and centrifuged again. The effluent from two

centrifugations was combined and sent to refinery for processing (Conaway). Cambiella et

al. (2006) found that a small amount of a coagulant salt (CaCh, in the concentration range

of 0.01-0.5 M) could improve the water-oil separation process by centrifugation, with a

high oil separation efficiency of 92-96% being obtained. One recent US patent reported an

approach of recovering crude oil from oily sludge through centrifugation, with oily sludge

being mixed with demulsifying reagents at predefined ratio and then agitated to allow

homogenization, while the processed mixture is centrifuged to separate PHCs, water and

solids (Nahmad, 2012). Huang et al. (2014) developed a model to describe the effect of the

size distribution of solid particle and viscosity of oily sludge on the W /0 recovery

performance by centrifugation treatment. According to their results, decreasing the

viscosity by adding solvents or by preheating could lead to a better solid removal rates in

oily sludge treatment (Huang et al. , 2014).

Figure 2.4 presents a process of using centrifugation for oily sludge treatment. Oily

sludge is firstly mixed with demulsifying agent or other chemical conditioners. The

mixture is then treated by hot steam in a pre-treatment tank in order to reduce its viscosity.

42

The less viscous sludge is mixed with water at a certain sludge/water ratio for high-speed

centrifugation. After centrifugation, the separated water containing high concentration of

PHCs is drained for further wastewater treatment, and the separated oil (still containing

water and solids) is sent to a gravimetric separator for further separation to obtain the

recovered oil. The separated water from the separator is sent to wastewater treatment. The

sediments from centrifugation and separator are collected as solid residues for further

treatment. In general, centrifugation is a relatively clean and mature technology for oily

sludge treatment, and its oil separation from sludge is effective. Another advantage is that

centrifugation equipment usually does not occupy much space (da Silva et al., 2012).

However, this process requires high energy consumption to generate strong enough

centrifugation force to separate oil from petroleum sludge. The use of centrifugation has

been limited to small scales because of the high equipment investment and limitations (Nii,

2009). In addition, centrifugation can bring noise problems (da Silva et al. , 2012).

Moreover, the introduction of demulsifying agents and tensioactive chemicals for sludge

pre-treatment not only increase the processing cost, but also bring environmental concerns.

43

Oily sludge .... r -

Pretreatment tank -r

mulsifying agent or De oth er chemical conditioners

,J' .... Condensate

Steam

Water

Treated water discharge ,J' ....

JI\

Wastewater treatment unit

Recovered oil

'\

Oil phase to - Gravimetric be purified

, separator

'\

I Waste solids

'I\

r-, r, -, Centrifuge .... ,

L,..J

,J' w Wastewater ....

Figure 2.4 Schematic view of centrifugation used in oily sludge treatment

2.4.3 Surfactant enhanced oil recovery (EOR)

The application of surfactant for removing organic pollutants from solid matrices is a

cost effective and relatively fast process, and it has the potential to treat a large volume of

contaminants. Surfactant is usually an amphiphilic compound, and its molecule consists of

a hydrophobic tail and a hydrophilic tail. The hydrophilic tail makes surfactant molecule

dissolve in the water phase and increase solubility of PH Cs, while the hydrophobic tail

makes it tend to gather at the interfaces to decrease the surface or interfacial tension and

thus enhance the mobility of PHCs (Mulligan, 2009). It was reported that chemical

surfactants, such as sodium dodecyl sulphate (SDS), Corexit 9527®, Triton X-100®, Tween

80® and Afonic 1412-7®, can be used to increase the concentration of PH Cs in aqueous

44

I

phase (Christofi and Ivshina, 2002; Grasso et al., 2001; Cuypers et al., 2002; Prak and

Pritchard, 2002). Abdel-Azim et al. (2011) used three sets of surfactant (nonyl phenol

ethoxylates) based demulsifier mixture to break down petroleum sludge emulsion, and they

found that more than 80% of water can be separated from oily sludge. Similar results can

be found in Dantas et al. (200l)'s research. However, using chemical surfactants can be

associated with a range of problems such as environmental toxicity and resistance to

biodegradation (Christofi and Ivshina, 2002; Mulligan et al., 2001; Whang et al., 2008;

Cort et al., 2002). As compared to chemical surfactant, bio-surfactant has received

increasing attention since it exhibits greater environmental compatibility, more diversity,

better surface activity, lower toxicity, higher emulsification ability, higher selectivity, and

higher biodegradability (Syal and Ramamurthy, 2003; Urum et al., 2004; Ron and

Rosenberg, 2002; Mulligan, 2005; Pekdemir et al., 2005; Chin et al., 2009) . For instance,

Lima et al. (2011) compared the toxicity of five bio-surfactants and one chemical

surfactant called SDS to petroleum degrading microorganisms, and their results showed

that bio-surfactants had a significantly lower toxicity than SDS. Edwards et al. (2003) also

found that the toxicity of bio-surfactants to estuarine invertebrate species were much lower

than that of chemical surfactants.

Bio-surfactants are produced by yeast or bacterial from various substrates including

sugars, oils, alkanes and wastes. They can be grouped into five categories, including (1)

glycolipids, (2) lipopeptides, (3) phospholipids, fatty acids, and neutral lipids, polymetric

bio-surfactant, and particulate bio-surfactant (Mulligan, 2009). Most of the bio-surfactants

are either anionic or neutral, and only a few are cationic such as those containing amine

groups. Their structures include amphiphilic molecules with a hydrophobic moiety (i.e.

fatty acid) and a hydrophilic moiety ( e.g., carbohydrate, carboxylic acid, phosphate, amino

45

acid, cyclic peptide or alcohol) (Mulligan, 2005). A very wide spectrum of microbial

species can be used to produce bio-surfactants, and three groups of bio-surfactants have

been extensively studied, including (1) rhamnolipids (i.e. a type of glycolipids) produced

from Pseudomonas aeruginosa, (2) sorphorolipids (i.e. a type of glycolipids) produced

from Candida bombicola, and (3) surfactins (i.e. a type of Lipopeptides) produced from

Bacillus subtilis (Abalos et al., 2001; Noordman et al., 2002; Rahman et al., 2002; Straube

et al., 2003; Olivera et al., 2000; Makkar and Rockne, 2003). Various laboratory- and field-

scale studies have been conducted to use bio-surfactants in oily sludge treatment. For

example, Lima et al. (2011) isolated five bacterial strains for bio-surfactant production,

and they found that bio-surfactants produced by Dietzia maris sp., Pseudomonas

aeruginosa, and Bacillus sp. respectively recovered up to 95, 93 and 88% of the total oily

sludge as oil, but only 2% of the oil present in oily sludge was recovered when not using

bio-surfactant. Yan et al. (2012) proved the promising oil recovery performance of bio-

surfactant produced by Pseudomonas aeruginosa, and an oil recovery rate of up to 91.5%

was obtained when using bio-surfactant and centrifugation for oil separation from refinery

oily sludge. Long et al. (2013) also applied rhamnolipid to a pilot-scale (100 L) waste crude

oil treatment, and their results indicated that rharnnolipid could recover over 98% of crude

oil from the wastes, while the recovered oil contained less than 0.3% of water. In general,

surfactant enhanced oil recovery method is a simple but a relatively fast and effective

process, and it has a potential to treat a large volume of oily sludge. In spite of the

successful application of surfactants, several factors should be taken into account when

selecting surfactants for oil recovery, including effectiveness, cost, public and regulatory

acceptance, biodegradability, degradation products, toxicity, and ability to recycle. In

particular, the costs of producing bio-surfactants may limit their commercial applications.

46

The related costs can be reduced by improving yields, recovery, and using inexpensive or

waste substrates (Calvo et al., 2009).

2.4.4 Freeze/thaw treatment

One important process of oil recovery from oily sludge is to remove water from a

W/0 emulsion by separating oil and water into two phases, and this process is called

demulsification. Freeze/thaw treatment used for sludge dewatering in cold regions has

been reported as an effective demulsification method (Chen and He, 2003; Jean and Lee,

1999; Lin et al., 2007). As shown in Figures 2.5, two different mechanisms are responsible

for the demulsification. The first one occurs when water phase in emulsion becomes frozen

ahead of oil phase. The volume expansion of frozen water droplets leads to their

coalescence and cause the inner disarrangement of emulsion, and the oil phase is gradually

frozen with temperature dropping (Figure 2.5b ). During the thawing process, the oil phase

coalesces as a result of interfacial tension, and the oil-water mixture can thus be

delaminated into two bulk phases driven by gravitational force (Figure 2.5f) (Lin et al.,

2008). The second mechanism occurs when oil phase becomes frozen ahead of water phase.

This would form a solid cage that encapsulates water droplets during the freezing process

(Figure 2.5c). These water droplets are gradually frozen with temperature dropping, and

the volume expansion of frozen droplets breaks the oil cage. This could produce fine

crevices that allow the unfrozen water droplets to permeate and contact with each other,

forming a large network of micro-channels (Figure 2.5d). During the thawing process, this

network fuses with water droplet coalescence that leads to phase inversion (Figure 2.5e),

and such unstable oil-water mixture can then be delaminated into bulk phases driven by

gravitational force (Figure 2.5f) (Lin et al., 2008).

47

Figure 2.5 Schematic diagram of the mechanism of freeze/thaw-induced demulsification

for W/0 emulsion: (a) original emulsion, (b) water droplets freezing,

expansion and coalescing, ( c) oil phase freezing to form a solid cage, ( d) water

droplets freezing and expanding to break the cage, ( e) emulsion thawing and

water droplets coalescing, and ( f) gravitational de lamination

Jean et al. (1999) found that over 50% of oil can be separated from refinery oil-in-

water emulsions by freeze/thaw method. Chen and He (2003) reported that freeze/thaw

treatment removed nearly 90% of water from a high-water-content oily sludge. In another

research (He and Chen, 2002), freeze/thaw showed satisfactory performance on the

dewatering of lubricating oily sludge, and over 90% of water was removed using this

method. Lin et al. (2008) applied freeze/thaw treatment to break water/oil emulsions, and

the volume expansion of water turning to ice and interfacial tension of oil-water interface

were determined as the main driving forces of demulsification. They investigated the effect

48

of four different freezing methods (i.e. freezing in a refrigerator, cryogenic bath, dry ice

and liquid nitrogen), and found that the best freezing method was freezing in cryogenic or

dry ice, with over 70% dehydration efficiency being achieved for the emulsions containing

60% of water. Zhang et al. (2012) found that freeze/thaw worked effectively for oil

recovery from refinery sludge with an oil recovery rate of 65.7%. In general, the

performance of freeze/thaw on demuslification can be affected by a number of factors,

such as freezing and thawing temperatures, treatment duration, water content, salinity of

aqueous phase, presence of surfactants, and solid contents in emulsion (Lin et al., 2007;

Zhang et al., 2012). For example, Chun et al. (2004) found that the dewatering rate at -

40 °C was obviously higher than that at -20 °C, and the lower thawing temperature was

also beneficial to dewatering process. At high thawing temperature, the frozen sludge

could be melted rapidly which may break down the aggregated floes of oil/solids, thus

leading to poor dewaterability. Yang et al. (2009) compared the effect of three thawing

approaches on W /0 demulsification, and they found that rapid thawing (i.e. microwave

heating) could significantly enhance the water-oil separation efficiency, with more than

90% (v/v) of emulsion being separated. Similar results can be found in Rajakovic and Skala

(2006). Oh and Deo (2011) also proved that water content can affect oil recovery, with

higher yield stresses being observed in W/0 emulsion with lower water content. In

summary, freeze/thaw treatment is a promising method for dewatering and oil recovery

from oily sludge. However, its industrial application should consider the required freezing

time and related costs. Chun et al. (2004) suggested that 8 h was almost sufficient for

freezing at -20 °C. In addition, freezing could be a relatively slow process that requires

intensive energy and high cost (Jean and Lee, 1999). Thus, the application of freeze/thaw

49

treatment for oil recovery from oily sludge might be more promising for cold regions

where natural freezing is possible.

2.4.5 Pyrolysis treatment

Pyrolysis is the thermal decomposition of organic materials at elevated temperatures

(500-1000 °C) in an inert environment. It is different from gasification which transforms

organic materials to combustible gas or syngas with the existence of 20-40% of oxygen.

The pyrolysis process produces hydrocarbons with lower molecular weight in

condensation (i.e. liquid) and/or non-condensable gases. It also generates a solid product

called char (Fonts et al. , 2012). Depending on the operational conditions, the main product

of pyrolysis can be either char, liquid, or gas, and they may have a more elevated heating

value than the raw oily sludge (Zhang et al. , 2005). For example, the major product of fast

pyrolysis treatment (i.e. a pyrolysis process that rapidly heats feedstock to a controlled

temperature and then very quickly cools the volatile products formed in the reactor) can

be a liquid (i.e. pyrolysis oil), which could be used as a fuel or a source of other valuable

chemical products (Liu et al. , 2011; Bridgwater et al. , 1999). Several research studies have

been reported to use pyrolysis for fuel recovery from oily sludge. Shen and Zhang (2003)

observed that oil yield increased initially with pyrolysis temperature with a maximum oil

yield (30 wt.% of the feed oily sludge) occurring at 525 °C, but decreased when the

temperature was above 525 °C due to secondary decomposition reactions which could

break the oil into lighter and gaseous hydrocarbons. Liu et al. (2009) found that about 80%

of total organic carbon content (TOC) in oily sludge could be converted into usable

hydrocarbons when using a pyrolysis process, with a significant hydrocarbon yield

occurring in the temperature range of 327-450 °C. Schmidt and Kaminsky (2001) found

50

that the separation of oil from oily sludge occurred from 460 to 650 °C, and 70 to 84% of

the oil could be separated from sludge by a fluidized bed reactor. Chang et al. (2000)

applied the pyrolysis process to treat oily sludge, and they observed a maximum production

rate of hydrocarbons (mainly low-molecular-weight paraffins and olefins, 51.61 wt% of

PH Cs) at 440 °C, and the distillation characteristics of liquid product from pyrolysis were

close to that of diesel oil. Their results also indicated that the major gaseous products from

pyrolysis excluding N2 are CO2 (50.88 wt%), hydrocarbons (25.23 wt%), H20 (17.78 wt%),

and CO (6.11 wt%) (Chang et al., 2000). Karayildirim et al. (2006) illustrated that the main

decomposition of oily sludge occurred in the temperature range of 100-350 °C, while the

inorganic materials started decomposition when the temperature rose to 400 °C, and the

carbonaceous residues accounted for 38 wt% of the original sludge at the final pyrolysis

temperature of 900 °C. Wang et al. (2007) found that pyrolysis of oily sludge started at a

low temperature of 200 °C, and the maximum hydrocarbon production occurred in the

range of 350-500 °C, with improved oil yield and quality being observed by maintaining

temperature at 400 °C for 20 min.

Pyrolysis can be affected by a number of factors, such as temperature, heating rate,

characteristic of oily sludge and chemical additives. Punnaruttanakun et al. (2003)

investigated the influence of different heating rate (i.e. 5, 10 and 20 °C/min) on the

pyrolysis of API separator sludge, and they found a lower pyrolysis rate at a heating rate

of 20 °C/min than those at 5 and 10 °C/min, but the heating rate did not affect the amount

of solid products. Shie et al. (2003) found that chemical additives (such as sodium and

potassium compounds) in the pyrolysis process could enhance the reaction rate within a

pyrolysis temperature range of 377-437 °C, and the highest fuel yield was obtained as

73.13 wt.% with the addition of KCl, while the maximum improvement effect on the

51

quality of pyrolysis oil was achieved by KOH, then followed by KCl, K2C03, NaOH,

Na2C03, and NaCl. Results from other studies also indicated that additives of metal

compounds (i.e. aluminum and iron compounds) and catalytic solid wastes (i.e. fly ash,

oily sludge ash, waste zeolite and waste polymer of polyvinyl alcohol) could affect the

conversion, reaction rate, yield and quality of oil products from oily sludge pyrolysis

process (Shie et al., 2003; 2004). In terms of application, three types of pyrolysis

configuration can be used, including ablative pyrolysis, fluid bed and circulating fluid bed

pyrolysis, and vacuum pyrolysis. The commercial-scale oil recovery application has

mainly adopted the fluid beds or circulating fluid beds and the associated auxiliary systems,

such as sludge and nitrogen feeding, char collection, and vapor condensation (Figure 2.6)

(Fonts et al., 2012).

I

Oily sludge Cyclone & fly Liquid .... ----? '

l ., ash collector condensing .,

system

Fluidized Oily sludge bed reactor w

... Liquid product pretreatment

,

~ Char collector Inertial gas .... .,

Figure 2.6 Typical schematic view of fluidized bed systems used in sludge pyrolysis treatment

Gaseous product

Pyrolysis has the advantages of producing a liquid product that can be easily stored

and transported, and the recovered oil was proved to be comparable to low-grade

52

petroleum distillates from commercial refineries and could be directly used in diesel

fuelled engines (Czernik and Bridgewater, 2004; Chiaramoni et al. , 2007) . As compared

to the incineration process, pyrolysis of oily sludge generates a lower emission ofNOx and

SOx, and it also enables heavy metals in oily sludge to be concentrated in the final solid

product (i.e. char) (Shen and Zhang, 2003). The char usually accounts for 30 to 50 wt% of

the original oily sludge, and it can be applied as an adsorbent for the removal of different

pollutants such as H2S or NOx in gaseous streams (Fonts et al., 2012). It can also be used

as a soil conditioner to increase the nutrient availability for plants (Lehmann and Rondon,

2006). The metals enriched in solid char can be more resistant to leaching than those

concentrated in the ash obtained from incineration (Mohan et al. , 2007). In spite of these

advantages, the large-scale implementation of pyrolysis could be restricted by the low

economic value of liquid products and the relatively complex processing equipments

(Bridle and Pritchard, 2004; Lim and Parker, 2008). The high operational cost is mainly

due to the large amount of external energy required for the endothermic reaction to take

place (Fonts et al. , 2012). In addition, oily sludge usually contains a relatively high water

content, and thus the dewatering of oily sludge before pyrolysis treatment may

significantly increase the overall cost of pyrolysis (Bridle and Unkovich, 2002). Additional

concern may include that P AHs, the well-known highly carcinogenic substances, could be

existed as a large portion in the liquid products of sludge pyrolysis (Kwah et al. , 2006).

2.4.6 Microwave irradiation

The microwave frequency ranges from 300 MHz to 300 GHz, but the industrial

application is usually performed at a frequency either close to 900 MHz or near 2450 MHz

(Appleton et al. , 2005). Microwave energy can directly penetrate the material through

53

molecular interaction with the electromagnetic field, and provide a quick heating process

at improved heating efficiencies as compared with conventional techniques. Such heating

effect can be used for the demulsification of W/0 emulsion by rapidly increasing the

temperature of emulsions, leading to the reduction of viscosity which could accelerate the

settlement of water droplets in emulsion (Tan et al. , 2007). The rapid temperature increase

can also break heavy hydrocarbons into lighter ones. For materials with low dielectric loss,

microwave can pass through them with little energy absorption. For materials with high

dielectric loss, microwave energy can be absorbed based on the electric field strength and

the dielectric loss factor. When using microwave for treating a mixture of materials with

different dielectric properties, a selective heating could occur (Chan and Chen, 2002;

Shang et al., 2006). For W/0 emulsion such as oily sludge, the inner phase is water with a

relatively higher dielectric loss, and it can absorb more microwave energy than oil. Such

energy absorption could result in the expansion of water and press the water-oil interfacial

film to become thinner, which could facilitate water/oil separation (Tan et al., 2006).

Moreover, microwave irradiation could lead to molecular rotation by rearranging the

electrical charges surrounding water molecules. This could destroy the electric double

layers at the oil/water interface, resulting in the reduction of zeta potential. Under reduced

zeta potential, water and oil molecules can move more freely in the emulsion so that the

water or oil droplets can collide with each other to form coalescence (Tan et al., 2006).

The above mechanisms can lead to the separation of an emulsion (Kuo and Lee, 2010).

Several laboratory- and field-scale studies have demonstrated the benefits of using

microwave irradiation for W/0 emulsion treatment. Xia et al. (2003) observed a nearly

100% of demulsification efficiency within a very short time by using microwave radiation

for treating W/0 emulsions, which is much higher than that when using conventional

54

heating approaches. Fang and Lai (1995) applied microwave irradiation for a field test to

demulsify 188 barrels ofW/0 emulsion in tanks, and their results showed that the emulsion

was separated into 146 barrels of oil and 42 barrels of water, while the water separation

efficiency from emulsions was higher than that when using conventional heating. They

also illustrated that microwave irradiation could have particular effect on the partial

removal of polar compounds (Fang and Lai, 1995).

The performance of microwave irradiation on oily sludge demulsification can be

affected by many factors such as microwave power, microwave duration, surfactant, pH,

salt, and other properties of oily sludge such as water-oil ratio (Fortuny et al., 2007). The

increase of pH could decrease the stability of sludge W /0 emulsion due to the increase in

hydrophilicity (Fortuny et al., 2007). On the other hand, the salt-assisted microwave

irradiation has been found to be effective due to increased conductivity of water phase that

could accelerate the heating rate (Xia et al., 2004). For example, Chan and Chen (2002)

found that electrolyte could lower the Zeta potential of oil droplets and result in the

destabilization of W /0 emulsions, and they observed that the demulsification rate

increased with electrolyte concentration (NaCl, KCl, NaN03, and Na2S04) in the range of

dilute solutions (< 0.5 M) when using microwave irradiation. The optimum microwave

irradiation power and treatment time were suggested to be 420 Wand 12 seconds in their

research (Chan and Chen, 2002). Fortuny et al. (2007) investigated the demulsification

effect of salt (Na Cl) on a crude oil emulsion by microwave irradiation, and they found that

the dissolved salts significantly increased the heating efficiency and destabilization of

emulsions. Tan et al. (2007) found that the performance of microwave treatment on W/0

emulsions could be improved by the addition of chemical demulsifiers, and a satisfactory

water-oil separation efficiency (95v/v%) was achieved by adding 50 ppm of demulsifier

55

when using microwave irradiation (700 W, 2450 MHz). Abdulbari et al. (2011) combined

microwave heating (900 W, 2450 MHz) with synthetic surfactants (i.e. Triton X-100, low-

sulfur waxy residue, sorbitan monooleate, and SDS) to separate water from petroleum

emulsions, and their results indicated that over 90% of water can be separated. In general,

as compared to other heating methods, microwave irradiation can more rapidly raise the

energy of molecules inside the medium, leading to higher reaction rates and superheating

within several minutes (Robinson et al., 2008). The short heating time makes microwave

irradiation a high energy-efficient and easy-to-control approach for breaking emulsions. In

addition to its high demulsification efficiency, the low temperature of reactor wall during

the direct heating inside the bulk medium caused by microwave irradiation could lead to

an extended aromatization reaction. This could result in an increased yield oflight aromatic

compounds. These light compounds can have a much lower toxicity as compared to P AHs

with high molecular weight in the liquid products generated during the pyrolysis process

(Dominguez et al., 2005). However, the application of microwave irradiation to industrial-

scale oily sludge treatment is limited due to the specific equipment required and possible

high operating costs.

2.4. 7 Electrokinetic method

Electrokinetic process utilizes a low-intensity direct current across an electrode pair

on each side of a porous medium, causing the electro osmosis of liquid phase, migration

of ions and electrophoresis of charged particles in a colloidal system to the respective

electrode (Virkutyte et al. , 2002; Yang et al. , 2005). The separation of different phases

(water, oil, and solids) from oily sludge using electrokinetic process can be based on three

main mechanisms. Firstly, colloidal aggregates in oily sludge can be broken under the

56

influence of an electrical field, leading to the movement of colloidal particles of oily sludge

and solid phase towards the anode area as a result of electrophoresis, and the movement of

the separated liquid phase (water and oil) towards the cathode area as a result of electro-

osmosis (Yang et al., 2005). As mentioned before, oily sludge is a W /0 emulsion stabilized

by several kinds of emulsifiers such as asphaltenes, resins, organic acids, and finely

divided solids. These emulsifiers usually occur as a film on the surface of dispersed water

droplets, and the fine particles can be adsorbed at the water droplet surface and act as a

barrier to prevent droplet coalescence (Ali and Alqam, 2000). The separation of colloidal

particles and fine solids from W /0 emulsion could remove such barrier and thus accelerate

the coalescence of water droplets in continuous oil phase. Secondly, following this process,

the electro-coagulation of the separated solid phase could occur near the anode area,

leading to increased solid phase and sediments concentration. Lastly, the separated liquid

phase (water and oil, without colloid particles and fine solids) can produce an unstable

secondary oil-in-water emulsion which could be gradually electro-coalesced near the

cathode area through charging and agglomeration of droplets, thus forming two separated

phases of water and oil (Elektorowicz et al., 2006).

The electrokinetic treatment performance can be affected by several factors such as

resistance, pH, electrical potential, and spacing between electrodes. This process may be

improved through the use of surfactants or reagents to increase the contaminant removal

rates at the electrodes (Electorowicz and Hatim, 2000). Elektorowicz and Habibi (2005)

applied electrokinetic process to treat oily sludge, and they found that this process could

reduce the amount of water by nearly 63% and light hydrocarbon content by about 43%,

and the light hydrocarbon content was removed by 50% when combining electrokinetic

treatment with surfactant. Yang et al. (2005) conducted an oily sludge dewatering study

57

using electrokinetic treatment, and their results showed that the water removal efficiency

reached to 56.3% at a 4 cm electrode spacing and an electric potential of 30 V, while the

solids content at the anode area increased from 5 to 14.1 %. A larger-scale experimental

study (i.e. a capacity of 4 L) revealed that more than 40% of water was removed and a very

efficient oil separation from oily sludge was achieved using electrokinetic process at 60 V

with an initial spacing of 22 cm (Ali and Alqam, 2000), and the temperature in the system

was observed to increase significantly in a short time due to the application of high electric

current. In general, the application of electrokinetic process for oil recovery from oily

sludge requires less amount of energy than other oil recovery methods such as

centrifugation and pyrolysis. However, most of the electrokinetic studies on oily sludge

have been conducted at the laboratory level, and the performance and costs at a large scale

still need further investigation. It is expected that using oily sludge storage pools as the

electrokinetic cell at the field scale could considerably reduce the treatment cost

(Elektorowicz et al., 2006).

2.4.8 Ultrasonic irradiation

Ultrasonic irradiation has proved to be effective for removing adsorbed materials

from solid particles, separating solid/liquid in high-concentration suspensions, and

decreasing the stability ofW/0 emulsion (Li et al., 2013; Song et al., 2012; Kim and Qang,

2003; Ye et al., 2008). When ultrasonic wave propagates in the treatment medium, it

generates compressions and rarefactions. The compression cycle exerts a positive pressure

on the medium by pushing molecules together. The rarefaction cycle exerts a negative

pressure by pulling molecules from each other, and microbubbles can be generated and

grow due to such negative pressure. When these microbubbles grow to an unstable

58

dimension, they can collapse violently and generate shock waves, resulting in very high

temperature and pressure in a few microseconds (Pilli et al. , 2011). Such cavitation

phenomenon can increase the temperature of the emulsion system and decrease its

viscosity, increase the mass transfer of liquid phase, and thus lead to destabilization of

W/0 emulsion (Chung and Kamon, 2005). Other studies suggested that under the influence

of ultrasonic irradiation, smaller droplets in emulsion can move faster than the larger ones,

and this can increase their collision frequency to form aggregates and coalescence of

droplets, which then promotes the separation of water/oil phases (Yang et al. , 2009;

Gholam and Dariush, 2013). Ultrasonic irradiation can not only clean the surface of solid

particles but also penetrate into different regions of a multiphase system that are

inaccessible when using other separation methods (Swamy and Narayana, 2001). This

mechanism is called ultrasonic leaching which enables solvents or leaching reagents to

more readily enter the interior of solids pores, and increases the mass transfer of

contaminants through the solids matrix (Feng and Aldrich, 2000). A number of studies

have been conducted to investigate the efficiency of ultrasonic irradiation for oil recovery

from oily sludge. Xu et al. (2009) applied ultrasonic cavitation with a frequency of28 kHz

in an ultrasonic cleaning tank to strip oil constituents from the surface of solid particles in

oily sludge, and an overall oil separation rate of 55.6% was obtained. They also found that

the optimal temperature, acoustic pressure, and ultrasonic power for oil recovery from

sludge were 40 °C, 0.10 MPa, and 28 kHz, respectively, while both too high or too low

ultrasonic energy input could inhibit oil recovery process since high ultrasonic energy

input can prevent oil droplets from merging and low ultrasonic energy input makes it

difficult to detach oil from solid particles. Zhang et al. (2012) reported an oil recovery rate

59

of up to 80% from oily sludge-water matrix (sludge/water ratio of 1:2) after 10 min of

ultrasonic treatment using a 20 kHz ultrasonic probe system at a power of 66 W.

In general, the performance of oil recovery from oily sludge using ultrasonic

treatment can be affected by a variety of factors, such as ultrasonic frequency, sonication

power and intensity, water content in emulsion, temperature, ultrasonic treatment duration,

solid particle size, initial PHCs concentration, salinity, and presence of surfactant (Kim

and Wang, 2003; Feng and Aldrich, 2000; Na et al. , 2007) . For example, Xu et al. (2009)

found that lower ultrasonic frequency is more favorable for oily sludge treatment since

cavitation is more difficult to occur under high frequency ultrasound than that under low

frequency ultrasound, and they also indicated that too high or too low temperature is not

suitable for oily sludge treatment by ultrasound. In the research by Jin et al. (2012), a high

oil recovery rate (above 95%) was achieved when the ultrasonic treatment parameters were

28 kHz, 15 min, 400 Wand 60 °C, respectively, and no further enhancement of oil recovery

was observed when the ultrasonic power and treatment duration increased beyond 400 W

and 15 min. Overall, ultrasonic irradiation is a "green" treatment method which can process

oily sludge within a relatively short time. In spite of its high efficiency of oil recovery and

no secondary pollution, the application of ultrasonic irradiation to field-scale oil recovery

from sludge has rarely been reported, while the most commonly used laboratory ultrasonic

irradiation system was the ultrasonic probe system which can only be effective when

treating a small volume of oily sludge. The utilization of large ultrasonic cleaning tank

could be more promising in terms of treating a large amount of oily sludge, but its oil

recovery performance might be compromised due to the resulted low ultrasonic intensity

(Canselier et al. , 2007). The high cost of equipment and maintenance could also inhibit the

industrial application of this technology.

60

2.4.9 Froth flotation

Froth flotation is a surface chemistry-based unit operation for separating fine solid

particles from an aqueous suspension. It involves the capture of oil droplets or small solids

by air bubbles in an aqueous slurry, followed by their levitation and collection in a froth

layer (Urbina, 2003). Froth flotation has been successfully used to treat oily wastewater

from refineries, and its utilization for oily sludge treatment has recently received research

attentions (Ramaswamy et al. , 2007). Figure 2.7 presents a schematic overview of this

process. Oily sludge is firstly mixed with a given amount of water to form a sludge slurry.

Air injection then generates fine air bubbles which approach oil droplets in the slurry

mixture, and the water film between oil and air bubble can get thinning to reach a critical

thickness, causing the rupture of water film and the spreading of oil to air bubbles. Lastly,

the conglomerate of oil droplets with air bubbles can quickly rise to the top of water/oil

mixture, and the accumulated oil can be skimmed off and collected for further purification

(Moosai and Dawe, 2003). Ramaswamy et al. (2007) applied froth flotation treatment for

oil-water separation from oily sludge, and they observed an oil recovery of up to 55% at

the optimal flotation conditions. Addition of surfactants and extraction solvent could

enhance the oil recovery performance of flotation. Al-Otoom et al. (2010) applied a

modified fluidized flotation process for bitumen recovery from tar sand, and the maximum

bitumen recovery could reach 86 wt% when 0.35 wt% light cycle oil was added as an

extraction solvent at 80 °C. Stasiuk et al. (2001) found that the addition of surfactants such

as Tween 80 or Alcopol O during the froth flotation process significantly reduced water

contents (i.e. 10 v/v%) in the recovered oil.

61

• D • • • • • • • t Water phase

• C ··viu,. Feed water

Figure 2.7 Schematic view of froth flotation in oil/water separation process (A: Air

control valve; B: Motor; C: Stirrer; D: Skimmed oil collector)

In general, oil recovery performance using froth flotation can be affected by many

factors, such as oily sludge properties (i.e. viscosity, solid content and density), pH, salinity,

temperature, size of air bubble, presence of surfactant, and flotation duration (Al-Shamrani

et al., 2002; Faksness et al., 2004). Ramaswamy et al. (2007) reported that the oil recovery

rate increased with the duration of flotation but approached its maximum after 12 min of

treatment, and they also found that surfactant had significant influence on the oil recovery

process. Higher temperature could also enhance the oil recovery via flotation since

decreased sludge viscosity would facilitate oil separation and the subsequent flotation (Al-

Otoom et al., 2010). Froth flotation is a simple and less expensive approach for oily sludge

treatment, but it is usually effective when treating sludge with relatively low viscosity. It

has limited effect on the desorption of oil from solid matrix, and the oil constituents in

skimmed oil/solids mixture still need to be further purified. The recovered oil could also

contain relatively high moisture. Consequently, a number of limitations still exist when

62

Oil film

applying it to field-scale oily sludge treatment. Oily sludge needs to be pretreated to reduce

its viscosity and remove the coarse solid particles. Moreover, the froth flotation requires a

large volume of water when treating oily sludge with low moisture and high viscosity, and

thus oily wastewater treatment problem can be generated.

2.5 Oily Sludge Disposal Methods

In addition to oil recovery, a number of technologies are available for the disposal of

oily sludge, including incineration, stabilization/solidification, oxidation, and

biodegradation.

2. 5 .1 Incineration

Incineration is a process of complete combustion of oily wastes in the presence of

excess air and auxiliary fuels, and it is widely adopted in large refineries for sludge

treatment. Rotary kiln and fluidized bed incinerator are the most commonly used

incinerators. In rotary kiln incinerator, the combustion temperature is in the range of 980-

1200 °C and the residence time is around 30 min. In fluidized bed incinerator, the

combustion temperature can be in the range of 730-760 °C, and the residence time can be

in order of days (Scale and Chirone, 2004 ). Fluidized bed incinerator is especially effective

when treating low-quality sludge due to its fuel flexibility, high mixing efficiency, high

combustion efficiency and relatively low pollutant emissions (Zhou et al., 2009). The

incineration performance can be affected by a variety of factors, including combustion

condition, residence time, temperature, feedstock quality, presence of auxiliary fuels, and

waste feed rates. Liu et al. (2009) investigated the incineration of oily sludge by the

addition of an auxiliary fuel consisting of coal-water slurry (CSW) in a fluidized bed boiler,

63

and they found that the temperature of combustion could be stabilized by controlling the

CSW feeding rate, with a combustion efficiency of 92.6% being reached. They also found

that the gaseous emission from incineration and heavy metals in ash residues could meet

the corresponding environmental regulations. Wang et al. (2012) examined the

incineration of a petroleum coke-sludge slurry (PCSS) which was obtained by blending

oily sludge with a petroleum coke-water slurry (PCWS), and they found that the PCSS fuel

exhibited low viscosity and satisfactory stability for combustion process, but the

combustion efficiency, gaseous emission, and heavy metals in ash residues were not

investigated in their study. Sankaran et al. (1998) investigated the direct combustion of

three types of oily sludge in a fluidized bed incinerator without auxiliary fuels, and a high

combustion efficiency of 98-99% was observed for sludge with low water content, but the

combustion was difficult to occur for sludge with a relatively high water content (i.e. >

51 %). They also found that oily sludge was too viscous to flow as the feedstock, and a pre-

treatment such as heating was required to reduce the viscosity before incineration

(Sankaran and Pandey, 1998).

Through combustion in an incinerator, oily sludge can provide a valuable source of

energy, which can be used to drive steam turbines and as a heat source in a waste oil

reclamation factory. Moreover, a significant reduction in the volume of waste can be

achieved after incineration treatment. Although the oily sludge incineration has been

practised in a few developed countries (Naranbhai and Sanjay, 1999), it suffers from a

number of limitations. Oily sludge with high moisture needs to be pre-treated to improve

its fuel efficiency by lowering the excessive water content (Al-Futaisi et al. , 2007).

Auxiliary fuel is usually required to maintain a constant combustion temperature (Li et al. ,

1995; Bhattacharyya and Shekdar, 2003). In addition, fugitive emission of pollutants (e.g. ,

64

low molecular P AHs) from incineration and incomplete combustion could cause

atmospheric pollution problem (Li et al., 1995). Moreover, ash residue, scrubber water,

and scrubber sludge generated during the incineration process are hazardous and need

further treatment. In general, the oily sludge contains high concentration of hazardous

constituents that are resistant to combustion, and the incineration requires high capital and

operating costs, with a cost of more than $800 per ton of oily sludge incineration was

reported (Shiva, 2004).

2.5.2 Stabilization/solidification

Stabilization/solidification (S/S) is a quick and inexpensive waste treatment technique

aimed at immobilizing contaminants by converting them into a less soluble or a less toxic

form (i.e. stabilization), and encapsulating them by the creation of a durable matrix with

high structural integrity (i.e. solidification) (Malviya and Chaudhary, 2006). The use of

this disposal method for inorganic wastes has been widely reported (Leonard and

Stegemann, 2010). However, S/S was considered less compatible with organic wastes

since organic compounds may inhibit cement-based binder hydration and are generally not

chemically bound in the binder hydration products (Malviya and Chaudhary, 2006). As a

result, immobilization of organic contaminants depends mainly on physical entrapment in

the binder matrix and sorption onto the surface of binder hydration products. There is a

possibility of releasing undesirably high concentration of pollutants when exposed to

environmental leachants. Karamalidis and Voudrias (2007) studied the leaching behavior

of TPH, alkanes and P AHs from refinery oily sludge stabilized/solidified with Portland

cement, and their results showed that the waste was confined in the cement matrix by

macro-encapsulation, but they also reported that increased cement addition to oil refinery

65

sludge led to higher concentrations in leachates from batch extraction. Other leaching test

studies of oily sludge S/S treatment using Portland cement have shown relatively high

release of PAHs (Conner and Hoeffner, 1998) and methanol and 2-chloroaniline (Gussoni

et al., 2004). In general, a Portland cement only binder system is not effective for the

immobilization of several common organic contaminants. A possible method for

improving the effectiveness of S/S treatment for organic wastes is to use binders that

increase sorption of organic compounds ( e.g., combined use of cement and activated

carbon), thereby improving their immobilization and preventing their detrimental effects

on binder hydration. Leonard and Stegemann (2010) found that Portland cement with the

addition of high-carbon power plant fly ash (HCFA) significantly reduced the leaching of

PH Cs.

In addition to the immobilization of organic contaminants, an advantage of applying

S/S method is that some hazardous heavy metals in oily sludge can be immobilized into

the cement matrix. Karamalidis and Voudrias (2007) evaluated the leaching behavior of

heavy metals from stabilized/solidified refinery oily sludge and ash produced from oily

sludge incineration with Portland cement (OPC), and an immobilization of > 98% was

observed for metals of solidified ash at pH > 6 and of > 93% of solidified oily sludge at

pH > 7. They found that pH had the strongest influence on the immobilization of heavy

metals during the S/S process, and an extremely high Ni immobilization (> 98%) was

observed for solidified oily sludge samples at pH > 8, but the immobilization was very low

(47%) at pH of 2.5. Al-Futaisi et al. (2007) solidified tank bottom sludge mixtures using

three combinations of selected additives such as Portland cement (OPC), cement by-pass

dust (CBPD) and quarry fines (QF), and the toxicity characteristic leaching procedure

(TCLP) analyses results revealed that no extracts exceeded the regulated TCLP maximum

66

limits of metals. Although S/S technique has shown effectiveness on immobilizing

inorganic and organic contaminants in oily sludge, little is known from the current

literatures regarding the physical strength of S/S treated oily sludge and its cost for

handling a large volume of oily sludge. Further researches, including the use of other

binders such as pozzolanic substances ( e.g., fly ash) and other cement-like materials in the

S/S of oily sludge, are also worth to be conducted.

2.5.3 Oxidation treatment

Oxidation treatment has been used to degrade a range of organic contaminants

through chemical or other enhanced oxidation processes (Ferrarese et al., 2008). Chemical

oxidation involves introducing reactive chemicals into oily wastes to oxidize organic

compounds to carbon dioxide and water, or transform them to other nonhazardous

substances such as inorganic salts (Ferrarese et al., 2008). The oxidation can be brought

by Fenton ' s reagent, hypochlorite, ozone, ultrasonic irradiation, permanganate, and

persulfate, by generating a sufficient amount of radicals such as hydroxyl radicals (OH•)

which can quickly react with most organic and many inorganic compounds (Rivas, 2006).

Many studies have proven that chemical oxidation can effectively degrade PHCs and

P AHs in soils, and this method has recently been applied to oily sludge treatment. Mater

et al. (2006) found that a Fenton type reagent (i.e. 12 wt% of H202 and 10 mM of Fe2+) at

a low pH (i.e. pH= 3.0) can significantly reduce the concentrations of PAHs, phenols and

BTEXs in oily sludge contaminated soil. Ultrasonic irradiation has also recently been

investigated for its efficiency on oily sludge oxidation process. The sonication of water

can produce intermediate radicals such as hydrogen (H•), hydroxyl (OH•), hydroperoxyl

(H02•) and hydrogen peroxide (H202) with high oxidization power in and around the

67

cavitation bubbles (Mason, 2007). Through sonolysis reaction by these free radicals, long-

chain or aromatic petroleum hydrocarbons with complex structure and large molecular

weight can be broken into simple hydrocarbons which have higher solubility and

bioavailability. Ultrasound can be used to degrade many PHCs, such as chlorinated

aliphatic hydrocarbons (CAHs), aromatic compounds, polychlorinated biphenyls (PCBs),

poly aromatic hydrocarbons (PAHs), and various phenols (Adewuyi, 2001; Dewulf and

Langenhove, 2001; Collings et al., 2006; Lim et al., 2007). Zhang et al (2013) utilized a

combined process of ultrasonic and Fenton oxidation for the oily sludge treatment, and it

was found that ultrasonic irradiation could enhance the Fenton oxidation effect on oily

sludge degradation by improving the contact of hydroxyl radicals (OH•) with PHCs

compounds.

Other advanced oxidation methods such as supercritical water oxidation (SCWO),

wet air oxidation (W AO), and photocatalytic oxidation (PO) have been reported in recent

literatures of oily sludge treatment. SCWO method uses water above its critical point

(375 °C, 22.1 MPa) as a reaction medium where gases, oil and aromatics form a single

homogeneous phase, and the oxidation could proceed quickly and completely by

converting most H-C-N compounds to water, carbon dioxide, and molecular nitrogen

(Fujii et al., 2011). Cui et al. (2009) applied the SCWO method for oily sludge treatment,

and their results indicated that 92% of chemical oxygen demand (COD) in oily sludge was

removed after only 10 min of treatment. The W AO technology uses oxygen as the oxidant

under high temperature and high pressure to convert hazardous organic compounds to CO2,

H20 and other innocuous end products. Jing et al. (2011) found that the WAO process

could remove 88.4% of COD in oily sludge within 9 min of reaction at temperature of

330 °C and an 02 excess of 0.8, and the COD removal can be increased to 99.7% with the

68

addition of Ni2+ catalyst. Fe3+ catalyst can also improve the COD removal during oily

sludge treatment using W AO process. It was reported that a very promising COD removal

rate of 95.4% was achieved under the catalytic effect of Fe salt (i.e. 50 mg/L of Fe3+)

compared to that of 84.12% in the treatment without the addition of Fe3+ catalyst (Jin et

al., 2012).

Another enhanced oxidation treatment is the heterogeneous photocatalytic oxidation

which is based on the activation of the surface of a photo-conductor (e.g., titanium dioxide

in the Anatase form) by light (e.g., UV light or sunlight). The absorption of photons with

energy higher than the band gap of the photo-conductor produces pairs of conduction band

electrons ( e-) and valence band vacancies (h +). After this first step, the charge carriers either

recombine within the bulk of the material or migrate to the particle surface, where they can

react with adsorbed electron donors (D) and acceptors (A) species (Mazzarino and

Piccinini, 1999). The presence of oxygen and water is essential to the photocatalytic

process. Oxygen serves as the e- acceptor which can form the superoxide radical anion Oi"

•. The surface-bound water and the surface hydroxyl groups are oxidized by the vacancies

to form OH• radicals which are supposed to be the most active oxidizing species. da Rocha

et al. (2010) applied the heterogeneous photocatalytic oxidation (H202/UV/Ti02) for oily

sludge treatment, and they found that white light catalyzed oxidation was able to eliminate

100% of P AHs in sludge after 96 h of treatment. They also revealed that white light was

more effective than black light in terms of catalyzing oxidation on P AHs. However, the

effect of photocatalytic oxidation on other constituents such as asphaltenes and resins in

oily sludge was not investigated in their study. In general, oxidation requires a relatively

short treatment duration to degrade oily sludge, and it is relatively insensitive to external

disturbances ( e.g. , pollutant loading, temperature change, and the presence of biotoxic

69

substances, etc.). Its reaction products are usually more biodegradable than the raw waste

materials (Jin et al. , 2011). However, when treating a large volume of oily sludge, the

oxidation may require a large amount of chemical reagents. The advanced oxidation

methods such as W AO, SCWO and PO also require special equipments and considerable

energy inputs which could increase the treatment cost.

2.5.4 Bioremediation

Bioremediation is defined as the process of using micro-organisms to remove

environmental pollutants, and is commonly employed for the restoration of oil-polluted

environments through accelerating the microbial degradation of PHCs (Luquefio et al. ,

2011). The most intensively studied bioremediation approaches include land treatment,

biopile/composting, and bio-slurry treatment (Powell et al., 2007).

2. 5. 4.1 Land treatment

Land treatment involves the incorporation of wastes into soil and then the use of

various processes to degrade contaminants in that soil (Hejazi et al., 2003). Biological

activity usually accounts for most of the degradation of organic pollutants, while physical

and chemical removal mechanisms such as evaporation and photo-degradation may also

be important for some compounds. Landfarming is a widely employed land treatment

approach, and it spreads the well-mixed oily sludge and fresh soil on the ground surface

of a treatment site. The treatment efficiency can be improved by maintaining appropriate

sludge application rate, aeration, fertilization, moisture content, and pH to maintain

microbial density and enhance their activity in the sludge/soil mixture. Marin et al. (2005)

applied landfarming method to clean up oil refinery sludge in a semi-arid climate, and their

results showed that 80% of the PHCs were removed within 11 months of treatment, while

70

half of this removal occurred during the first three months. Admon et al. (2001) also

observed similar degradation pattern through experiments on the landfarming of refinery

oily sludge, and 70-90% of PH Cs degradation occurred within 2 months, while a relatively

high biodegradation activity was observed in the first 3 weeks of treatment. Hejazi and

Husain (2004) compared the influence of three operating parameters (i.e. tilling, addition

of water, and addition of nutrients) on a 12-month oily sludge land treatment under arid

conditions, and they found that tilling was the main parameter responsible for achieving

the highest PHCs removal rate of 76%. Mishra et al. (2001) combined bioaugmentation

(i.e. introducing extraneous bacterial consortium) and biostimulation (i.e. addition of

nutrients and water) for oily sludge decontamination, and their results indicated that up to

90.2% of TPH can be degraded in 120 days in the land block which had extraneous

bacterial consortium inoculum and nutrients enhancement, while only 16.8% ofTPH was

eliminated in the control land block.

As compared to other oily sludge disposal technologies, landfarming holds many

merits such as relatively low capital costs, simple operation, high potential for success, low

energy consumption, and the capability of treating a large volume of oily sludge (Hejazi

et al. , 2004). However, landfarming of oily sludge requires a large area of land, and is a

very time-consuming process (e.g. , usually 6 months to 2 years or even longer) since the

soil/sludge conditions favoured by biodegradation are largely uncontrolled, particularly for

recalcitrant and heavy PHC compounds. In addition, temperature can greatly affect the

contaminant biodegradation efficiency, and the performance of oily sludge landfarming in

cold regions can be compromised. Moreover, landfarming can bring various environmental

issues, such as the emission of volatile organic compounds (VOCs) and the risk of

groundwater pollution due to the migration of leachate that may contain PHCs, phenols

71

and heavy metals (Bhattacharyya and Shekdar, 2003). For example, Hejazi et al. (2004)

reported that weathering (i.e. evaporation) and not biodegradation was the overall

dominant degradation mechanism occurring in a landfarm under hot and arid climate

conditions. Results from a 12-month field study showed that up to 76% of oil and grease

in oily sludge has been evaporated as a result of weathering, while relatively long chain

alkanes such as C17 and C1 s were biodegraded in the landfarm (Hejazi and Husain, 2004).

Another advanced land treatment approach is secure landfill after oily sludge dewatering.

This technology isolates sludge wastes from air and water through the use of thick layers

of impermeable clay and synthetic liner, and it also employs a leachate collection system

above the bottom liner (Moses et al. , 2003 ; Butt et al., 2008). Secure landfill is popular in

developed countries such as the USA, UK, Canada, and Germany, and it can greatly

address the environmental problems associated with landfarming, but its treatment cost is

much higher (Bhattacharyya and Shekdar, 2003).

2.5.4.2 Biopile/composting

Biopile/composting of petroleum wastes has received increased attention as a

substitute technology for landfarming which often requires a large land area. Biopile refers

to the turning of waste materials into piles or windrows usually to a height of 2-4 m for

degradation by indigenous or extraneous micro-organisms. The piles may be static with

installed aeration piping, or turned and mixed by special devices. The bio-treatment

efficiency can be improved with moisture adjustment, air blowing, and the addition of

bulking agent and nutrients. Bulking agents usually include straw, saw dust, bark and wood

chips, or some other organic materials. Addition of bulking agents results in increased

porosity in soil-sludge piles, which leads to better air and moisture distribution in the

matrix. This technology is termed as composting if organic material is added (Marin et al. ,

72

2006). The biodegradation rate can be enhanced by manipulating a number of operating

parameters such as controlling carbon : nitrogen : phosphrous (C:N:P) ratio, air blowing

or tilling to improve aeration, and moisture and temperature maintenance to keep high

microbial activity (Butt et al., 2008). A number of studies have been reported to use

biopile/composting for refinery oily sludge treatment. Wang et al. (2012) found that during

the composting process of aged oily sludge, microbial metabolic activity and diversity

were significantly enhanced by the addition of bulking agent cotton stalk, with a TPH

removal rate of 49.62% being observed in the middle layer of biopile after 220 days, but

the application of large amounts of nutrient (i.e. urea) had a suppressing effect on the

microbes. Liu et al. (2010) found that the addition of manure (as nutrients) to oily sludge

significantly increased the microbial activity and diversity, and the TPH in the treated

sludge was decreased by 58.2% after about 1 year ofbioremediation, but this number was

only 15.6% in the control plot.

Ouyang et al. (2005) investigated the effect ofbioaugmentation on the composting of

oily sludge, and they found that the TPH content decreased by 46-53% in the piles after 56

days of treatment, but only decreased by 31 % in the control piles. Kriipsalu et al. (2007)

reported the aerobic biodegradation of oil refinery sludge in composting piles with four

different amendments, and their results showed that after 373 days of treatment, the

reduction of TPH was 62, 51 , 7 4 and 49% in the piles with amendments of sand, matured

oil compost, kitchen waste compost, and shredded waste wood, respectively. As compared

to landfarming, biopile/composting is able to more efficiently remove PHCs in oily sludge

and could treat more toxic compounds since it creates controlled conditions more favoured

by biodegradation. Another noticeable characteristics of biopile/composting is that

temperature in the piles could increase up to 70 °C or more due to the heat generated by

73

intense microbial activity, and the application of this method for PH Cs degradation under

extreme climatic conditions such as sub Antarctica area has been proven successful

(Ouyang et al., 2005; Delille et al., 2007) . In addition, it is more environmentally friendly

since the composting can be conducted in treatment vessels and the VOCs emissions can

be controlled by auxiliary collection units. It is also easy to design and implement, and can

be engineered to fit different site conditions. However, the treatment capacity of

biopile/composting is much smaller than that of land treatment and it still requires a

relatively large area of land and long treatment duration for oily sludge degradation (Khan

et al., 2004) .

2.5.4.3 Bio-slurry treatment

Slurry-phase bio-treatment was reported to have faster pollutant removal than solid-

phase treatment (e.g., land treatment and composting), and has been successfully applied

to the cleanup of oil contaminated soils (Frank and Castaldi, 2003). This technology mixes

sludge-associated solids with water (i.e. 5-50% w/v) and dissolves contaminants into

aqueous phase to obtain a larger amount of solubilized pollutants. The microbial

degradation can then transform the pollutants to less toxic intermediates (e.g. , organic acids

and aldehydes) or end products of carbon dioxide and water. Slurry-phase biodegradation

usually occurs in designed slurry bioreactors where the contact between microorganisms,

PHCs, nutrients, and oxygen can be maximized (Weber and Kim, 2005). A variety of

bioreactor designs are available, such as the rotating drum equipped with lifters to provide

internal mixing, and the vertical tank equipped with an impeller for mixing (Woo and Park,

1999). Bio-slurry treatment has been successfully applied to oily sludge decontamination

(Christodoulatos and Koutsospyros, 1998). Ayotamuno et al. (2007) applied a bio-slurry

remediation approach to treat oily sludge through the addition of extraneous microbes as

74

well as regular mixing and watering, and they found that the TPH reduction in sludge was

40.7-53.2% and 63.7- 84.5% within two weeks and six weeks of treatment, respectively.

Machin-Ramirez et al. (2008) reported that the addition of commercial fertilizer to the

sludge slurry could greatly enhance the degradation of weathered oily sludge, with a TPH

removal of 24% occurring within 25 days of treatment. Ward et al. (2003) investigated the

biodegradation of oily sludge in slurry phase with sludge concentration in the range of

1.55-12.8%, and they found that the TPH degradation was in the range of 80-99% within

10-12 days by using three different bio-surfactant-producing microbial consortiums.

Large-scale application of bio-slurry remediation on oily sludge also showed

promising results. For example, Maga et al. (2003) reported that a 10,000-gallon

sequencing batch reactor (SBR) was used for the on-site biodegradation of oily sludge, and

the micro-organisms degraded the PHCs in sludge from 20,000 ppm to less than 100 ppm

within two weeks of treatment. According to Ward and Singh (2003 ), a large-scale bio-

slurry reactor system with a 4.55 x 106 L capacity was designed to treat oily sludge at the

Gulf Coast refinery, while air sparging and mechanical agitation system were incorporated

to improve the homogenization of oily sludge slurry, with 50% of oil and grease removal

being achieved after 80 to 90 days of treatment. Bio-slurry degradation is a rapid and

effective approach for oily sludge disposal which can greatly decontaminate oily sludge

within a short treatment period. Unlike other biodegradation treatments, bio-slurry

processing only requires a small area of land. A major concern with the application of this

technology to field-scale oily sludge treatment is the relatively high treatment cost. The

oily sludge slurry is non-homogeneous and clayey mixtures which can cause operation

problems, and thus needs pretreatment. During the bio-slurry treatment process, volatile

gaseous compounds can be generated and thus may require treatment. After bio-slurry

75

treatment, the mixtures also need dewatering. All of these pre-treatments and/or post-

treatments could significantly increase the overall cost. It was estimated that the operating

cost of the bio-slurry treatment of refinery oily wastes was above $625 per ton, while the

operating cost of landfarming was around $155 per ton (Frank and Castaldi, 2003).

2.5.4.4 Factors affecting oily sludge biodegradation

The oily sludge biodegradation can be affected by a variety of factors, such as the type

of microorganisms, treatment duration, temperature, nutrients, concentration and

characteristics of oily sludge. Many microorganisms, mainly bacteria and fungi, are

capable of degrading PH Cs, but there is no single microbial strain which has the metabolic

capacity to degrade all of the components found in oily sludge (Bassam and Mohammed,

2005). The degradation of PH Cs in sludge may involve progressive or sequential reactions

where a group of micro-organisms may initially degrade the petroleum constituents into

intermediate compounds, and these intermediates are then utilized by a different group of

microorganisms for further degradation (Janbandhu and Fulekar, 2011). As a result, the

biodegradation of oily sludge typically needs a microbial consortium with a succession of

species. It was reported that the employment of a mixed bacterial culture was more

advantageous in comparison with pure cultures due to synergistic interactions among

microbial species (Mukred et al., 2008). In addition, the biodegradation can be affected by

treatment duration. Generally, the degradation rate of PH Cs decreases with time, reaching

an apparent plateau associated with pollutant residues which are recalcitrant compounds

and have very slow degradation. The PHCs characteristics can also affect the

biodegradation efficiency. It was reported that the degradation was higher for

saturates/alkanes, followed by light aromatics, high aromatics and polar compounds, and

asphaltenes (Yerushalmi et al., 2003; Vasudevan and Rajaram, 2001). In addition, the

76

initial PHCs concentration can affect the bioremediation performance. For example, Lazar

et al. (1999) observed that the best biodegradation occurred for the bio-treatment that

received 10% of oily sludge (equivalent to 3.2% initial TPH). However, another study

reported that the addition of oily sludge to soil (with equivalent TPH concentration of

9,000-61,000 mg/kg) resulted in a TPH reduction of 70-90% during two months of

treatment regardless of the initial PH Cs concentrations (Admon et al., 2001).

Nutrient is an important factor affecting the degradation of PH Cs, and thus nutrient

supplementation is the foremost strategy employed for biodegradation stimulation. Admon

et al. (2001) and Y erushalrni et al. (2003) found that the removal of PH Cs was observed

only after nutrients were amended to oily sludge contaminated soil at a C:N:P ratio of

50: 10: 1. Roldan-Carrillo et al. (2012) investigated the biodegradation of oily sludge under

different nutrient conditions, and they found that after 30 days of treatment, the highest

TPH removal was 51% in the sludge which had a C:N:P ratio of 100:1.74:0.5. However,

Tahhan et al. (2011) observed that adding nutrients caused the inhibition of oily sludge

biodegradation probably because of the high nutrient concentration already present in the

original sludge, and such inhibitory effect increased with the addition of nitrogen and

phosphorous. In addition to the nutrient effect, the degradation of PHCs are usually

restricted by their high hydrophobicity or low solubility (Anna et al. , 2007). One effective

way to improve this is the use of surfactants to enhance the desorption and solubilization

of PHCs, thereby enhancing their bioavailability and facilitating their assimilation by

microorganisms (Kuyukina et al., 2005; Zhang et al., 2010). For example, Shuchi et al.

(2006) used Bacillus sp. SV9 to produce bio-surfactant, and approximately 59% of TPH

was degraded within 5 days. Vanessa et al. (2011) isolated five bacteria from

petrochemical oily sludge with two of them being capable of producing bio-surfactants,

77

and the mixed bacterial consortium degraded 90.7% of the aliphatic fraction and 51.8% of

the aromatic fraction in oily sludge within 40 days of treatment. Rahman et al. (2003)

investigated the combined effect of rhamnolipid bio-surfactant and nutrients (phosphorus

and potassium solution) on the biodegradation of refinery oily sludge, and they found that

a maximum degradation was achieved after 56 days of treatment, with the degradation rate

of n-alkanes being 100%, 83-98%, 80-85% and 57-73% for nCs-nC11, nC12-nC21, nC22-

nC31 and nC32-nC40, respectively. Cameotra and Singh (2008) also investigated the effect

of two additives (a nutrient mixture and a crude bio-surfactant) on oily sludge

biodegradation process, and they found that the individual use of any one additive along

with a microbial consortium brought about a TPH removal of 91-95% within 4 weeks of

treatment, with the crude bio-surfactant being more effective, but more than 98% of TPH

removal was obtained when both additives were used with the consortium. As a result, bio-

surfactants and inorganic nutrients can significantly enhance the biodegradation process,

and will have great potential to be used in the application of oily sludge bioremediation

(Haritash and Kaushik, 2009).

2.6 Discussion and conclusion

Oily sludge generation is an inevitable problem during the operation of petroleum

industry. Due to its toxicity and adverse environmental effect, oily sludge needs effective

treatment. A variety of oil recovery and sludge disposal technologies have been developed,

and some of them have been applied to field-scale treatment. Tables 2.1 and 2.2

summarizes the treatment methods introduced in this paper based on their development

status, performance, treatment duration, costs, advantages and limitations. In general, a

particular technology cannot satisfy all of the reuse and disposal requirements for different

78

oily sludge wastes. Some treatments may be very promising in terms of fuel recovery

and/or the decontamination of unrecoverable residues, but their capital and/or operating

costs could be very high, or their implementation to large-scale treatment might be

infeasible. Other treatments such as land treatment and composting may have great

applicability and low operating costs for large-scale treatment, but their microbial

degradation process can be very time-consuming. The selection of oily sludge treatment

technologies should depend on sludge characteristics, treatment capacity, costs, disposal

regulatory requirements, and time constraints. Some methods such as centrifugation,

surfactant enhanced oil recovery, freeze/thaw, froth flotation, and bio-slurry treatment may

be more suitable for treating oily sludge with high moisture content. However, other

methods such as incineration, pyrolysis and stabilization/solidification require sludge

pretreatment to reduce its moisture content. Since the technology selection involves a

variety of criteria, it is difficult to evaluate the overall performance of the available

technologies. Some multi-criteria decision analysis approaches may help develop an

overall technology evaluation system and help practitioners to select the most suitable

treatment methods.

As shown in Tables 2.1 and 2.2, each oily sludge treatment method is associated with

some advantages and limitations. A more effective treatment performance might be

achieved by integrating different methods into a process train. For example, the ultrasonic

irradiation method can be combined with freeze/thaw treatment of oily sludge to obtain a

promising oil recovery performance (Zhang et al. , 2012). The Fenton ' s oxidation can be

combined with solidification/stabilization method to offer a safer way of oily sludge

disposal (Mater et al., 2006). The addition of bio-surfactants during froth flotation and/or

bioremediation of oily sludge can also enhance the overall efficiency (Ramaswamy et al. ,

79

2007; Rahman et al., 2003). Other combination of oily sludge treatment technologies such

as emulsion liquid membrane process and microwave irradiation, microwave and

freeze/thaw, and oxidation and pyrolysis can be found in literatures (Shie et al., 2004; Yang

et al., 2009; Rajakovic and Skala, 2006; Chan and Chen, 2002). Moreover, most studies

were either focusing on oil recovery from oily sludge or aiming at the maximum removal

of PH Cs from oily sludge. The oil recovery may significantly reduce the volume of sludge

wastes for further disposal, but the unrecoverable residues may contain more recalcitrant

and toxic components which could increase the difficulty of disposal. It is expected that

the oil recovery technologies and sludge disposal treatments introduced in this chapter

have the potential to be used in conjunction to reduce the overall adverse environmental

impacts, but their combinations have still rarely been reported. Thus, the economic and

environmental performance of combined oil recovery and sludge disposal is worthy of

further investigation.

80

Method Status'

Solvent extraction F

Centrifugation F

Surfactant EOR F

Freeze/thaw L

Pyrolysis F

Microwave irradiation F

Electrokinetic L

Ultrasonic irradiation L

Froth flotation L

Table 2.1 Summary and comparison of oil recovery methods (Hu et al., 2013)

Efficiency' Cost3

C-B B-A

D-C B-A

B B-A

C

C-B B-A

A

C

C-B

C

Duration• By-products

A VOCs, unrecoverable

sludge slurry

A

A

A

A

A

A

A

A

Wastewater,

unrecoverable sludge

slurry

Wastewater and

unrecoverable sludge

slurry

Wastewater and

unrecoverable slurry

voes and chars

voes, wastewater and

unrecoverable solids

Wastewater and sludge

slurry

Wastewater and

unrecoverable sol ids

Large amount of

wastewater

Advantages

Easy to apply, fast and efficient

Easy to apply, fast and efficient,

high throughput, no need of

chemical addition

Easy to apply, fast and efficient,

Limitations

Large amount of organic solvents is used, high cost and not

environmentally sound, unable to treat heavy metals

High capital and maintenance cost, large energy consumption, noise

problem, viscosity reducing pre-treatment requirement, unable to treat

heavy metals

High cost, chemical surfactants could be toxic, surfactants need to be

limited effect on treatment of heavy removed from recovered oil

metals

Easy to apply, short treatment

duration, suitable for cold regions

Lower efficiency, cost could be high due to high energy consumption

for freezing, unable to treat heavy metals

Fast and efficient, recovered oil can High capital , maintenance, and operating cost, high consumption of

be upgraded, large treatment

capacity

Very fast and efficient, no need of

chemical addition

Fast and efficient, no need of

chemical addition, limited effect on

treatment of heavy metals

Fast and efficient, no need of

chemical addition

Easy to apply, no intensive energy

requirement

energy, and not suitable for oily sludge with high moisture content

Special designed equipment, high capital and operating cost, high

consumption of energy, small treatment capacity, unable to treat

heavy metals

Low treatment capacity and not easy to apply

High equipment cost, small treatment capacity, unable to treat heavy

metals

Relatively low efficiency, large amount of water is used, not suitable

for treating oily sludge with high viscosity, unable to treat heavy

metals

F: Field scale, L: Laboratory scale. 2 Efficiency (oil recovery rate) A: >90%; B: 75-90%; C: 50-75%; D: <50%. 3 Cost (US$/m3)

A: >200; B: 100-200; C: 50-100; D: <50; -: unknown. 4 Treatment duration A:< 1-2 days; B: 1-6 months; C: 6-12 months; D: 1-2 years.

81

Table 2.2 Summary and comparison of oily sludge disposal methods (Hu et al., 2013)

Method Status' Efficiency2 Cost3 Duration4 By-products Advantages Limitations

Incineration F A A A Hazardous gas Rapid and complete removal of High cost of equipment and auxiliary fuels , gas emissions and

emissions and fly ash PHCs in oily sludge, heat value can fly ash need further treatment, pre-treatment of moisture

be reused removal is required, unable to treat heavy metals

Stabilization/solidifi L A B B Stabilized/solidified Fast and efficient of encapsulating Moisture content in oily sludge need to be reduced first, PHCs

cation products PHCs in stabilized/solidified cannot be completely removed, loss of recyclable energy,

products, low cost, able to stabilized/solidified products require proper management

encapsulate heavy metals in oily

sludge

Oxidation L A A A CO2, oxidation Rapid and complete removal of Large amount of chemical reagents is utilized, high cost and

intermediates PHCs in oily sludge, relatively not environmentally sound, loss of recyclable energy, high cost

insensitive to external disturbances if advanced oxidation process taken place, limited effect on

heavy metals

Land farming F B D-C D-C voes, contaminated Low cost and do not need much Very slow process, pollutants may build up on repeated

soils, hazardous maintenance, large treatment applications, VOCs emission problems, risk of underground

leachate capacity water pollution, occupy a very large area of land

Landfill F B C D-C Contaminated soils, Large treatment capacity, relatively Higher cost than landfarming, very slow process, occupy a very

relatively small low cost, VOCs emission is large area of land

amount ofVOCs collected and groundwater pollution

could be prevented

Biopile/ F B D-C C-B voes emissions if Relatively large treatment capacity, Higher cost and smaller treatment capacity than land treatment,

composting 110 voes collecting faster process than land treatment, still requiring a large area ofland

units exist less land area requirement than

landfarming, suitable for cold

regions and various terrain

82

Bioslurry F A C-8 C-8 Sludge slurry Fastest biodegradation approach,

great PHCs removal performance,

small land area requirement

High cost, small treatment capacity, need skilled operation,

maintenance and monitoring, slurry residues need proper

management

1 F: Field scale, L: Laboratory scale. 2 Efficiency (PHCs removal rate) A: >90%; B: 75-90%; C: 50-75%; D: <50%. 3 Cost (US$/m3)

A: >200; B: 100-200; C: 50-100; D: <50; -: unknown. 4 Treatment duration A: < l-2 days; B: l-6 months; C: 6-12 months; D: 1-2 years.

83

Chapter 3 Ultrasonic oil recovery and salt removal from refinery tank bottom sludge *

* The results of this work have been published as: Hu, G.J., Li, J.B., Ronald, R.W., and

Arocena, J., 2014. Ultrasonic oil recovery and salt removal from refinery tank bottom

sludge. Journal of Environmental Science and Health, Part A, 49, 1425-1435.

3 .1 Introduction

A significant volume of oily sludge is generated from the oil and gas industry during

various activities such as drilling, production, processing, and distribution. This kind of

sludge usually exists as a complex mixture of various petroleum hydrocarbons (PHCs),

water, metals, and solid particles (Hu et al., 2013). It has been recognized as a hazardous

waste in many countries because it contains a high concentration of toxic components,

including polycyclic aromatic hydrocarbons (PAHs) and heavy metals (Yan et al., 2012).

Any improper treatment of oily sludge poses serious threats to the environment and human

health, and thus requires effective management. In the past several years, various physical,

chemical, and biological methods have been proposed for oil recovery from oily sludge

through various methods or their combinations (Hu et al., 2013). These methods include

solvent extraction, centrifugation, surfactant enhanced oil recovery, freeze/thaw treatment,

pyrolysis, microwave heating, electro-kinetic treatment, froth flotation, and ultrasonic

irradiation (Hu et al., 2013; da Silva et al., 2012). Like many others, these methods are

associated with different advantages and limitations. Amongst them, ultrasonic irradiation

represents an emerging and promising technique due to its several inherent merits such as

high efficiency, short treatment duration, and chemical-free application (Bendicho et al.,

2012; Jin et al., 2012; Li et al., 2013). It has been reported to effectively remove adsorbed

84

materials from solid particles and decrease the stability of water/oil emulsions (Bendicho

et al., 2012; Nii et al., 2009; Pilli et al., 2011). The cavitation collapse due to ultrasonic

irradiation does not only affect the surface of solid particles but also penetrates into the

solid matrix, and could thus improve the separation of oil from solid particles and slurries

(Zhang et al., 2012; Abramov et al., 2009). Several studies to investigate the effect of

ultrasound on oil recovery from sludge have been conducted. For example, Xu et al. (2009)

found that oil constituents can be effectively stripped from the surface of solid particles in

oily sludge when using combined ultrasonic cavitation (28 kHz) and mechanical vibration

in an ultrasonic cleaning tank, and an overall oil recovery efficiency of 55.6% was

observed. Jin et al. (2012) combined ultrasonic irradiation and thermochemical cleaning

methods to treat oily sludge, and they found an oil recovery rate of up to 99.3%. Zhang et

al. (2012) reported an oil recovery rate of up to 80% from an oily sludge-water matrix by

using combined ultrasonic irradiation and freeze/thaw treatment process, but the effect of

temperature was not investigated.

Although ultrasonic irradiation shows great potential in oil recovery, a number of

issues still need to be addressed. An important one is that the recovered oil from sludge

still needs reprocessing for refinery feedstock; however, the salt content (e.g., sodium,

calcium, potassium, and magnesium chlorides) was either unknown or unreported in

previous oil recovery studies (Zhang et al., 2012). When compared to fresh crude oil,

recovered oil may contain more salts which can cause undesirable consequences in the

refining process (Ye et al., 2008). For example, high salt content in feedstock oil corrodes

equipment, obstructs oil pipes, and poisons catalysts in refineries (Abdul et al., 2006;

Mahdi et al., 2008). It is thus of great importance to have desalting treatment for the

recovered oil before its entry to a refinery. Ultrasonic irradiation could serve as an effective

85

salt removal method, as it has been reported to enhance the desalting process of crude oil.

For example, Ye et al. (2008) observed that the desalting efficiency of crude oil by

ultrasonic irradiation could reach 87.9%, and the salt content in ultrasound treated oil was

3 .85 mg/L which is suitable for refining. Gholam and Dariush (2013) also reported that the

salt removal from crude oil when using ultrasonic irradiation was 84% under optimal

conditions, and the treated crude oil could meet the salt requirement for refinery feedstock.

However, the desalting effect of ultrasonic irradiation on oily sludge has rarely been

reported. In fact, if ultrasonic irradiation treatment can reduce the salt content in recovered

oil, the cost for further desalting of such recovered oil as refining feedstock could then be

decreased.

Another issue with ultrasonic irradiation enhanced oil recovery is that it may generate

a large volume of wastewater which contains a high concentration of unrecoverable PH Cs

and undesirable salts (Zhang et al., 2012). Removal of such pollutants in wastewater could

increase the overall cost of oily sludge treatment. However, most previous studies have

focused on the improvement of oil recovery either by exploring the optimal ultrasonic

operating condition or by investigating the combined effect of ultrasonic irradiation with

other oil recovery methods (Jin et al., 2012; Xu et al., 2009). Little attention has been paid

to the wastewater generated from oily sludge treatment and the variation of pollutants in

wastewater under different treatment conditions. The objective of this chapter was then to

investigate the effects of ultrasonic irradiation on both oil recovery and salt removal from

refinery tank bottom sludge, as well as the concentrations of PH Cs and salt in wastewater

generated from sludge treatment. A number of factors including ultrasonic power,

treatment duration, sludge to water ratio, and sludge-water slurry temperature were studied

to examine their impacts on the treatment performance.

86

3 .2 Experimental materials and methods

3.2.1 Materials

Oily sludge was collected from a crude oil tank bottom in an oil refinery plant in

western Canada. The sludge was black and sticky (Figure 3.1), and was stored in a sealed

can at 4 cc after collection. Characteristics of the sludge sample are listed in Table 3.1. All

of the organic solvents and reagents ( e.g. , cyclohexane, dichloromethane and toluene) used

for sample extraction were of high performance liquid chromatography (HPLC) grade

(>99%), and were purchased from Sigma Aldrich, Canada. Silica gel (from Sigma)

activated at 105 cc for 12 h was used to clean up the extraction solution, and anhydrous

sodium sulfate (from Sigma) dried at 400 cc for 12 h was used to remove trace amounts

of water from the extractant. In this study, the PHCs fractions of recovered oil were

analyzed and compared with those from oily sludge, fresh crude oil, and diesel. The crude

oil and diesel samples were obtained from an oil refinery plant and a gas station in western

Canada, respectively.

Table 3.1 Characteristics of tank bottom oily sludge sample for ultrasonic treatment

Parameter

Total petroleum hydrocarbons (TPH)

Water content

Solids content

Salt content

87

Concentration

62.2% (mass percent)

22.1 % (mass percent)

15.7% (mass percent)

4580 mg/kg

Figure 3.1 Tank bottom sludge sample

3.2.2 Methods

Laboratory experiments were conducted to examine the efficiency of ultrasonic

irradiation on oil recovery and salt removal from oily sludge under different operating

conditions. Figure 3.2 illustrates the ultrasonic treatment system which utilized a 20 kHz

Misonix Sonicator 3000 generator. The effects of several experimental factors on oil

recovery and desalting performance were investigated, and these include ultrasonic power,

treatment duration, sludge-to-water (SW) ratio, and the initial temperature of sludge-water

slurry. The high viscosity of sludge was reduced by using different volumes of water and

by adjusting the temperature of sludge-water slurry. As a result, the sludge to water ratio

and the initial slurry temperature were considered to be two important experimental factors

in this study (Table 3.2). No chemicals were added to the treatment process because the

purpose of this study was to evaluate the sole effect of ultrasonic irradiation. For the

ultrasonic irradiation treatment, about 5 g of oily sludge was placed in a glass beaker and

88

mixed with a given volume of ultrapure water produced by a water purification system

(Milli-Q® Advantage AlO) according to the specified SW ratio, while the initial

temperature of slurry was controlled by using a hot plate (Thermo Scientific Cimarec ?x?

Ceramic Hotplate). When the pre-set initial temperature (e.g., 20, 40, 60 and 80 °C) was

reached, the hot plate was removed and ultrasonic irradiation was then applied. The 1.27

cm diameter titanium sonic probe was placed into the center of the SW slurry mixture for

ultrasonic irradiation. The changes in slurry temperature during treatment were recorded

using a noncontact infrared thermometer (OAK.TON® TempTestr IR). The experiments

were conducted using different combinations of experimental factor levels, and each

experiment was conducted in triplicate. In this study, the effect of direct heating on oily

sludge treatment was also investigated. The direct heating experiments were conducted by

heating the sludge-water slurry using a hot plate from an initial temperature of about 10 °C.

Once the slurry was heated to the pre-set temperature, the glass beaker was immediately

removed from the hotplate.

89

CD 0 .00 000 000 OQO

Ultrasonic generator

Converter

Sludge-water slurry

Figure 3.2 Schematic diagram of the ultrasonic irradiation set-up for oily sludge

treatment

Table 3.2 Experimental factors and values of ultrasonic irradiation treatment

Factors Values used in experiments

Ultrasonic power (Yv) 21 48 75

Initial temperature of slurry (0 C) 20 40 60 80

Treatment duration (min) 2 4 6 8 10

Sludge-to-water (SW) ratio (g/mL) 1: 1 1:2 1:4 1:6 1:8

3.2.3 Sample analysis

After ultrasonic or direct heating treatment, the SW slurry mixture was transferred

into a 50-mL centrifuge tube. The sample was centrifuged (Thermal Scientific Sorvall

Legend Xl) at 1000 rpm for 10-min to remove unrecoverable solids residues. The oil and

90

aqueous phases after centrifugation were then separated using a separatory funnel, and the

mass of the oil layer separated was measured as oil recovery. The concentrations of salt

and TPH in both the separated oil layer and aqueous phase (i.e., wastewater) were analyzed.

The salt content in oil was analyzed using the method given in American Society for

Testing and Materials (ASTM) D3230, while in the wastewater it was measured using an

electrical conductivity meter (VWR Symphony Conductivity Meter) (Majumdar et al.,

2002). To estimate the TPH concentration in the oil layer, approximately 1 g of recovered

oil sample was dissolved in 15 mL cyclohexane and then shaken (Talboy 3500 Orbital

Shaker) for mechanical extraction at 150 rpm for 1 h. After shaking, the extraction solution

was passed through a glass column packed with silica gel and anhydrous sodium sulfate

for cleanup. The cleaned extraction solution was collected and then evaporated using a

rotary evaporator (Yamato RE400) to remove the solvent and concentrate the PH Cs. The

concentration of TPH in the recovered oil was then calculated using the measured PH Cs

mass and oil sample mass. Similarly, the TPH concentration in the oily sludge sample was

also determined. In terms of measuring TPH in wastewater, the wastewater sample was

subjected to liquid-liquid extraction (3X) using cyclohexane (with a volume ratio of 8:3),

with the remaining procedures being similar to those used for measuring TPH in the oil

layer sample. The detailed descriptions of sample extraction and analysis can be found in

Zhang et al. (2012). The PHC recovery rate was defined as follows:

(1)

91

Where R0 is oil recovery rate (%), Co1 and Cs are TPH concentrations (mg/g) in the

recovered oil layer and original sludge, respectively, Mo, is the total mass of recovered oil

(g) after separation, and Ms is the mass of oily sludge (g) used for each experimental

treatment.

The proportions of different PH Cs fractions in recovered oil, oily sludge, fresh crude

oil, and diesel were analyzed using a Varian CP-3800 Gas Chromatograph with flame

ionization (GC-FID). The external standard method was used to calculate PHCs

concentration, while decane (CJO), hexadecane (C16), tetratriacontane (C34), and

pentacontane (Cso) were used as standard compounds to determine PHCs fractions (CCME,

2001). The PHCs fractions Fl, F2, F3 and F4 were defined as the group of petroleum

hydrocarbons from C6 to CJO, CJO to C16, C16 to C34, and C34 to Cso, respectively. About 1

g of recovered oil, oily sludge, fresh crude oil, and diesel were dissolved in 20 mL of

cyclohexane, and then the solution was passed through the silica gel column cleanup and

subsequent rotary evaporation as discussed above. After evaporation, PHCs in the flask

were completely dissolved in 20 mL of toluene, and 2 mL of each solution was collected

for GC-FID analysis. The ZB-capillary column (Phenomenex Torrance, CA) with 30 m x

0.25 mm inner diameter and 0.24-µm film thickness was used. The GC analysis conditions

were set up with an injection volume of 1 µL, injector and detector (FID) temperature at

320 °C, and carrier gas of helium at a constant flow rate of 1.5 mL/min. The splitless

injection mode was performed on the 1079 PTV injector and after 0.7 min the split mode

was activated at a split ratio of 10: 1. The capillary column temperature was initially held

at 50 °C for 1 min, then ramped at 15 °C/min to 110 °C and further increased at 10 °C/min

to 300 °C and then held for 11 min. The total running time for each sample analysis was

about 3 5 min.

92

3.3 Results and discussion

3.3.1 Temperature change during ultrasonic treatment

Figure 3.3 presents the temperature change in the SW slurry mixture (with an initial

slurry temperature of 25 °C) after ultrasonic irradiation treatment; as seen, this change can

vary significantly. A more dramatic increase was found for samples treated with high

ultrasonic irradiation power than for those treated with low ultrasonic power. For example,

when using an ultrasonic power of75 W, the slurry temperature increased from 25 to 77 °C

within 6 min of ultrasonic treatment before reaching a plateau. However, at an ultrasonic

power of21 W, the slurry temperature only increased to 61 °C after 6 min of treatment and

thereafter it remained stable. This is generally in agreement with results obtained from

previous studies on sewage sludge treatment by ultrasonic irradiation (Dewil et al. , 2006).

When ultrasonic irradiation is applied to the SW medium, micro-bubbles form due to the

different pressures caused by ultrasonic waves, and the violent collapse of these micro-

bubbles when they reach an unstable size would produce shock waves and rapid increase

in temperature (Pilli et al. , 2011). As it can be seen in Figure 3.4, three layers including oil

layer (top), water layer (middle), and solid residue (bottom) were separated after ultrasonic

irradiation treatment. The water layer was not clear, indicating it might contain unwanted

impurities such as fine solids and dispersed PH Cs, which requires proper treatment before

discharging into the environment.

93

90

80 -u 70 !.-4) ... .a 60 (I)

~ a.. so E CJ l- -0-21 W

40 -0-48 W

30 ~ 75W

20

10 0 1 2 3 4 5 6 7 8 9 10

Treatment duration (min)

Figure 3.3 Temperature change in oily sludge-water slurry mixture treated by ultrasonic

irradiation ( experimental condition: sludge to water ratio of 1 :4 and initial

slurry temperature of 25°C)

Oil layer

Wastewater layer

Bottom residues

Figure 3.4 Three layers separated from ultrasonic irradiation treatment on oily sludge

94

3.3.2 Effect of temperature and ultrasonic irradiation

Temperature has proven to be an important factor affecting the demulsification of

W/0 emulsions (Peng et al., 2012). Increased temperature in oily sludge reduces viscosity

and accelerates the settling of water droplets in oily sludge, thus promoting the separation

of water and oil (Nii et al., 2009). To investigate the effect of temperature on oily sludge

treatment, the oil recovery and salt removal performance of both direct heating and

ultrasonic irradiation were further examined. Figure 3.5a presents the variation of oil

recovery rate and salt content in the recovered oil from direct heating. As shown, direct

heating was able to recover part of the oil from oily sludge and also reduced the salt content

in the recovered oil. In addition, the oil recovery rate increased with the rise of bulk

temperature, and accordingly the salt content in the recovered oil decreased. When the bulk

temperature of oily sludge-water slurry mixture rose to 80 °C by direct heating, an oil

recovery rate of 32.9% and a salt content of 6.0 mg/L in recovered oil were obtained.

Figure 3.4b presents the results ofTPH and salt concentration variation in wastewater with

the temperature of oily sludge-water slurry due to direct heating. It is shown that a higher

slurry temperature led to a higher salt content but a lower TPH concentration in wastewater.

Using direct heating, the lowest TPH concentration and highest salt content are found to

be 533.3 mg/Land 159.8 mg/L under 80 °C, respectively.

95

45 a 12 :]' -C')

40 E 10 ::::-

~ 35 ·a 0 "'O .._ Q) 30 8 Q) - .... «l Q) .... > ~ 25 0

(.) Q) 6 ~ > 0 20 .£; (.) Q) -.... 15 4 C:

Q)

6 -:ll- Oil recovery rate -C: 10 0

-0-Salt content 2 (.) -5 «l

(/)

0 0 20 40 60 80

Temperature (°C) 700 b 250

::J' -- ::J" O> g 600 -O') ... 200 E 2 -(,;) 500 ..... s: Q) -2 Ctl If> 150 ~ ~ 400 -(/) .!: ro C 300 ~ ,g 100 .E ~ -c C

200 Q) (!) -<.;) -i:l-TPH concentration C: C 50 0 0 (.) <.;) 100 ..... I -0-Salt content Ctl Q. (/) I- 0 0

20 40 60 80 Temperature (°C)

Figure 3.5 Effect of direct heating on oily sludge treatment at different temperatures, (a)

oil recovery rate and salt content in recovered oil, (b) TPH and salt

concentration in wastewater ( error bar represents standard deviation)

The combination of direct heating and ultrasonic irradiation enhances oil recovery

and salt removal (Figure 3.6). For example, oil recovery increased from 19.2% to 32.9%

when the slurry temperature increased from 20 to 80 °C using only direct heating method

96

(Figure 3 .Sa). However, it increased from 33 .8% to 51.6% when the initial slurry

temperature increased from 20 to 80 °C in the treatment using an ultrasonic power of 21

W and a treatment duration of 6 min, and accordingly the salt content in recovered oil

decreased from 7. 7 mg/L to 5. 0 mg/L. This shows that ultrasonic irradiation improved the

oil recovery of direct heating by 14.5-21.6% under lower ultrasonic power (i.e., 21 W)

conditions. However, the increase in initial slurry temperature by direct heating showed

very little performance enhancement when using relatively higher ultrasonic power. For

example, the oil recovery rates were 57.9% and 59.6% for treatments with ultrasonic power

of 48 W and 75 W when the initial slurry temperature was 80 °C, which were only slightly

higher than those (i.e., 52.8% and 59.1 %) for treatments with an initial temperature of

20 °C (Figure 3.6a). The salt content in recovered oil also showed little change (Figure

3.6b ). When the initial slurry temperature was increased from 20 to 80 °C by direct heating,

the TPH concentration in wastewater decreased from 1616 to 1330 mg/L (Figure 3.6c),

and the salt content in wastewater (Figure 3.6d) increased from 471 to 534 mg/L at an

ultrasonic power of 21 W, but a very slight change of such concentrations was observed at

an ultrasonic power of 75 Wand 48 W. The result shows that it is unnecessary to raise the

initial slurry temperature when applying relatively higher ultrasonic power for oily sludge

treatment because high ultrasonic power could rapidly increase the bulk temperature of

slurry to a high level as shown in Figure 3.3 .

97

;i" 0 ..._, Q) ro ... >. ... Q) > 0 (.) Q) ... 6

:::::i' Cl s ... Q)

ro ~ Q)

cii co ~ C C 0 ro ... c Q) (.) C 0 (.)

I a.. 1--

80 a 10 -21W -(;- 48W --i:s- 75 W _J -<r- 21 W -("-- 48 W -t::- 75W Cl 70 s 9

fr ·5 8 60 ~ ·-:l'). "C

' ' .J.. Q) 7 <' ~ J:!- I ... Q)

50 ·:i. · > 0 6 T (.) / T ..... Q) ,~....____ 40 ... 5 . A

1 C ;.l.._ ____ .,-,

c ~/ :r! 30 Q) 4

c '=' ':r ._:,

0 3 (.) 20 ..... co 2 Cf)

10 1

0 0 20 40 60 80 20 40 60 80

Initial temperature (°C) Initial temperature (°C)

2000 C 700 -<>- 21 W - 0-48W - b:- 75 W --<>- 21 W --::r- 48 W --fr-75 W

:::::i' 600 :i;

~ 6 1600 Cl ~~ I E '-" 'I' ... 500 Q)

ro 1200 ~

~ Q) 400

~---C~' I 1i5 co ~ ~

800 1i C 300 ~ Q) c 200 0

400 (.) ..... co 100 Cf)

0 ··············---l.--··-····---1--------··-· 0 20 40 60 80 20 40 60 80

Initial temperature (°C) Initial temperature (°C)

Figure 3.6 Effect of ultrasonic irradiation on oily sludge treatment at different initial slurry

temperatures, (a) oil recovery rate, (b) salt content in recovered oil, (c) TPH

concentration in wastewater, and ( d) salt content in wastewater ( experimental

condition: treatment duration of 6 min and sludge to water ratio of 1 :4; error bar

represents standard deviation)

98

b

d

3.3.3 Effects of ultrasonic power and treatment duration

As discussed above, there is no need to raise the initial slurry temperature because

ultrasonic irradiation greatly increases the temperature of the SW slurry medium within a

short time period, and an initial temperature between 20 and 40 °C achieved satisfactory

oil recovery performance (Figure 3.6). As a result, the initial slurry temperature of 25 °C

was selected in this study to further investigate the effects of ultrasonic irradiation on oily

sludge treatment. The oil recovery performance was improved either by increasing

ultrasonic power or treatment duration (Figure 3.7a). Under the same treatment duration,

higher ultrasonic power illustrated better oil recovery performance. For example, after 2

min of ultrasonic treatment, the oil recovery rate from sludge was 27.0%, 35.7%, and 49.3%

at an ultrasonic power of 21, 48, and 75 W, respectively. As shown in Figure 3.8b, no

obvious separation between water and oil was observed from the sludge-water mixture,

indicating that the low ultrasonic power and short treatment duration were not effective for

oil and water separation. However, such separation was noticeable when higher ultrasonic

power and longer treatment duration were applied (Figure 3.8c). For example, under an

ultrasonic power of 48 W, oil recovery significantly increased from 35.7% to 61.2% as the

treatment duration was extended from 2 to 8 min, and then it approached an asymptote

which indicates that further extension of treatment duration is unnecessary.

99

80

70

,....., 60 -£.

~ >, .... Q) 50 > 0 (.) Q) 40 a:: 6 30

20

10

0

s,:l 3500 a, E L.. Q) 3000 iii s: Q)

1n 2500 ~ C:

C: 2000 0

"" ~ c Q) 0 1500 C: 0 0

I 1000 0.. I-

500

0

a 16 -<>- 21 W ----c".)--- 48 W -&-75W :::;- -<>- 21W -6-48W -rr-75W C) E 14 '-'

·o T "O y Q) 12 Q) >

~ 0 (.) 10 ~ " ~ ', .5 c 8 Q) c 0 6 (.) -

1a (fJ 4

2

0 2 4 6 8 10 2 4 6 8

Treatment duration (min) Treatment duration (min)

-"- 21 W ~ ,.:- 48W -tr- 75 W C 700 '(> 21W ......-:r- 48W -tr- 75W ......... ...J --0)

E 600 '-' Q) (V; ii! ~ 500 ,$ $ 7/ l -i> VI (I)

~ 400 y / .5 c IT'(

Q) ;....

300 c ¢ 0 (.)

i\j 200 (fJ

100

.................. J. ................ .i. ................... .l .... - ........... 1.--.... 1 ................... l ......... - ..... _J ____ I -- i 0 2 4 6 8 10 2 4 6 8

Treatment duration (min) Treatment duration (min)

Figure 3. 7 Effect of ultrasonic irradiation duration on oily sludge treatment, (a) oil recovery

rate, (b) salt content in recovered oil, ( c) TPH concentration in wastewater, and

( d) salt content in wastewater ( experimental condition: sludge to water ratio of

1 :4 and initial slurry temperature of 25°C; error bar represents standard

deviation)

100

b

10

d

1>

10

Figure 3.8 Mixtures of oily sludge and water, (a) before ultrasonic treatment, (b) after

treatment under ultrasonic power of 21 W and duration of 2 min, ( c) after

treatment under ultrasonic power of 75 W and duration of 6 min ( experimental

condition: sludge to wat er ratio of 1 :4 and initial slurry temperature of 25°C)

Under an ultrasonic power of 75 W, the oil recovery rate also increased with treatment

duration from 2 to 6 min (i.e., highest oil recovery of 62.2% at 6 min), but it slightly

decreased with treatment duration beyond 6 min. Jin et al. (2012) utilized an ultrasonic

cleaning tank for oily sludge treatment and they also found that the oil recovery rate

increased significantly in the first 15 min of treatment due to the high rates of initial

desorption. However, when the treatment duration was extended, they found that the re-

adsorption of removed oil onto the solid particle surfaces increased, and a further

enhancement of oil recovery could not be achieved when the rate of desorption was equal

to that ofre-adsorption (Jin et al. , 2012)

Oil recovery performance under the same temperature using ultrasonic irradiation was

more promising when compared with only direct heating. For example, when the slurry

was treated for 2 min using an ultrasonic power of 21 W, its temperature increased to 40 °C

101

(Figure 3.3), and its oil recovery rate was 27.0% (Figure 3.7a), which is slightly higher

than that (i.e. , 23.3%) of direct heating to 40 °C (Figure 3.5a). The difference in oil

recovery rate between direct heating and ultrasonic irradiation became more significant at

higher temperature. For example, the oil recovery rate at 80 °C by direct heating was 32.9%

(Figure 3.5a), but reached 62.2% after 6 min of ultrasonic treatment at a power of 75 W

(Figure 3.7a) under which the slurry temperature increased from 25 to 75 °C (Figure 3.3).

The promising oil recovery performance of ultrasonic irradiation may be attributed to its

cavitation effect. The sudden and violent collapse of a considerably large number of micro-

bubbles during ultrasonic irradiation could generate strong hydro-mechanical shear forces

in the vicinity area of solid particles, which would then break the aggregates of solid

matrices and release the adsorbed or trapped PHCs (Pilli et al. , 2011 ; Xu et al. , 2009; Flores

et al., 2007). Chu et al. (2001) found that direct heating alone was insufficient to

disintegrate the floe structure in wastewater activated sludge, but the combined ultrasonic

irradiation and heating produced satisfactory results in sludge floe disintegration. The

strong hydro-mechanical shear forces generated during ultrasonic irradiation can also

cause disturbances in the bulk liquid and introduce a drag effect on the droplets. Under this

effect, small droplets of the dispersed-phase (i.e., water in W/0 emulsion) move faster than

the larger droplets, and the faster movement of these small droplets increases their collision

frequency with other water droplets, leading to an increased chance of forming larger water

droplets which would then enhance the separation of water and oil (Nii et al. , 2009).

As shown in Figure 3.7b, the salt content in recovered oil decreased with increasing

treatment duration and ultrasonic power. The decrease was more significant for the

treatment with an ultrasonic power of21 W, where salt content decreased from 12.8 to 5.7

mg/L when the treatment duration increased from 2 to IO min. Under an ultrasonic power

102

of 75 W, the salt content in recovered oil decreased from 5.7 to 4.3 mg/L when the

treatment duration was increased from 2 to 10 min, and the lowest salt content of 4.2 mg/,

which meets the salt content norm of 5 mg/Lin feedstock oil for most refineries (Ye et al.,

2008), was observed at a treatment duration of 8 min. This is in agreement with the results

from other studies of crude oil desalting using ultrasound. For example, Gholam and

Dariush (2013) reported that an increase of ultrasonic power to 50 W resulted in a decrease

of salt content in crude oil at a constant irradiation duration, and the increase of treatment

duration until 6 min under a given ultrasonic power also showed a positive desalting effect

on crude oil. However, a further increase in ultrasonic power beyond 50 Wand treatment

duration beyond 6 min didn't show further desalting of crude oil.

TPH concentration in wastewater decreased with increasing ultrasonic power and

treatment duration (Figure 3.7c and 3.7d). As Figure 3.9 shows, all wastewater separated

from oil recovery treatment were turbid because the presence of emulsified PHCs and

solids. The wastewater from the treatment using lower ultrasonic power and shorter

treatment duration showed darker colour than that from high ultrasonic power and longer

treatment duration, indicating higher TPH concentration within it. However, the salt

content in wastewater varied in an opposite trend. The wastewater contained less TPH but

more salt content from oily sludge treatments with higher ultrasonic power and longer

treatment duration. Under an ultrasonic power of 21 W, the TPH concentration in

wastewater decreased significantly from 2600.0 to 1622.2 mg/L, and the salt content

increased from 279.0 to 448.8 mg/L when the treatment duration increased from 2 to 6 min.

As the treatment duration increased to 10 min, the TPH concentration slightly decreased

to 1444.4 mg/Land salt content slightly increased to 506.1 mg/L.

103

The high TPH concentration (i.e., 2600.0 mg/L) in wastewater under low ultrasonic

power and short treatment duration (i.e., 21 Wand 2 min) could be the result of insufficient

oil-water separation. Insufficient separation could also lead to a large portion of soluble

salt in the aqueous phase which was not separated from the oil-water matrix, and thus

resulted in lower salt content in wastewater. On the other hand, under an ultrasonic power

of 75 W, the TPH concentration in wastewater decreased significantly from 2133.3 to

1044.4 mg/L, and the salt content increased from 554.0 to 588.2 mg/L when the treatment

duration increased from 2 to 6 min, respectively. When the treatment duration increased to

10 min, the TPH concentration only slightly decreased to 911.2 mg/Land the salt content

to 608.1 mg/L.

Figure 3.9 Wastewater generated from oil recovery treatment using different ultrasonic

irradiation powers and durations

When the slurry temperature was the same, higher concentrations of TPH and salt

were found in the wastewater separated from the treatment by ultrasonic irradiation as

compared with that from direct heating. For example, when the slurry was treated at an

104

ultrasonic power of 75 W, the temperature rose to 77 °C in 8 min (Figure 3.3), and the

concentration of TPH and salt in wastewater were 955.4 mg/L (Figure 3.7c) and 609.9

mg/L, respectively (Figure 3.7d). In contrast, when the slurry temperature was increased

to 80 °C by direct heating, the TPH and salt content in the wastewater were 383.1 and

195.8 mg/L, respectively (Figure 3.5b ). In general, soluble salts in oily sludge were

dissolved in the water phase of W /0 emulsion, and this water phase was surrounded by a

continuous oil phase. Under higher ultrasonic power and longer treatment duration, a

higher water-oil separation efficiency (also meaning a higher oil recovery rate) was

obtained.

The higher oil recovery could then lead to a lower TPH concentration in wastewater.

Such higher demulsification efficiencies of oily sludge could also lead to less salt content

in the separated oil phase but higher salt content in the separated wastewater (Gholam and

Dariush, 2013). When the treatment duration was extended beyond 6 min, the further

enhancement effect of ultrasonic irradiation on oil recovery cannot be achieved as

indicated above, and thus the TPH concentration in wastewater didn't show further

decrease nor the salt content in wastewater showed further increase. Although the TPH

concentration in wastewater was significantly reduced under a relatively high ultrasonic

power and a longer irradiation time, it is still very high (i.e., 911.1 mg/Lat 75 Wand 10

min) as compared to that in common petroleum refinery wastewater effluents (Fakhru ' l-

Razi et al., 2009). The high salt content in refinery wastewater is also of environmental

concern (Zhao et al., 2011 ). Therefore, wastewater from oily sludge treatment by ultrasonic

irradiation could be a secondary pollution problem and must therefore be properly handled.

105

3 .3 .4 Effect of sludge to water ratio

Higher ultrasonic power was generally associated with a higher oil recovery rate and

a lower salt content in recovered oil (Figure 3.7). Under the same ultrasonic power level,

the oil recovery rate increased with decreasing sludge to water ratio, but further decrease

of sludge to water ratio beyond a certain value instead resulted in a decrease of oil recovery.

Too much water in the oily sludge-water mixture could have negative impact on the

demulsification process by ultrasonic irradiation. When the sludge to water ratio was below

1 :4, both oil recovery and desalting were reduced under different ultrasonic power levels.

For example, under an ultrasonic power of 75 W, oil recovery reached a maximum of 56.6%

(Figure 3 .1 Oa) and the salt content in recovered oil decreased to as low as 3 .2 mg/L (Figure

3 .1 Ob) at a sludge to water ratio of 1 :4. When the ultrasonic power was 48 W, the highest

oil recovery (i.e., 51.4%) occurred at sludge to water ratio of 1 :2, but no significant

difference of oil recovery was observed between oily sludge treatments at sludge to water

ratio of 1 :2 and 1 :4.

Too low water content (i.e. , too high sludge to water ratio) could result in increased

viscosity of the oily sludge-water slurry mixture which could then impede the formation

and collapse of cavitation micro-bubbles, leading to weakened sonication effect on oil

recovery and desalting (Pilli et al. , 2011). For example, Feng and Aldrich (2000)

investigated the PHCs contaminated soil remediation by ultrasonic irradiation, and they

found that the increased solid concentration to above 50% in the oily slurry could

significantly inhibit the ultrasonic cavitation process. Our previous study reported that the

highest oil recovery rate (i.e. 80%) occurred at an ultrasonic power of 66 Wand a sludge

to water ratio of 1 :2 when using a combined ultrasonic irradiation and freeze/thaw

treatment process (Zhang et al. , 2012). It is greater than that (i.e. 62.2%) observed in this

106

work ( e.g., 75 W, sludge-water ratio of 1 :4, when using only ultrasonic irradiation) (Figure

3.7a) because the freeze/thaw process could have a synergic effect on the ultrasonic

irradiation process (Zhang et al. , 2012).

107

~ e..... >, ... Q) > 0 (.) Q) er: 6

~ E .... .'!l "' 3: .,

1iS "' 3: -~ C 0 ~ .b C ., 0 C 0 0 I a. I-

70 -0- 21 W -C>- 48W -ls- 75 W a ::;- 16

r .A- 21W --<>- 48 W -ls- 75W Ol .s 14 60 ·5

-0 I Q) 12 so Q) r / { > 0

(.J 10 ~ 40 .!; c 8 Q)

30 c 0 6 (.J

20 iii C/J 4

10 2

0 ··'"······-···-···1 ...................................... 1 ....................................... 1 ··············-·-·······--·-·····J.. .... 0 ..................... - .......... .!

1:1 1:2 1:4 1:6 1:8 1:1 1:2 1:4 1:6 1:8

Sludge/water ratio Sludge/water ratio

3000 -<- 21 W --(-48W -&- 75W C ....... 1600 ---v- 21 W -:.:;;- 48 W -ir-' 75 W ...J Ol .s

2500 Q) 1400 1a 3: 1200 Q)

2000 Kl 3: 1000

.!; ?: 1500 800 Q)

?: 0 (.J

600 iii 1000 C/J

f 400 500

200

0 0 1:1 1:2 1:4 1 :6 1:8 1:1 1 :2 1:4 1:6

Sludge/water ratio Sludge/water Ratio

Figure 3.10 Effect of sludge to water ratio on the performance of oily sludge treatment by

ultrasonic irradiation, (a) oil recovery rate, (b) salt content in recovered oil, ( c)

TPH concentration in wastewater, and ( d) salt content in wastewater

( experimental condition: treatment duration of 6 min and initial slurry

temperature of 25°C; error bar represents standard deviation)

108

1:8

b

d

The TPH and salt concentration decreased with the increase in water addition which

significantly reduces pollutant concentrations in wastewater (Figure 3.10c and 3.10d).

Under an ultrasonic power of 75 W, the TPH concentration decreased from 2066.3 to 992.1

mg/L as the sludge to water ratio changed from 1: 1 to 1 :8, and accordingly the salt

concentration decreased from 1248.7 to 320.7 mg/L. Lower ultrasonic power (i.e., 21 W)

treatment was generally associated with a higher TPH and a lower salt concentration in

wastewater, indicating that a lower ultrasonic power was insufficient for effective oil and

water separation at any sludge to water ratio. Although increased water content in the

sludge-water slurry mixture system reduces the TPH and salt concentrations in wastewater,

less promising oil recovery and desalting performance (shown in Figure 3.10a and 3.10b)

could be a negative consequence. Moreover, a high water content in the slurry consumes

a large volume of fresh water, which makes the ultrasonic irradiation treatment not

environmentally sound. As a result, the sludge to water ratio of 1 :4 was selected as the

optimal value for oily sludge treatment.

3.3.5 Characteristics of recovered oil

The characteristics of recovered oil were compared with those of different samples

including original oily sludge, fresh crude oil, and diesel. As Figure 3 .11 a shows, the F2,

F3, and F4 fractions accounted for 24.9%, 66.1 %, and 9.0% of TPH in the original oily

sludge, 24.7%, 66.3%, and 9.0% ofTPH in the recovered oil, 54.3%, 39.8%, and 5.9% of

TPH in fresh crude oil, as well as 47.8%, 52.2%, and 0% of TPH in diesel, respectively.

The GC-FID profiles of oily sludge, recovered oil, fresh crude oil, and diesel was shown

in Appendix I. As compared to the original sludge, the recovered oil contained nearly the

109

same percentage ofF2, F3 and F4 fractions, indicating that ultrasonic irradiation under the

operating conditions in this study did not rupture the long-chain and heavy PHCs. Fresh

crude oil and diesel contained more F2 fraction (i.e., light PHCs) than recovered oil, but

more F3 and F4 fractions (i.e., heavy PHCs) were found in the latter. Figure 3.10b shows

that the recovered oil contained a higher percentage of TPH (i.e., 80.4%) than fresh crude

oil (i.e. , 71.7%).

As shown in Figure 3.1 lb, oily sludge contained the lowest TPH content (i.e. , 62.2%),

while the refined product (i.e., diesel) showed the highest TPH content (i.e., 87.7%). These

results revealed that ultrasonic irradiation could significantly increase TPH content in the

recovered oil, and the recovered oil contains more valuable PHCs content than crude oil.

The present study also showed that the recovered oil from the ultrasonic irradiation of oily

sludge contained low salt content (i.e. , as low as 3.5 mg/L). Consequently, the recovered

oil is of better quality as an energy source when compared with the original oily sludge

and crude oil, and it can be used as feedstock in conventional refmeries.

110

80 ~ Oily sludge • Recovered oil it Crude oil o Diesel

70

~ 60 0 -C 50 a :.::; :::s .c 40 ·;:: ...... (/)

0 30

0 F2 F3 F4

PHCs fractions

100 b 90

....-.. 80 '::?. 0 70 -Q) C) 60 ro c

50 Q) (.) L.. <l) 40 a. I 30 a.. I- 20

10

0 Oily sludge Recovered oil Crude oil Diesel

Samples

Figure 3.11 Distribution of PH Cs fractions and TPH in different hydrocarbon samples, (a)

PH Cs fraction percentage, (b) TPH percentage ( error bar represents standard

deviation)

111

3 .4 Conclusions

Ultrasonic irradiation was illustrated as a promising approach for oil recovery and

desalting of refinery tank bottom sludge. About 62.2% of PH Cs was recovered from the

sludge at an ultrasonic power of 75 W and a treatment duration of 6 min. The optimal

sludge-to-water ratio was found to be 1 :4, yet too high or too low water content in the

sludge slurry system led to lower oil recovery rate. It was found that the salt content in

recovered oil was decreased to less than 5 mg/L after 6 min of ultrasonic irradiation at a

power of 75 W, which can meet the salt content requirement for feedstock oil in most

refineries. The recovered oil contained about 80.4% TPH which is higher than in crude oil

(i.e., 71.7%). Its TPH contained similar percentages of F2, F3 and F4 fractions as the

original sludge, which implies that ultrasonic irradiation did not break the heavy PHCs. It

was found that the separated wastewater from ultrasonic irradiation contained relatively

high concentrations of TPH and salt which require proper treatment, although a higher

ultrasonic power led to a higher salt concentration and a lower TPH concentration. In

summary, ultrasonic irradiation is shown to be an effective method for refinery oily sludge

treatment for the purpose of recovering high-quality oil with more PH Cs and less salt.

112

Chapter 4 Oil recovery from refinery oily sludge through ultrasonic assisted extraction *

*The results of this work have been accepted for publication in the Journal of

Environmental Science and Health, Part A. in 2016.

4.1 Introduction

Among various oil recovery treatment methods, solvent extraction as a simple process

of separating PHCs from various matrixes has been proven effective (Zubaidy and

Abouelnasr, 2010; Taiwo and Otolorin, 2009). In this process, oily waste and solvent are

mixed to ensure adequate miscibility of oil in solvent, followed by the separation of oil

from solvent through distillation of the oil/solvent mixture. Most water and solids are

rejected as impurities which can be removed by gravitational settling or centrifugation (Al-

Zahrani and Putra, 2013) . Several efforts have been made to investigate the oil recovery

from oily wastes by solvent extraction in recent years. For example, Avila-Chavez et al.

(2007) employed supercritical ethane and dichloromethane to treat oily sludge and found

that more than 50% of oil can be recovered; Taiwo and Otolorin (2009) reported an oil

recovery up to 67.5% when using hexane and xylene extraction on tank bottom sludge;

Zubaidy and Abouelnasr (2010) found that methyl ethyl ketone (MEK) and light petroleum

gas condensate (LPGC) can recover 39% and 32% of oil from oily sludge when the

solvent-to-sludge ratio was 4: 1.

Traditional solvent extraction process is usually associated with a number of

drawbacks, such as long extraction time, low extraction efficiency, large solvent volume,

and high cost. Some alternative approaches, including pressurized liquid extraction (PLE),

microwave assisted extraction (MAE), supercritical fluid extraction (SFE), and ultrasonic

113

assisted extraction (UAE), have been used to enhance the utility of solvent extraction

(Bossio et al., 2008). These enhanced methods were regarded as environmentally friendly

processes with high extraction efficiency, improved selectivity, and rapidity. However,

they still face some particular challenges. For example, PLE, MAE and SFE require

complex instrumentation and large amount of energy consumption ( e.g., external heat and

pressure) (Gorn;alves and Alpenurada, 2005). In comparison, UAE offers several

advantages including relatively lower equipment cost, simplicity of operation, lower

extraction temperature, and the ability to process large sample volume (Banjoo and Nelson,

2005).

Ultrasonic treatment could produce cavitation bubbles which generate preferable

internal temperature and pressure, thus causing the extraction solvent to propagate outward

and collide with matrix particles (Bossio et al. , 2008). As a result, the matrix could be

broken up to expose more surface area for solvent extraction (Capelo and Mota, 2005).

Such unique extraction characteristics make UAE become an effective approach for

extracting organic contaminants from solid matrices, showing great potential in the

remediation of contaminated soils. For example, Gonvalves and Alpendurada.(2005)

reported a high extraction efficiency ( 69-118 %, average 8 8 % ) when using U AE process to

extract various pesticides from horticulture soils. Aydin et al. (2007) investigated the

effect of UAE on the removal of persistent organic pollutants including P AHs,

polychlorinated biphenyls (PCBs), and organochlorine pesticides (OCPs) from airborne

particles, and found that more than 90% of these pollutants can be removed with a shorter

extraction time and a lower solvent volume compared with traditional Soxhlet extraction.

Bossio et al. (2008) applied UAE to extract organic anthropogenic waste indicators (AWis)

from soil and sediment samples, and found that the method recoveries ranged from 46.1 %

114

to 110.0% in the fortified soil and from 49.2% to 118.6% in the fortified sediment samples.

However, limited studies have been reported on using UAE for the rapid extraction of oil

from refinery oily wastes.

The objective ofthis work was therefore to evaluate the effect ofUAE process on oil

recovery from refinery oily sludge by comparing to conventional mechanical shaking

extraction (MSE). Two types of UAE system, including ultrasonic assisted extraction

probe (UAEP) system and ultrasonic assisted extraction bath (UAEB) system, were

investigated. Several experimental factors including solvent type, solvent-to-sludge ratio,

ultrasonic treatment duration, ultrasonic power, ultrasonic bath temperature, and the

number of extraction cycles, were examined for their effects on the oil recovery

performance. Their impacts on solvent recovery were also investigated. The results would

be valuable for the development of improved oily waste recycling approaches.

4.2 Experimental materials and methods

4.2.1 Materials and reagents

Tank bottom sludge from an oil refinery in western Canada was used as the study

sample. After collection, the sludge was stored in a sealed glass jar at 25 °C, and was well

stirred manually before use in the experiments. Table 4.1 lists the properties of oily sludge.

The TPH concentration was measured using EPA method 8015C (2007), and the water

content was analyzed based on ASTM D297 4-00 Method A (2013 ), while the solid content

was calculated according to the TPH and water contents. Asphaltene content was analyzed

using ASTM D3279 method (2007), and metal elements were measured using Inductively

Coupled Plasma (ICP) analysis based on ASTM D5 l 85 (2009).

115

Table 4.1 Properties of API separator oily sludge for extraction treatment

Property Value

Total petroleum hydrocarbon (TPH) (mass percentage) 35.5%

Water content (mass percentage)

Solid content (mass percentage)

Asphaltene content (mass percentage)

Arsenic (As) (mg/kg)

Chromium (Cr) (mg/kg)

Cobalt (Co) (mg/kg)

Copper (Cu) (mg/kg)

Lead (Pb) ( mg/kg)

Nickel (Ni) (mg/kg)

Vanadium (V) (mg/kg)

Zinc (Zn) (mg/kg)

36.0%

28.5%

12.2%

24.5

11.0

1.2

24.2

32.2

7.4

2.1

44.4

Based on the consideration of toxicity, oil miscibility, economics, viscosity, and

vapor pressure (Hamad et al. 2005; Kheireddine et al. 2013), three organic solvents were

selected in this study for the treatment of oily sludge, including cyclohexane (CHX), ethyl

acetate (EA), and methyl ethyl ketone (MEK). These solvents with HPLC grade (>99%)

were purchased from Sigma Aldrich, Canada. Silica gel (70-230 mesh, supplied by Sigma

Aldrich, Canada) activated at 105 °C for 12 h was used to clean up extraction solution, and

anhydrous sodium sulfate (Sigma Aldrich, Canada) dried at 400 °C for 12 h was used for

the removal of moisture in extraction solution.

116

4.2.2 Methods

4.2.2.1 Optimization of three treatment processes

The optimal operational conditions of three extraction processes, including

conventional mechanical shaking extraction (MSE), ultrasonic assisted extraction probe

(UAEP) system, and ultrasonic assisted extraction bath (UAEB) system, were examined

using multifactor orthogonal design followed by single factor experiments (Li et al. , 2012).

The impact factors and their experimental levels were presented in Table 4.2. For each

extraction process, the effect of four impact factors were investigated at three levels using

a L9 (34) orthogonal array (Table 4.3). The oil recovery rate and solvent recovery rate were

selected as the response variables. Three statistical coefficients (i.e., K, R, and F-ratio)

were used for the evaluation of experimental results and the identification of optimal

operational condition. Coefficient K is the sum of percentage of oil or solvent recovery for

each impact factor at each level, respectively. The higher the K value, the higher oil or

solvent recovery the factor contributes. Coefficient R is the extreme difference which can

represent the fluctuation degree of oil or solvent recovery in accordance with the variation

of impact factors. A larger R value means more influence of the corresponding factor on

the results. F-ratio was used to determine the statistical significance level of impact factors

by ANOVA analysis (Gonder et al. 2010). Single factor experiments were carried out

(Table 4.4) to validate the results obtained from orthogonal experiments.

Table 4.2 Impact factors and levels of each factor in three extraction processes for the

orthogonal experiments

Extraction Levels Impact factors Symbol

Processes 1 2

117

3

MSE Solvent type A CHX EA MEK

Solvent-to-sludge ratio (v/m) B 2: 1 3:1 4:1

Extraction duration (min) C 10 30 60

Shaking speed (rpm) D 100 175 250

UAEP Solvent type A CHX EA MEK

Solvent-to-sludge ratio (v/m) B 2:1 3: 1 4:1

Treatment duration (s) C 10 15 20

Ultrasonic power (W) D 21 48 66

UAEB Solvent type A CHX EA MEK

Solvent-to-sludge ratio (v/m) B 2: l 3:1 4:1

Treatment duration (min) C 5 10 15

Ultrasonic bath temperature (0 C) D 20 40 60

Table 4.3 A L9 (34) orthogonal array of factors and levels for each extraction process

Impact factors & levels Runs

A B C D

1 1 1 1 1

2 1 2 2 2

3 1 3 3 3

4 2 1 2 3

5 2 2 3 1

6 2 3 1 2

118

Extraction

process

MSE

UAEP

7

8

9

3

3

3

1

2

3

3

2

2

3

1

Table 4.4 Single factor experimental conditions of the three extraction processes

Runs Experimental factors

#

Solvent type S/S ratio Extraction duration Shaking speed (rpm)

(v/m) (min) for MSE, or for MSE, or ultrasonic

treatment duration ( s) for power (W) for UAEP,

UAEP, or treatment or bath temperature

duration (min) for (°C) for UAEB

UAEB

1 CHX, EA, MEK 4:1 60 250

2 CHX 2:1, 4:1 , 8:1 60 250

3 CHX 4:1 10, 60, 120 250

4 CHX 4:1 60 150, 250, 300

Runs Solvent type S/S ratio Treatment duration ( s) Ultrasonic power (W)

(v/m)

1 CHX, EA, MEK 4:1 20 21

2 CHX 2:1 , 4:1 , 8:1 20 21

3 CHX 4:1 10,20, 30 21

4 CHX 4:1 20 21,66, 75

119

UAEB Runs Solvent type S/S ratio Treatment duration (min) Bath temperature (0 C)

(v/m)

1 CHX,EA,MEK 4:1 15 20

2 CHX 2:1, 4:1, 8:1 15 20

3 CHX 4:1 5, 15, 30 20

4 CHX 4:1 15 20,60,80

4.2.2.2 Extraction treatment process

Oily sludge sample and solvent were mixed at designated solvent-to-sludge (S/S)

ratio in a centrifugation tube, and then this mixture was subject to different extraction

processes. In terms ofMSE process, the tube was put on a mechanical shaker (Talboy 3500

Orbital Shaker, 75 W) for shaking extraction (Schwab et al. 1999). In UAEP process, a 20

kHz Misonix Sonicator 3000 generator was utilized to generate ultrasonic power, and the

ultrasonic probe was placed into the center of the sludge-solvent mixture. The MSE and

UAEP were operated at room temperature (i.e., 25 ± 2°C). For UAEB treatment, a 25 kHz

ultrasonic tank (Branson IC 1216) was used. Centrifugation tubes containing sludge-

solvent mixture were put in the pre-heated water bath in the ultrasonic tank and then subject

to ultrasonic treatment. All sludge-solvent mixture after these three extraction processes

was allowed to settle for 24 h at room temperature, and three layers were observed after

settling, including an extractant layer (i.e., mainly oil and solvent mixture) on the top, a

water layer in the middle, and the semi-solid residues at the bottom. The top extractant

layer was collected for distillation in a vacuum rotary evaporator (Yamato RE400) at a

temperature of 40 °C. The solvent collected from the vacuum evaporation was considered

as the solvent recovery, while the remaining part in the evaporator flask was considered as

120

the recovered oil. The solvent recovery rate was determined as the ratio of the mass of

recovered solvent to that of original solvent used in the extraction experiment. For the

simplicity of comparison, the oil recovery rate in this study was calculated as the ratio of

the recovered oil mass to the original sludge mass as reported in previous literatures

(Zubaidy and Abouelnasr, 2010; Taiwo and Otolorin, 2009). All the solvent extraction

experiments were carried out in triplicates.

Figure 4.1 Experimental setups of the three different extraction processes

4.2.2.3 Extraction cycles

In this study, the impact of extraction cycles on recovered oil yield was investigated

(Al-Zahrani and Putra, 2013 The obtained semi-solid residue from the first extraction was

subject to the second extraction immediately following the extraction procedure as

described above, while such extraction was repeated for another two times. All of the

extractions were operated using three solvents at the optimal conditions which were

121

identified from section 2.2.1. The oil recovery rate of the ith ( ORi) extraction was calculated

based on the mass of recovered oil from the ith extraction (Xi):

ORi (%) = L~Xi x 100 s

(4.1)

Where Ms is the mass of original oily sludge samples.

4.2.2.4 Sample analysis

The recovered oil from different treatments was purified to remove water, particles,

and unwanted polar organic compounds by column cleanup (CCME, 2001). About 1.5 g

of recovered oil was completely dissolved in 20 mL of solvent (i.e., 1: 1

cyclohexane/dichloromethane) and passed through the column packed with silica gel and

anhydrous sodium sulfate. After column cleanup, the oil sample was sent for vacuum

evaporation. After evaporation, the final leftover was weighed as relatively pure PHCs

(Zhang et al. 2012), and the TPH content was calculated as the mass percentage of PH Cs

in recovered oil. The distribution of PH Cs fractions (F2, F3 and F4) in the recovered oil

by different extraction processes was also analyzed. The fractions of F2, F3 and F4 were

defined as the group of petroleum hydrocarbons from C10 to C16, C16 to C34, and C34 to Cso,

respectively (CCME, 2011). The analysis was conducted using a Varian CP-3800 Gas

Chromatograph with flame ionization detector (GC-FID) following the CCME standard

method (CCME 2001). The physical properties and heavy metals in recovered oil were

analyzed according to ASTM methods as described in section 2.1.

122

4.3 Results and discussion

4. 3 .1 The effect of mechanical shaking extraction (MSE)

Figure 4.2a presents the effects of four impact factors on oil recovery from MSE

treatment. It can be seen that all three solvents were effective on oil recovery from oily

sludge. CHX can recover more than 50% of oil from oily sludge, followed by EA (i.e.,

37.9%) and MEK (i.e., 33 .3%). The amount of recovered oil by CHX was much higher

than the TPH content in original oily sludge (Table 4.1). This is because the recovered oil

by CHX contained a considerable amount of impurities such as water and solid particles,

and thus the calculated oil recovery of CHX extraction was higher than the those of EA

and MEK extractions. It was reported that CHX can be an effective solvent for oily sludge

treatment since it has high affinity with various petroleum constituents in oily sludge

compared with other solvents (Taiwo and Otolorin 2009; Liang et al. 2014). Other factors

such (SIS ratio, extraction duration, and shaking speed) also affect oily recovery.

Particularly, solvent type illustrated the most significant influence on oil recovery,

followed by shaking speed, extraction duration, and SIS ratio according to the order of Rj

value (Table 4.5). The highest oil recovery was observed as 63. 7% in CHX extraction with

the SIS ratio of 4: 1, extraction duration of 60 min, and shaking speed of 250 rpm,

respectively. According to the ANOV A analysis shown in Table 4.6, the effects of four

impact factors on oil recovery were significant. It is widely reported that high SIS ratio

could be beneficial to solvent extraction treatment of tank bottom sludge (Zubaidy and

Abouelnasr, 2010) and used lubricating oil (Al-Zahrani and Putra, 2013; Rincon et al.

2005). This is because the increase in the ratio of SIS can enhance the mutual solubility of

oil in solvent, thereby increasing the amount of oil recovered (Al-Zahrani and Putra, 2013).

Longer extraction time is also preferred since it allows sufficient contact time for oil and

123

Runs

solvent. In MSE treatment, the oil recovery reached equilibrium after 60 min. Higher

shaking speed could generate larger turbulence in extraction system, which can positively

affect the mixture of oil and solvent, and thereby enhance the oil recovery performance.

60

'i: I; 40 (I)

I :> 0 . u I~ 20

.. . a ,

CHX EA MEK Solvent type

2:1 3:1 4:1 S/S ratio (v/rn)

10 30 60 Extraction duration (min)

100 175 250 Shaking speed (rpm)

I 100 1 'i:

Ii 80

, U 60 - ~ I +-'

I] 40 I~ I 20

I l

CHX EA MEK Solvent type

- --

2:1 3:1 4:1 10 30 60 100 175 250 S/S ratio (v/ml_ Extraction duration (min) . . Shaking~~ed (q~mj

Figure 4.2 Influence of experimental factors on the (a) oil and (b) solvent recovery in

MSE process

Table 4.5 MSE oil recovery results and statistical analysis

Oil recovery(%) Impact factors ------------. Statistical analysis

Rep. I Rep.2 Rep.3 A B C D

124

46. 7 48 .9 45 .1 K11 483.0 354.6 360.6 347.2

2 46.1 53.2 57.1 K21 340.6 374.8 356.3 378.5

3 59.5 63.7 62.7 KJJ 299.7 394.0 406.5 397.6

4 36.0 36.3 36.4 ka J 9 9 9 9

5 36.1 36.9 42.3 KJ/kJ 53.7 39.4 40.1 38.6

6 38 .5 39.5 38.8 K2/kJ 37.8 41.6 39.6 42.1

7 35.5 34.8 35.1 KJ/kJ 33.3 43.8 45.6 44.2

8 32.6 32.4 38.1 R1 20.4 4.4 5.6 5.6

9 33.5 29.0 28.8 Order A>D>C>B

Optimal level 3 3 3

Optimal organization A1D3C3B3

a The number of the appearance for a specific level

Table 4.6 ANOV A for oil and solvent recovery results from MSE

Oil recovery: Sources of variation ss a df b MS C Fvalue Fa.d Significance e

(A) Solvent type 2056.7 2 1028.4 134.4 F oos(2,18) = 3.6 **

(B) SIS ratio (vim) 86.3 2 43 .1 5.6 F o.01(2,18) = 6.0 *

(C) Extraction duration (min) 172.0 2 86.0 11.2 **

(D) Shaking speed (rpm) 144.0 2 72.0 9.4 **

Error 137.7 18 7.6

Total 2596.7 26

Solvent recovery: Sources of variation ss df MS F value Fa. Significance

(A) Solvent type 383.7 2 191.9 6.6 F oos(2,18) = 3.6 **

(B) SIS ratio (vim) 535.7 2 267.8 9.3 F o.01(2,18) = 6.0 **

125

(C) Extraction duration (min) 174.5 2 87.2 3.0

(D) Shaking speed (rpm) 28 .9 2 14.5 0.5

Error

Total

520.7 18 28.9

1643.5 26

a Sum of squares; b Degrees of freedom; c Mean square; d the critical F value; e ** P < 0.01 , * P < 0.05 .

Solvent recovery is another important aspect of extraction treatment because low

solvent recovery could not only increase treatment cost but also cause environmental

problems such as escaped solvent vapor emission. As shown in Figure 4.2b, more than 70%

of solvent can be recovered after MSE treatment. Among the three solvents, MEK showed

the highest solvent recovery of 83.3%. Using higher solvent-to-sludge (S/S) ratio could

produce higher solvent recovery rate. With the increase of S/S ratio from 2: 1 to 4: 1, the

solvent recovery rate increased from 72.8% to 83.1 %. The influence of solvent type and

S/S ration on solvent recovery was significant. However, extraction duration and shaking

speed had little effect on the solvent recovery as the F-ratio was much lower than the

critical F-value (Table 4.6) . This might be due to the fact that the extraction process was

performed in an enclosed system (i.e., sealed tubes), and thus higher shaking speed and

longer extraction duration would not cause more solvent loss. Solvent loss might be due

to the unavoidable evaporation and adhesion to the inner wall of container during the

extraction process of mixing, settling, and liquid transferring. The solvent loss in a nearly

identical extraction system (i.e., same container and distillation apparatus) could be similar

for the same solvent, so the solvent recovery would be calculated as lower when the lower

S/S ratio was chosen for extraction.

126

80

'*- 60 ~ Q) > 0 u 40 Q) ... 0

20

0

The results of single factor MSE experiment (Figure 4.3) indicated that under the

same extraction conditions, the MSE treatment using CHX could achieve higher oil

recovery rate than EA and MEK. There was little difference in oil recovery between the

treatment of EA and MEK. The results also revealed that higher values of S/S ratio,

extraction duration, and shaking speed were preferable for oil recovery. Oil recovery can

reach equilibrium when the S/S ratio, extraction duration, and shaking speed reached to

4: 1, 60 min, and 250 rpm, respectively. Further increase the value of these factors only

brought limited enhancement, but could increase the cost and duration of treatment.

Therefore, it is reasonable to set the optimal MSE condition as S/S ratio of 4: 1, extraction

duration of 60 min, and shaking speed of 250 rpm for economical consideration.

- ~ ~ CHX EA MEK

Solvent type

2:1 4:1 8:1 S/S ratio (v/m)

--10 60 120

Extraction duration (min)

Figure 4.3 Results of single-factor experiments ofMSE

127

150 250 300 Shaking speed (rpm)

4.3.2 Ultrasonic assisted extraction using a probe system (UAEP)

Figure 4.4a presents the effects of four experimental factors on UAEP performance.

Similar to MSE process, CHX showed much higher oil recovery than the other two

solvents. Oil recovery of UAEP increased with S/S ratio, treatment duration and ultrasonic

power. It is noteworthy that the highest oil recovery obtained by CHX was 59.8% in UAEP,

which is higher than the highest oil recovery of 53.7% in MSE by the same solvent. The

extraction using MEK was also improved by UAEP with the highest oil recovery of38.3%

compared to 33 .3% by using MSE. According to the Rj value shown inin Table 4.7, the

influence of four factors on oil recovery was ranked in the order of solvent type, S/S ratio,

treatment duration, and ultrasonic power. Among these four factors, ultrasonic power

showed insignificant effect on oil recovery while the effects of other three factors were

significant (Table 4.8), indicating that the ultrasonic power of 21 W was sufficient to

generate enough agitation for the mixing of oil and solvents. As compared to the longer

extraction duration required by MSE, the UAEP can achieve satisfactory performance

within a very short duration (i.e. in seconds).

128

Runs

1

2

3

4

5

1 80 ' 1~60 ~ (lJ

: ~ 40 · U I ~ I Q 20

r 100 ~ i 80

I~ :rl 60 ... ... 11 40 . V'l

CHX EA MEK Solvent type

..-• •

2:1 3:1 4:1 5/S ratio (v/m)

2:1 3:1 4:1 S/S ratio (v/m)

• • •

10 15 20 Treatment duration (s)

a

• • •

21 48 66 Ultrasonic power (w)

b

10 15 20 21 48 66 Treatment duration (s) Ultraso~ic power (w~

Figure 4.4 Influence of experimental factors on the (a) oil and (b) solvent recovery in

UAEP process

Table 4. 7 UAEP oil recovery results and statistical analysis

Oil recovery(%) Impact factors Statistical analysis

Rep.I Rep.2 Rep.3 A B C D

55.7 55.3 49.8 K11 538.3 375.2 386.8 388.5

58.4 55.6 67.4 K21 323.6 406.0 400.8 405.7

63.9 64.4 67.8 K11 344.7 425.3 419.0 412.3

33.9 33.7 34.2 k·" 1 9 9 9 9

37.2 36.7 36.2 K1/"1 59.8 41.7 43 .0 43.2

129

6 36.1 37.5 38.0 Ki/"1· 36.0 45.1 44.5 45.1

7 37.1 38.1 37.5 KJ/"1· 38.3 47.3 46.5 45.8

8 38.4 38.7 37.3 R1 23.9 5.6 3.6 2.6

9 36.5 37.8 43.2 Order A>B>C>D

Optimal level 3 3 3

Optimal organization A1B3C3D3

a The number of the appearance for a specific level

Table 4.8 ANOVA for oil and solvent recovery results from UAEP

Oil recovery: Sources of variation ssa df b MSC Fvalue Fad Significance e

(A) Solvent type 3112.8 2 1556.4 206.6 Foos(2,18) = 3.6 ** (B) S/S ratio (v/m) 141.6 2 70.8 9.4 Foo1(2,18) = 6.0 ** (C) Extraction duration (s) 57.7 2 28.8 3.8 * (D) Ultrasonic power (W) 33.4 2 16.7 2.2

Error 135.6 18 7.5

Total 3481.1 26

Solvent recovery: Sources of variation ss df MS Fvalue Fa Significance

(A) Solvent type 548.5 2 274.2 24.0 Foos(2,18) = 3.6 ** (B) S/S ratio (v/m) 148.6 2 74.3 6.5 Fo.01(2,18) = 6.0 ** (C) Extraction duration (s) 25 .0 2 12.5 1.1

(D) Ultrasonic power (W) 44.4 2 22.2 1.9

Error 206.0 18 11.4

Total 972.4 26

a Sum of squares; b Degrees of freedom; c Mean square; ct the critical F value; e ** P < 0.01, * P< 0.05.

130

The effects of the experimental factors on solvent recovery are shown in Figure 4.4b.

It can be seen that more than 80% of CHX and MEK can be recovered after UAEP

treatment, while a lower recovery rate (i.e. , 76.7%) was observed for EA. Higher S/S ratio

was beneficial to the solvent recovery. For example, the solvent recovery increased from

79.8% to 84.8% as the S/S ratio increased from 2: 1 to 4: 1. Similar to MSE, the solvent

type and S/S ratio showed significant effect on solvent recovery in UAEP treatment (Table

4.8). Moreover, longer treatment duration and higher ultrasonic power showed

insignificant effects on solvent recovery (Table 4.8). The results indicate that UAEP would

not cause negative influence on solvent recovery by noting the higher solvent recovery

than MSE treatment.

The results of single factor experiments ofUAEP are illustrated in Figure 4.5. Under

the same condition, CHX can recover as high as 68.8% of oil from oily sludge, while EA

and MEK achieved an oil recovery of 35.0% and 37.2%, respectively. The extraction

performance can be significantly improved (p<0.05) by UAEP using CHX and MEK with

about 8% (CHX) and 3% (MEK) increment in oil recovery as compared to the highest oil

recovery obtained in MSE by the same solvent. However, little difference (p>0.05) was

observed between UAEP and MSE treatments when using EA as the extraction solvent.

Similar to MSE, a promising oil recovery can be obtained when the S/S ratio reached 4: 1

in UAEP, and further increase S/S ratio lead to no improvement in oil recovery. The oil

recovery rate was also increased from 55. 7% to 68.4% as treatment duration extended from

10 s to 20 s, but there is no need to further extend the treatment duration because increasing

duration from 20 s to 3 0 s brought no obvious increase in oil recovery. In fact, longer

treatment duration might cause higher amount of solvent loss and energy consumption.

131

80

~ 60 c:-C1J > 0 u 40 C1J ... 0

20

0

Furthermore, there is no significant effect on oil recovery by alteration of ultrasonic power

from 21 W to 75 W, indicating low ultrasonic power could still produce sufficient

turbulence for mixing the oil and solvent in the extraction bulk system. Therefore, the

optimum extraction condition for UAEP was identified as: ultrasonic power of 21 W,

treatment duration of 20 s, and S/S ratio of 4: 1. In summary, compared with 60 min of

MSE, UAEP process can recover more oil from oily sludge in a much shorter duration (i.e.,

20 s) without compromising solvent loss, indicating that it is a high efficiency extraction

approach for oily waste treatment.

- - -CHX EA MEK

Solvent type

2:1 4:1 8:1 S/S ratio (v/m)

10 20 30 21 66 75 Treatment duration (s) Ultrasonic power (w)

Figure 4.5 Results of single-factor experiments ofUAEP

The high extraction efficiency of UAEP might be due to its cavitation phenomena

when acoustic power input is sufficiently high to generate microbubbles at nucleation spots

in the extraction system. The violent collapse of microbubbles when they reach an unstable

size would produce shock waves and rapid increase in temperature (Pilli et al., 2011). The

shock waves and fluid jets caused by cavitation can produce turbulence in the extraction

system and break the aggregation of solids, therefore expose more oil components to

132

solvent and increase the miscibility of oil in solvent in a short time (Pilli et al., 2011 ). The

temperature increase in the extraction system may also reduce the viscosity of oily sludge

and thus improve solvent extraction. Numerous studies have documented that ultrasonic

treatment is a fast, affordable, and highly effective extraction method for crude oil from

oil sands (Abramov et al. , 2009), organic compounds from soil and sewage sludge (Flores

et al., 2007; Dewil et al., 2006), and vegetable oil from agricultural products (Chemat et

al., 2011). This study has confirmed that ultrasonic extraction could be an effective

alternative approach for oil recovery from oily sludge.

4.3.3 Ultrasonic assisted extraction using a bath system (UAEB)

As illustrated in Figure 4.6a, oil recovery ofUAEB process can be affected by solvent

type, S/S ratio, and treatment duration. However, the effect of ultrasonic bath temperature

was not significant. The highest oil recovery of 62.6% was found in the treatment using

CHX, while EA and MEK can recover nearly equal amount of oil (i.e., 36.9%) from oily

sludge. Oil recovery also increased with S/S ratio and treatment duration. As shown in

Table 4.9, the influence of the four experimental factors was ranked in an order of solvent

type > SIS ratio > treatment duration > bath temperature. All of the factors showed

significant effects on oil recovery except bath temperature, indicating raising bath

temperature was unnecessary in UAEB treatment (Table 4.10).

133

Runs

1

2

3

,--- - - -80

e6o \._. a

I ~ I Q.J • • • ~40

1@20 I

0 CHX EA MEK 2:1 3:1 4:1 5 10 15 20 40 60

Solvent type S/S ratio {v/m) Treatment duration (min) Bath temperature (0() - """'- - - --,

100

* ~ ... ·-so • • If I ~ 60 ... ....

1140 b

l 20

CHX EA MEK 2:1 3:1 4:1 5 10 15 20 40 60 Solvent type S/S ratio (v/m) Treatment duration (min) Bath temperature (0

()

- - - ---'

Figure 4.6 Influence of experimental factors on the (a) oil and (b) solvent recovery in

UAEB process

Table 4.9 UAEB oil recovery results and statistical analysis

Oil recovery (%) Impact factors Statistical analysis

Rep.I Rep.2 Rep.3 A B C D

55.8 57.6 56.8 KJJ 563.7 372.0 394.9 408.8

58.4 66.8 60.4 K2j 332.5 422.2 397.4 402.4

72.0 69.6 66.2 K3j 332.6 434.6 436.5 417.6

134

4 32.9 32.5 31.4 kl 9 9 9 9

5 38.9 40.5 44.3 K1/ k1 62.6 41.3 43.9 45.4

6 42.0 35.1 34.8 K2/'9 36.9 46.9 44.2 44.7

7 35.0 35.6 34.3 K3/'9 37.0 48.3 48.5 46.4

8 37.6 37.8 37.5 RJ 25.7 7.0 4.6 1.7

9 40.0 37.8 37.1 Order A>B>C>D

Optimal level 1 3 3 3

Optimal organization A1B3C3D3

a The number of the appearance for a specific level

Table 4.10 ANOV A of oil and solvent recovery results from UAEB

Oil recovery: Sources of variation ss• df b MS C Fvalue Fad Significance e

(A) Solvent type 3957.6 2 1978.8 313.7 Fo.os(2, 18)=3.6 **

(B) S/S ratio (v/m) 244.5 2 122.2 19.4 F o.0 1(2,18)=6.0 **

(C) Extraction duration (min) 120.8 2 60.4 9.6 **

(D) Bath temperature (°C) 12.8 2 6.4 1.0

Error 113.6 18 6.3

Total 4449.2 26

Solvent recovery: Sources of variation ss df MS Fvalue Fa Significance

(A) Solvent type 293.7 2 146.8 30.4 Foos(2,18)=3.6 **

(B) S/S ratio (v/m) 249.3 2 124.6 25.8 Fo.o, (2, 18)=6.0 **

(C) Extraction duration (min) 33.4 2 16.7 3.5

135

(D) Bath temperature (0 C) 92.6

87.0

755 .9

2

18

26

46.3

4.8

9.6 ** Error

Total

a Sum of squares; b Degrees of freedom; c Mean square; d the critical F value; e ** P <

0.01, * P < 0.05.

The effects of the experimental factors on solvent recovery from UAEB are shown in

Figure 4.6b. Compared to CHX and EA, MEK showed higher solvent recovery

performance. Similar to MSE and UAEP, the solvent recovery was improved as more

solvent was used for extraction. There was no obvious variation in solvent recovery as the

treatment duration extended from 5 to 15 min, but bath temperature showed a negative

effect on solvent recovery for the reason that solvent recovery declined by about 4% as the

bath temperature increased from 20 to 60 °C.

The results of single-factor experiments are shown in Figure 4. 7. The oil recovery of

CHX, EA, and MEK in UAEB treatment was 67.6%, 35.0%, and 34.7%, respectively.

There was no significant difference in oil recovery between UAEP and UAEB treatment

when the same solvent was applied, but the oil recovery of UAEP and UAEB were

significantly higher than that of MSE treatment using the same solvent. The oil recovery

increased from 56.3% to 67.3% as the SIS ratio changed from 2:1 to 4:1, but only a trivial

increment in oil recovery was observed when this ratio further increased from 4: 1 to 8: 1.

The variation of oil recovery caused by the extension of ultrasonic treatment duration also

showed similar trends. A significant increase in oil recovery was obtained as treatment

duration increased from 5 to 15 min, but little enhancement was observed when the

treatment duration exceeded 15 min. In agreement with the results from orthogonal

136

experiments, the bath temperature of UAEB process had limited effect on oil recovery.

Therefore, the optimum condition for UAEB treatment was set as: S/S ratio of 4: 1,

treatment duration of 15 min, and bath temperature of 20 °C.

The results indicated that in comparison to MSE, UAEB can obtain a relatively higher

oil recovery within shorter treatment duration. The oil recovery of UAEB and UAEP by

three solvents under each optimal condition was similar, but longer treatment duration (i.e.,

20 s of UAEP as compared to 15 min of UAEB) and higher energy consumption (i.e. 21

W of UAEP as compared to 80 W of UAEB) was required by UAEB treatment. For

example, the total energy consumption of each treatment by UAEP, UAEB, and MSE

processes were 1.17x 1 o-4, 0.02, and 4.5 KWh, respectively. Higher energy consumption

in UAEB process might be due to the fact that UAEP treatment is a much more powerful

process because the ultrasonic energy is delivered on a small area around the ultrasonic

probe, but ultrasonic energy can be highly attenuated by the water contained in the bath

and the walls of the container used in UAEB treatment (Chemat et al. 2011). As a result, a

higher ultrasonic power and longer extraction duration are required by UAEB to achieve

satisfactory extraction performance. Nevertheless, UAEB process can handle a much

larger volume of oily sludge than UAEP in one treatment, and thus it has great potential to

be applied to large-scale treatment. Many safety, equipment design, and treatment cost

issues should be thoroughly considered before introducing UAE process to large-scale oily

sludge treatment. For example, the safety concerns may include but not limited to the

escaping emission of solvent vapor, the corrosion of pipelines caused by water in oily

sludge, and the heat and pressure generated during distillation. Treatment cost also should

be carefully calculated in terms of the loss of solvent by evaporation and/or leaking, energy

137

consumption, and the maintenance of equipment for the reason that extra costs might occur

from these aspects.

l 80 70

~ 60 QJ > 0 u Q) .... 0

so 40 30 20 10

0 - ~ ~ CHX EA MEK

Solvent type

2:1 4:1 8:1 S/S ratio (v/m)

5 15 30 Treatment duration (min)

Figure 4.7 Results of single-factor experiments ofUAEB

4.4.4 Impact of extraction cycles

--20 60 80

Bath temperature (0 C)

As seen in Figure 4.8, the oil recovery generally increased with the extraction cycles

for the three extraction processes. In MSE treatment, CHX recovered 60.5% of oil from

oily sludge in the first extraction, and oil recovery increased to 68.5% in the second

extraction. For EA and MEK, the increment in oil recovery was 8.3% and 7.9%,

respectively as the second extraction was introduced. After two extraction cycles, there

was only trivial increase in oil recovery, suggesting that it is unnecessary to have more

than two extraction cycles in MSE for economical consideration. For UAEP treatment,

CHX recovered 67.3% of oil in the first extraction, which was significantly higher than

that of MSE. However, there was only 4.6% increase in oil recovery in the second

extraction cycle, which was much lower than that ofMSE. This might be because most oil

was extracted in the first extraction in UAEP treatment, but in MSE process there was

138

some oil still trapped in oily sludge after the first extraction. The trapped oil can be

extracted in the second extraction, so that a relatively larger increase in oil recovery can

be observed in MSE. The effect of extraction cycles on oil recovery ofUAEB was similar

to that of UAEP. After the second extraction, the oil recovery of CHX, EA, and MEK in

UAEB increased by 4.4%, 6.6%, and 5.9%, respectively. The results indicate that two

extraction cycles could be sufficient for all treatment processes to achieve a satisfactory

oil recovery.

139

r- 80

I 'ii 60 c:-(J) > 40 0 u (J) L.

0 20

O CHX O EA A MEK

0 1 2 3 4

Number of cycles 80

b0- e e E) 'cf.

60 c:-(J) > A 0 0 i u 40 (J) 4fs ... 0

20 O CHX ¢ EA ti MEK

0 1 2 3 4

Number of cycles 80

l C e e E)

60 er-->-... (J) > 0 u 40 Q.I ... 6

20 O CHX ~ EA li MEK

0 1 2 3 4

Number of cycles

Figure 4.8 Effect of extraction cycles on oil recovery in (a) MSE, (b) UAEP, and (c)

UAEB treatment

140

4.3 .5 Characteristics of recovered oil

The appearance of recovered oil using different solvents was shown in Figure 4.9.

Recovered oil by CHX was darker than the recovered oil by MEK and EA. In addition, it

can be seen from Figure 4.9 that higher amount of oil can be recovered by CHX than by

EA or MEK from the same amount of oily sludge. The oil recovery rate was used for the

comparison of three solvent extraction processes as described above. To make a further

evaluation, we used TPH recovery and the characteristics of recovered oil for comparison.

TPH recovery was calculated as the ratio of TPH mass in the recovered oil to the TPH

mass in the original oily sludge. Figure 4.1 Oa presents the TPH recoveries of the three

extraction processes under their optimal operating conditions. Among the three solvents,

CHX was associated with the highest TPH recovery for each extraction process. This might

be due to the higher miscibility of CHX with asphaltenes in oily sludge (John and Joseph,

2008), which are the compositions of heavy TPH fractions (e.g. , F3 and F4 fractions) .

Generally, more than 80% of TPH in oily sludge can be recovered by the three extraction

processes. However, the ultrasonic assisted solvent extraction (UAE) showed a better TPH

recovery as compared to MSE. For example, UAE treatment increased TPH recovery by

about 7% and 3% when using CHX and MEK solvents as compared to MSE, although

little improvement was observed in treatment using EA. The TPH recovery reached 92.8%

in UAEP treatment using CHX (Figure 10a). High TPH recovery indicated that after

solvent extraction only a low amount of TPH were left in the treatment residues which

could be easier for remediation than the original oily sludge.

As shown in Figure 1 Ob, the TPH content of recovered oil using CHX was the lowest

for all the three extraction processes. For example, it is 47.1 % when using CHX in MSE

treatment, which is lower than that for EA (79. 8%) and MEK extraction (81.6% ). Although

141

CHX showed the highest oil and TPH recovery in all the three extraction processes, the

recovered oil contained less valuable TPH content, suggesting that the recovered oil by

CHX contained a higher amount of unwanted impurities. The impurities could be remnant

water, natural polar substances and fine solid particles. It has been reported that about 40%

of heptane insoluble asphaltenes in petroleum residua dissolves in CHX (John and Joseph,

2008). Asphaltenes in extractants tend to form asphaltenic colloids because they have a

propensity to aggregate, flocculate, and adsorb onto interfaces of water and fine solid

particles, resulting in those remnant water and fine solid particles which cannot be readily

removed from extractant by gravitational setting (Spiecker et al., 2003). It can be found

from Figure 1 Ob that in general the UAE is associated with a higher TPH in the recovered

oil than MSE. UAEP treatment obtained the highest TPH content ( e.g., 51.6%, 83 .6%, and

85.8% when using CHX, EA, and MEK, respectively), although the TPH content in

recovered oil by UAEB treatment was only slightly higher than that by MSE treatment.

These results indicated that UAE process could improve both the quantity and quality of

recovered oil as compared to MSE.

142

Figure 4.9 Recovered oils by different solvents from UAEP treatment

143

100 a c;J CHX l!!IEA ll!MEK

90

*-c:- 80 C1J > 0 u C1J ... :I: 70 Q.. I-

60

50 MSE UAEP UAEB

Treatments

100 ::;J CHX CEA liii!MEK

80

*-C1J 1:11) 60 ra .... C: C1J

)~~::) u ... C1J 40 a. ~::::::::::::~ :I: ~~~~~~~ Q.. I- ~::~~::::~

20 ~!1!1!1111 f:::::~:::a

YYYYYYYYY-.,",

0 .v":-':.)'.v:'.v':.-;'!

MSE UAEP UAEB

Treatments

Figure 4.10 TPH recovery ( a) and TPH content in recovered oil (b) of three extraction

processes (MSE: SIS ratio of 4: 1, shaking speed of 250 rpm, and duration of 60 mins;

UAEP: SIS ratio of 4: 1, ultrasonic power of 21 W, and duration of 20 s; UAEB: SIS ratio

of 4: 1, duration of 15 min, and bath temperature of 20 °C)

Figure 4.11 presents the PHCs distribution in recovered oil by three extraction

processes. The recovered oil mainly consists of F2 and F3 fractions, while there is also a

144

significant amount of F4 fraction in the recovered oil extracted by CHX, showing that

CHX has better affinity to the heavy PHC components in oily sludge. The PHCs

distribution was similar when the same solvent was utilized, regardless of the extraction

processes. For example, the F3 fraction percentage of recovered oil by EA extraction from

MSE, UAEP, and UAEB were 44.4%, 42.7% and 41.3%, respectively. Zhang et al. (2012)

investigated the individual effect of ultrasonic irradiation on oil recovery from oily sludge,

and found that ultrasonic irradiation alone could not cause significant change of PHCs

distribution in recovered oil. Similar PH Cs distribution in recovered oil of three different

extraction treatments indicates that there was no significant destruction effect on PHCs by

ultrasonic treatment under the experimental conditions in this study.

145

1 60

~ 40 (1) ll.D ro ..... C: Q) 20 u '-Q)

CL.

0

60

~ Q.J 40 ll.D ro ..... C: Q) u '-Q) 20 CL.

0

60

if: 40 Q) ll.D ro ..... C: Q) u 20 '-Q)

CL.

I I

L

CHX

b

'' :

CHX

C

CHX

ml F2 0 F3 ~ F4

EA Solvents

LI F2 D F3 IZI f 4

.·l!

EA Solvents

~ F2 DF3 El:IF4

EA Solvents

MEK

MEK

' l

J MEK

Figure 4.11 Distribution of PH Cs fractions in the recovered oil by using three solvents in

(a) MSE, (b) UAEP, and (c) UAEB (MSE: S/S ratio of 4:1, shaking speed of250 rpm,

146

and duration of 60 mins; UAEP: S/S ratio of 4: 1, ultrasonic power of 21 W, and duration

of 20 s; UAEB: S/S ratio of 4: 1, duration of 15 min, and bath temperature of 20 °C)

The properties of recovered oil by different solvents in UAEP process are shown in

Table 4.11. Comparing with crude oil, recovered oil by MEK and EA showed higher

density, carbon residue, and viscosity. The calorific value of recovered oil by MEK and

EA was also higher than that of crude oil, indicating that more heat can be obtained if using

recovered oil as an energy source. The asphaltene content in recovered oil by MEK and

EA were much lower than that in crude oil, which is preferable in terms of ensuing refining

process. In contrast, recovered oil by CHX contained higher water and asphaltene content

than other oils, and this was in agreement with the TPH content analysis. The heavy metals

concentrations in recovered oil were generally very low, indicating that most heavy metals

in oily sludge were transferred to solid residues after extraction treatment.

Table 4.11 Properties of crude oil and recovered oil (RO) by different solvents in UAEP

Parameters Methods Crude oil R.O. CHX R.O.MEK R.O.EA

Density @25°C (g/mL) ASTMD4052 0.8271 0.8260 0.8465 0.8487

Carbon Residue (%) ASTMD4530 0.22 0.31 0.49 0.42

Water content (mg/kg) ASTMD6304 1349 2253 118 330

Viscosity @40°C (mm2/s) ASTMD445 4.02 3.92 4.72 4.47

Calorific Value (BTU/lb) ASTMD240 16197 16400 18169 19317

Surfur content(%) ASTMD4294 0.56 0.64 1.165 1.037

Asphaltenes (%) ASTMD3279 1.59 2.21 0.39 0.46

147

Metals (mg/kg) ASTMD5185

Iron (Fe) 11 34 2 4

Aluminum (Al) <1 2 <1 <1

Chromium (Cr) <1 <1 <1 <1

Copper (Cu) <1 <1 1 <1

Lead (Pb) <1 <1 <1 <1

Ni (Ni) 1 1 <1 <1

Zinc (Zn) 1 < l <1

Vanadium (V) 4 2 1 1

4.4 Conclusions

The effects of two ultrasonic assisted solvent extraction (i.e., UAEP and UAEB)

processes on oil recovery from oily sludge were investigated in comparison with that of

MSE. The impacts of several operational factors on each treatment process were examined

through an orthogonal experimental design. In MSE treatment, solvent type, solvent-to-

sludge (S/S) ratio, extraction duration, and shaking speed had significant effects on oil

recovery. The highest oil recovery ofCHX, EA, and MEK were 63.7%, 35.2%, and 34.8%

under a S/S ratio of 4: 1, an extraction duration of 60 min, and a shaking speed of 250 rpm,

respectively. In UAEP process, solvent type, S/S ratio, and treatment duration illustrated

significant effects on oil recovery while the influence of ultrasonic power was not obvious.

The oil recovery by CHX, EA, and MEK in UAEP reached 68.8%, 35.0%, and 37.4%,

respectively. Compared to MSE, UAEP can achieve a higher oil recovery within a much

shorter duration (i.e., 20 s) and less energy consumption (i.e., 21 W). The oil recovery

148

result of UAEB treatment was also satisfactory as compared to that of MSE, while its

treatment duration (i.e. , 15 min) was shorter than MSE. The results showed that two

extraction cycles could be sufficient to obtain a promising oil recovery in all extraction

treatments, while UAE processes would not cause more solvent loss than MSE treatment.

Among all three solvents, the TPH content in recovered oil using CHX was the lowest (i.e. ,

about 50%), while EA and MEK can recover relatively high TPH content in oil (i.e., >80% ).

The recovered oil mainly consisted ofF2 and F3 fractions, and its physical properties were

similar to those of crude oil, indicating that it can be utilized as a feedstock for downstream

upgrading process. By comprehensively considering the oil recovery, solvent recovery,

TPH recovery, as well as the TPH content and other characteristics of the recovered oil,

the UAE process using MEK would provide promising performance as compared to

conventional MSE process. In summary, UAE could improve both the quantity and quality

of recovered oil from oily sludge within a short treatment duration, and thus it represents

an attractive alternative extraction method for the treatment of refinery oily sludge.

149

Chapter 5 A combination of solvent extraction and freeze thaw for oil recovery from refinery wastewater treatment pond sludge *

* The results ofthis work have been published as: Hu, G.J., Li, J.B., and Hou, H.B., 2015.

A combination of solvent extraction and freeze thaw for oil recovery from petroleum

refinery wastewater treatment pond sludge. Journal o_f Hazardous Materials, 283, 832-840.

5 .1 Introduction

The petrochemical refineries generate a large volume of oily sludge from a wide range

of sources, including crude oil tank bottom sediments, slop oil emulsion solids, residues

from oil/water separators, and dredged sludge from on-site wastewater treatment pond (Hu

et al., 2013). As a very recalcitrant mixture, oily sludge is mainly consisted of water,

petroleum hydrocarbons (PHCs), and solids (Hu et al., 2013; Ramaswamy et al., 2007). It

has been recognized as a hazardous waste in many countries, and its improper management

can adversely impact the environment and public health (Hu et al., 2013; Robertson et al.,

2007; Al-Mutairi et al., 2008). Given the relatively high oil content in sludge, oil recovery

could be the most desirable management option since it can not only generate profit but

also reduce waste volume and pollutant concentration (da Silva et al., 2012; Hu et al., 2013;

Elektorowicz and Habibi, 2005). Driven by increasingly stringent regulations which have

banned the direct land disposal of oily sludge, various physical, chemical, and biological

methods have been developed for its treatment (Hu et al., 2013; da Silva et al., 2012).

Among them, solvent extraction is a simple process by mixing oily waste and solvent in

an appropriate proportion to ensure adequate miscibility of oil in solvent, while most water

and solids are rejected as unwanted impurities which can be removed by gravitational

settling or centrifugation. The oil and solvent mixture (i.e. extractant) can then be separated

150

by distillation for both oil and solvent recycling (Al-Zahrani and Putra, 2013). Many

studies have reported the use of various solvents for recovering oil from oily sludge. Avila-

Chavez et al. (2007) employed supercritical ethane and dichloromethane to treat oily

sludge in a Soxhlet extraction system, and found that more than 50% of oil can be

recovered. Taiwo and Otolorin (2009) utilized hexane and xylene as the solvents to extract

PH Cs from oily sludge, and the highest oil recovery rate of 67.5% was observed by using

hexane. Abouelnasr and Zubaidi (2010) found that methyl ethyl ketone (MEK) and light

petroleum gas condensate (LPGC) were able to extract 39% and 32% by mass of the

original sludge as recovered oil, respectively. The quality of recovered oil in terms of solids

and asphaltene content was also observed to improve with increasing amount of solvent

use (Abouelnasr and Zubaidi, 2010).

However, most of the previous solvent extraction studies focused on tank bottom

sludge, and little attention has been paid to the petroleum refinery wastewater treatment

pond sludge which contains more water and less oil as compared to tank bottoms. In

general, such high-moisture sludge is prone to form emulsified water in the solvent

extraction system under vigorous agitation, and thus the water cannot be effectively

removed by ensuing gravitational settling, leading to a high remnant water content in the

extractant (Yang et al., 2009; Lin et al., 2008). Moreover, the extractant could contain

undesirable impurities (e.g., dissolved salts and fine solids) which could compromise the

quality ofrecovered oil. If the recovered oil is used as a feedstock for further reprocessing,

it can cause problems such as the erosion of refining equipment and the poisoning of

refining catalysts (Abdul et al., 2006; Mahdi et al., 2008). As a result, it is imperative to

dewater the extractant (He and Chen, 2002). Freeze/thaw (FIT) treatment has been reported

as a cost-effective dewatering process (Chen and He, 2003). The volume expansion of

151

water droplets when turning to ice could cause the coalescence of emulsified water droplets

and the change of interfacial tension between water and oil phases, and these were the main

driving forces of dewatering (Lin et al., 2007). Chen and He (2003) reported that

freeze/thaw treatment removed nearly 90% of water from a high-moisture oily sludge.

Another study on used lubricating oil based emulsion also found a >90% of water removal

by freeze/thaw treatment (He and Chen, 2002). Zhang et al. (2012) reported that

freeze/thaw could reduce the concentration of total petroleum hydrocarbon (TPH) in the

wastewater separated from oily sludge treated with ultrasonic irradiation. However, limited

studies have been reported to examine the combined effect of solvent extraction and

freeze/thaw.

The objective of this chapter is to investigate the performance of oil recovery from

high-moisture petroleum sludge by using a combination of solvent extraction and

freeze/thaw treatment. The efficiency of several readily available solvents was firstly

compared based on their oil extraction capability from sludge, together with the impacts

of various factors such as solvent-to-sludge ratio, extraction duration and extraction cycles.

The solvent recovery rate and the mass reduction of sludge were also examined. Based on

their extraction effectiveness, three solvents were selected to further evaluate the

performance of freeze thaw treatment on improving the quality of recovered oil in terms

of its TPH content. Three groups of experiments were conducted, including freeze/thaw

treatment alone, solvent extraction alone, and combined solvent extraction with

freeze/thaw. The TPH content and PHCs fraction distribution in the recovered oil were

reported, and the properties of solid residue as the by-product of the treatment process were

also analyzed. In this way, the enhancement effect of freeze/thaw on oil recovery using

solvent extraction was evaluated.

152

5 .2 Experimental materials and methods

5.2.1 Materials

The dredged sludge from a petroleum refinery wastewater treatment pond in western

Canada was used as the study sample. A simple random sampling method as recommended

in US EPA SW-846 guideline was used for collecting sludge, while its TPH was

determined as the target contaminant that is of concern in statistical analysis (EPA, 1994).

Figure 5.1 shows the appearance of dredged oily sludge. Compared to tank bottom oily

sludge, dredged wastewater pond sludge was more watery and less sticky. The sample was

stored in a sealed glass jar at 25 °C. It was well stirred manually to obtain an evenly mixing

before use in the experiments. Table 5.1 lists the property of wastewater pond dredged

sludge. The TPH concentration was measured using the US EPA method 80 l 5C (EPA,

2007), and the water content was analyzed by sample drying based on ASTM D2974-00

Method A (ASTM, 2013), while the solid content was calculated according to the

measured TPH and water contents. The metal elements were measured using Inductively

Coupled Plasma (ICP) analysis based on ASTM D5 l 85 (ASTM, 2009). As shown in Table

5.1, the dredged oily sludge contained relatively high amount of water but less amount of

oil, which could pose negative effect such as causing undesirable water-oil emulsions on

oil recovery treatment. In this study, the PHCs in the recovered oil were analyzed and

compared with those of crude oil sample which was obtained from a petroleum refinery in

western Canada. Based on the consideration of toxicity, oil miscibility, economics,

viscosity, and vapor pressure (Hamad et al., 2005; Kheireddine et al., 2013), five organic

solvents were selected for experiments, including cyclohexane (CHX), dichloromethane

153

(DCM), methyl ethyl ketone (MEK), ethyl acetate (EA), and 2-propanol (2-Pro ). These

solvents with HPLC grade (>99%) were purchased from Sigma Aldrich, Canada.

Figure 5.1 Dredged oily sludge from a petroleum refinery wastewater treatment pond

Table 5.1 Properties of petroleum refinery wastewater treatment pond sludge

Parameter value Parameter Concentration (mg/kg)

TPH 19.9% (by mass) Antimony (Sb) 3.3

Water content 63.0% (by mass) Arsenic (As) 68.7

Solid content 17.1 % (by mass) Beryllium (Be) 0.3

154

5.2.2 Solvent extraction

5.2.2.1 Extraction procedure

Cadmium (Cd)

Chromium (Cr)

Copper (Cu)

Zinc (Zn)

Lead (Pb)

Nickel (Ni)

Cobalt (Co)

2.0

70.0

159.0

2040.0

64.7

58.5

8.4

Two main factors including solvent-to-sludge (S/S) ratio (v/m) and extraction

duration were investigated in this study to find the desired solvent extraction condition for

oil recovery, and Table 5.2 lists the experimental factor levels, with S/S ratio ranging from

1: 1 to 8: 1, and extraction duration varying from 5 to 90 min. In terms of the extraction

experiment for each solvent, about 5 g of oily sludge was placed in a 50-mL centrifugation

tube, and a given volume of solvent was added according to the specified S/S ratio in Table

5.2. The centrifugation tube was then sealed and placed on a mechanical shaker (Talboy

3500 Orbital Shaker) for shaking extraction at 150 rpm under ambient conditions (Shwab

et al. , 1999). After shaking for a specified extraction duration as shown in Table 5.2, the

mixture was allowed to settle for 24 h at room temperature (25 °C), and three layers were

observed after settling, including an extractant layer (i.e. , oil, solvent and emulsified water

mixture) on the top, an aqueous layer in the middle and the sediment at the bottom. The

middle aqueous (water) layer was removed by a Pasteur pipette, and the top extractant

155

layer was collected to a 100 mL round bottom flask for distillation in a vacuum rotary

evaporator (Yamato RE400) at a temperature of 40 °C. The solvent in this extractant was

evaporated and then condensed as the solvent recovery, while the distillate bottoms in the

flask were considered as the recovered oil (Zubaidy and Abouelnasr, 2010). The remaining

semi-solid residue in the centrifugation tube was air dried in the fume hood for 48 h to

remove any solvent, and then weighed (Mr) and compared with the original sludge mass

(Ms). The residue was also subjected to analyze its properties. The waste reduction rate

was calculated as follows:

M-M Waste reduction rate(%)= - 5 _r X 100

Ms (5.1)

The solvent recovery rate was determined as the ratio of the mass of recovered solvent

to that of original solvent used in the extraction experiment, while the oil recovery rate was

calculated as the ratio of the recovered oil mass to the original sludge mass (Zubaidy and

Abouelnasr, 2010). The quality of the recovered oil was evaluated based on its TPH

content which is determined using GC analysis. All the solvent extraction experiments

were carried out in three replicates. One-way ANOVA was used to determine the

significance (p < 0.05) of statistical difference between treatments.

Table 5.2 Experimental factors and values of solvent extraction

Factors Levels

Solvent type Cyclohexane Dichloromethane Methyl ethyl ketone

Ethyl acetate 2-Propanol

156

Solvent to sludge ratio (mL/g) 1:1 2:1 3:1 4:1 5:1 6:1 7: 1 8: 1 9:1

Extraction duration (min) 5 10 15 30 60 90

5.2.2.2. Extraction cycles

In this study, the impact of extraction cycles on recovered oil yield was investigated

because one extraction cycle might not be sufficient for oil recovery treatment (Al-Zahrani

and Putra, 2013). After the first solvent extraction from sludge, the obtained semi-solid

residue was subjected to the second extraction immediately following the extraction

procedure as described above, while such extraction was repeated for another two times.

All of the extractions were operated at the optimal solvent-to-sludge ratio and extraction

duration which were identified from the above extraction experiments. The oil recovery

rate of the ith ( ORi) extraction was calculated based on the mass of recovered oil from the

ith solvent extraction (Xi):

ORi (%) = L~Xi x 100 s

(5 .2)

5.2.3 Freeze/thaw treatment

Freeze/thaw treatment was introduced in this study to investigate its dewatering effect

on the extractant. Three groups of experiments were conducted, including freeze/thaw

alone, solvent extraction alone, and combined solvent extraction with freeze/thaw. In terms

of freeze/thaw treatment alone, a given volume of ultrapure water (Milli-Q® Advantage

Al 0) instead of solvent was added into the oily sludge, and the mixture of water and oily

sludge was then subjected to mechanical shaking for a given duration. The volume of water

157

added and the duration of shaking were determined from the above-identified optimal

solvent-to-sludge ratio and extraction duration. After that, the water and oily sludge

mixture was placed in a freezer at -20 °C for 12 h for freezing treatment (Figure 5.2), and

it was then thawed at room temperature (Zhang et al. , 2012). In terms of solvent extraction

alone, the experiment was conducted following the extraction procedure described above

using the same shaking duration as freeze/thaw experiment. In terms of the combined

solvent extraction and freeze/thaw treatment, the solvent extraction experiment was firstly

carried out at the same solvent volume addition and extraction duration, and then the

obtained extractant was frozen at -20°C for 12 h, and then thawed at room temperature.

The process of combined solvent extraction and freeze/thaw was shown in Figure 5 .3. The

recovered oil quality was compared for different groups of treatments based on its TPH

content. The impact of freeze/thaw cycles on improving the recovered oil quality was also

investigated by implementing another freeze/thaw cycle for the extractant after the first

freeze/thaw treatment, and a total of three freeze/thaw cycles were carried out. All of these

three groups of experiments were carried out in four replicates.

158

Solvent

Figure 5.2 Freeze treatment on different extractants at -20 °Cina fridge

Mechanical shaking extraction

... I

'-Water layer Solids

I I

Extractant

Freeze/thaw of extradant

' Emulsified , water

Water layer

Distillation

Figure 5.3 Flow chart of combined solvent extraction and freeze/thaw treatment

5.2.4 Sample analysis

The TPH in recovered oil was measured using GC analysis following the Canadian

Council of Ministers of the Environment (CCME) method (CCME, 2001). The distribution

159

of PH Cs fractions (F2, F3 and F4) was also analyzed, and was compared with that in the

fresh crude oil. The fractions of F2, F3 and F4 were defined as the group of petroleum

hydrocarbons from C10 to C16, C16 to C34, and C34 to Cso, respectively (Zhang et al. , 2012).

The analysis was conducted using a Varian CP-3800 Gas Chromatograph with flame

ionization detector (GC-FID): ZB-capillary column with dimension of 30 m x 0.25 mm x

0.25 µm (Phenomenex Torrance, CA); inject volume of 1 µL; injector and detector (FID)

temperature at 320 °C; carrier gas (helium) at a constant flow rate of 1.5 mL/min during

analysis. The air-dried solid residue after solvent extraction was digested using nitric acid,

and then the ICP method was used to analyze some selected heavy metal species (lead,

chromium, nickel, and zinc) (Al-Futaisi et al., 2007). The total carbon (TC) content in the

residue was analyzed using an ECS 4010 Elemental Combustion System (Costech

Instruments) at 1000 °C with a thermal conductivity detector.

5.3 Results and discussion

5.3.1 Impact of solvent-to-sludge (S/S) ratio

The solvents and solvent-to-sludge ratio had a significant influence on the oil

recovery rate. Figure 5.4a shows the oil recovery rate by different solvents at various

solvent-to-sludge ratio. It can be seen that the oil recovery rate for all solvents generally

increased with the solvent-to-sludge ratio until the equilibrium was reached, with the

lowest oil recovery rate at a ratio of 1: 1. Similar variation trend in oil recovery with

solvent-to-sludge ratio was observed using solvent extraction to treat tank bottom sludge

(Abouelnasr and Zubaidy, 2008; Zubaidy and Abouelnasr, 2010) and used lubricating oil

(Elbashir et al., 2002; Rincon et al. , 2005; Hamad et al., 2005). This is because the increase

in the ratio of solvent to sludge can enhance the mutual solubility of oil in solvent, thereby

160

increasing the amount of oil recovered (Al-Zahrani et al., 2013). The oil recovery rate of

different solvents can vary significantly. As Figure 5.5 shows, the extractant of treatments

using different solvents presented different colour. For example, the extractant using 2-

pro as extraction solvent was in light yellow, while extractant of MEK and EA were

brownish red. Since all solvents were colourless, the lighter colour the extractant was, the

less amount of PH Cs was mixed in that solvent. It was found in Figure 5.4a that 2-Pro was

associated with a much lower oil recovery than the other four solvents, while CHX was

associated with the highest oil recovery rate. The oil recovery rate of CHX, DCM, MEK,

EA, and 2-Pro reached up to 42.4%, 37.7%, 37.2%, 37.6%, and 17.1 %, respectively. A

considerable increase of oil recovery was observed for the four solvents (CHX, DCM,

MEK, EA) when the solvent-to-sludge ratio increased from 1: 1 to 4: 1, but further increase

of this ratio beyond 4: 1 resulted in very little increase of oil recovery. The results were in

accordance with the oil recovery rates of ultrasonic assisted extraction experiments. For

example, an oil recovery rate increased from 41.2% to 42.4% in the extraction using CHX

when the ratio increased from 4: 1 to 7: 1. Moreover, no significant difference of oil

recovery was observed in the extraction treatment using DCM, MEK, and EA when the

ratio reached 4: 1.

As shown in Figure 5.4b, the solvent recovery rate after oil extraction generally

increased with the solvent-to-sludge ratio. The CHX extraction treatment was associated

with a considerable increase of solvent recovery rate from 51.2% to 81.3% when the ratio

increased from 1: 1 to 4: 1, but further increase of this ratio beyond 4: 1 led to little increase

of solvent recovery. This is likely due to the fact that the solvent loss ( e.g., adhesion onto

the inner wall of distillation condenser or unavoidable evaporation) in an identical

extraction system (i.e., same containers and distillation equipment) could be similar for the

161

same solvent. As a result, the solvent recovery rate would be calculated as lower when

using a lower volume of solvent for extraction than using a higher volume of solvent under

similar solvent loss condition. It was also found that all the solvents except DCM showed

satisfactory solvent recovery rate, with the highest solvent recovery rate of only 48.5% for

DCM but > 80% for the other four solvents. This might be due to the much higher vapor

pressure of DCM (i.e. , 350 mmHg at 20 °C) than other solvents (i.e. , 73 mmHg for EA, 74

mmHg for MEK, 77.5 mmHg for CHX, and 15 mmHg for 2-Pro at 20 °C). Solvents with

high vapor pressure are prone to cause higher solvent loss because of evaporation during

the extraction process of mixing, settling, and liquid transferring (Kheireddine et al. , 2013).

In this regard, a closed solvent extraction system is desired as it is able to reduce

unnecessary solvent loss (Zubaidy and Abouelnasr, 2010).

162

50

40

~ 30 ~

~ ~ 20 0 u ~

6 10

0

* :~ l ~ 70 ~ (:! c: 60 0 ·.g 50 ::::J

] 40 Q) 30 ~

~ 20

10

0

a

___.sJ---0

1 2

1 2

C

1 2

<> CHX

O DCM

6.MEK

X EA

__co~--€Q,---'iQu---,0:,--~09---0 0 2-Pro

3 4 5 6 7 8 9 Ratio of solvent to sludge (v/m)

,J CHX !I DCM Bl MEK • EA D 2-Pro

3 4 5 6 7 8

ri CHX • DCM I:;! MEK • EA D 2-Pro

3 4 5 6 7 8 Ratio of solvent to sludge (v/m)

Figure 5.4 Effect of solvent to sludge ratio on solvent extraction performance, (a) oil

recovery rate, (b) solvent recovery rate, ( c) waste reduction rate ( experimental

163

condition: extraction duration of 120 min at 25 °C; error bar represents

standard deviation, n=3)

Figure 5.5 Extractant of extraction treatments using different solvents

In terms of waste reduction after solvent extraction, it is shown in Figure 5.4c that

more than 50% of original oily sludge mass was reduced in all the treatments. The residue

from solvent extraction treatment consists of mainly solid particles and trace amount of

water and non-extractable PHCs. It is noted that the waste reduction rate for DCM

extraction was the highest (i.e. 80%) among all the treatments, and this rate increased when

the solvent-to-sludge ratio increased from 1: 1 to 4: 1, but no further increase was observed

when the ratio was beyond 4: 1. A possible explanation for this might be that DCM has the

highest density (i.e. 1.326 g/mL) among the five solvents which might hinder the settling

of solid particles in the extractant, while a larger volume of high-density solvent

164

corresponding to a higher solvent-to-sludge ratio could retain more solid particles in the

liquid phase (Hamad et al., 2005; Kheireddine et al., 2013). As a result, the unseparated

solids associated with higher waste reduction rate could increase the content of unwanted

impurities such as fine particles, metals and asphaltene contents in the recovered oil, thus

leading to a lower oil quality. In terms of extraction by the other four solvents, the waste

reduction rate increased from 50.7% to 64.1 % (CHX), 56.3% to 61.5% (MEK), 53 .2% to

66.4% (EA), and 49.3% to 57.2% (2-Pro) as the solvent-to-sludge ratio increased from 1:1

to 4: 1, respectively. Consequently, a solvent-to-sludge ratio of 4: 1 was selected as the

optimal ratio in this study because of the associated high oil and solvent recovery and waste

reduction rates.

5.3.2 Impact of extraction duration

Extraction duration is another important factor that would affect the solvent extraction

performance because long enough duration can allow solvent to dissolve oil and enable

the impurities to aggregate to size big enough to be separated from liquid phase by settling

(Saari et al., 2007; Rincon et al., 2005). As shown in Figure 5.6a, the variation of oil

recovery rate with extraction duration depends on the solvent selected. The highest and

lowest oil recovery was observed for CHX and 2-Pro extraction, respectively, while 2-Pro

extraction achieved a much lower oil recovery than the other four solvents for all of the

extraction durations, indicating that 2-Pro was not an effective solvent for petroleum

refinery wastewater sludge extraction. Abouelnasr and Zubaidy (2008) reported similar

results that iso-propanol and iso-butanol can only recover 8% of oil from the tank bottom

sludge, suggesting that alcohol is not a suitable solvent for refinery oily sludge recycling.

Figure 5.6a also indicates that for extraction using CHX, MEK, and 2-Pro, the oil recovery

165

rate increased with the extraction duration until reaching equilibrium at the extraction

duration of 30 min. For example, the oil recovery rate of CHX extraction increased from

36.1 % to 41.5% as the extraction duration extended from 5 min to 30 min. However,

extraction duration showed little effect on oil recovery when using DCM and EA for

extraction, with a slight increase of oil recovery rate from 34.6% at 5 min to 36.6% at 30

min when using EA extraction. In terms of solvent recovery, as shown in Figure 5.6b, its

variation with extraction duration was not obvious for all the solvents although DCM

extraction was associated with the lowest solvent recovery rate (i.e. < 50%) as compared

with the other four solvents (i.e. > 80% of solvent recovery). Similarly, the extraction

duration illustrated little effect on the variation of waste reduction rate when using CHX,

MEK, EA, and 2-Pro extraction (Figure 5.6c). For DCM extraction, the waste reduction

increased from 62.6% to 77.7% as the extraction duration increased from 5 min to 30 min

when the equilibrium was reached. By considering the overall effect on oil recovery, the

mechanical shaking extraction 30 min was selected as the optimum in this study. This

duration would be long enough to achieve satisfactory miscibility between solvent and oil

in the petroleum sludge.

166

50 a

40 -::)-2-Pro e ->-CHX Q) 30 ..... ro -o-MEK .... > .... Q) 20 ~ EA > 0 Q 0 u -0 Q)

~ - CJ- DCM ....

·- 10 0

0 20 40 60 80 100

Extraction duration (min)

100

* b IIJ CHX • DCM ISi MEK • EA D 2-Pro

cf rf m rn ~ m Q) 80 ..... ro .... c:-Q) 60 > 0 u Q) .... ..... 40 C: Q) 2: 0 20 11'1

0 5 10 15 30 60 90

Extraction duration (min)

100 !/?ii CHX • DCM m MEK • EA D 2-Pro C

l <1J ..... 80 ro .... C: .Q 60 ..... u ;:J

a r:E cE a: dl :E

-u Q) .... 40 Q) ..... V'I ro s 20

0 5 10 15 30 60 90

Extraction duration (min)

167

Figure 5.6 Effect of extraction duration on solvent extraction performance, (a) oil recovery

rate, (b) solvent recovery rate, ( c) waste reduction rate ( experimental condition:

solvent-to-sludge ratio (v/m) of 4: 1 at 25 °C; error bar represents standard

deviation, n=3)

5.3.3 Extraction cycle

From the results presented above, three solvents including CHX, MEK, and EA were

selected for further investigation under the solvent-to-sludge ratio of 4: 1 and extraction

duration of 30 min. DCM and 2-Pro were excluded because of their low solvent recovery,

inadequate settling of solids, and poor oil extraction performance. As seen in Figure 5.7,

the oil recovery rate generally increased with the solvent extraction cycles. After one

solvent extraction cycle, CHX achieved a higher oil recovery (i.e., 41.0%) as compared to

that ofMEK (i.e. 35.4%) and EA (i.e. 35.8%). After the second extraction, the oil recovery

rate increased to 56.0% (CHX), 40.4% (MEK), and 38.3% (EA), respectively. Some heavy

oil components might still be attached to the solids and settled after the first extraction,

and the increase of extraction cycle could help further extract such components in the

solids although the oil recovery increment was not significant for MEK (i.e. 5.0%) and EA

(i.e. 2.5%). However, the increase of solvent extraction cycle number to three only resulted

in very slight oil recovery increment for all the three solvents (Figure 5.6), indicating that

only one or two extractions cycles could be enough for the study sludge.

168

70

es ~ 60

~ 'ii

Q) 50 +"' ro

~ ~ ... > 40 ... Q) ~ > 0 30 u Q) ... 0 20

<> CHX 6.MEK X EA 10

0 1 2 3 4

Cycles of extraction

Figure 5. 7 Effect of solvent extraction cycles on oil recovery rate

5.3.4 Freeze/thaw enhancement

The solvent-to-sludge or water-to-sludge ratio of 4: 1 and mechanical shaking duration

of 30 min were used to conduct the three groups of experiments to examine the effect of

freeze/thaw treatment. In this study, no noticeable oil separation was observed when using

freeze/thaw (FIT) treatment alone (experimental data not shown), indicating that FIT

treatment alone was not effective for oil recovery from refinery wastewater pond sludge.

The possible explanation is that in the original sludge, PH Cs were strongly attached to the

solids, which cannot be effectively separated by freeze/thaw alone. Figure 5.8 presents the

effect of solvent extraction alone and the combined solvent extraction with FIT treatment

on the oil recovery performance. As shown in Figure 5.8a, after solvent extraction

treatment, the extractant contained a large amount of emulsified water, which cannot be

effectively separated via gravitational settling. During freezing, these emulsified water

droplets formed ice lattice in the extractant (shown as the shadow area in the extractant),

leaving the mixture of solvent and oil unfrozen in the centrifugation tube. After thawing at

169

room temperature, it was observed that the height of separated water layer in the tube with

FIT treatment was generally higher than that without FIT treatment (Figure 5.8b-d), which

indicates an effective dewatering effect of FIT was taken place on extractant. Moreover,

the discrepancy in the height of the separated water between FIT treatment and non FIT

treatment was more obvious for MEK and EA as compared with CHX extraction,

suggesting that the FIT enhancement was more effective in dewatering of extractant from

treatments using EA and MEK. As a result, more emulsified water can be removed from

the extractant by using FIT for these two solvents, leading to reduced water content in the

recovered oil.

170

Figure 5.8 Effect of freeze/thaw enhanced solvent extraction and the comparison of

separated water level from the extractant with freeze/thaw treatment (left

yellow bar) and without freeze/thaw treatment (right red bar), (a) ice lattice

171

formation in the extractant from MEK extraction, (b) CHX extraction, ( c)

MEK extraction, and ( d) EA extraction

As shown in Figure 5.9a, the TPH contents in the recovered oil were generally lower

for solvent extraction alone in terms of all the three solvents. However, the combined

solvent extraction with freeze/thaw treatment can significantly (p < 0.05) increase the TPH

content in the recovered oil for both MEK and EA extractions. For example, with FIT

treatment, the TPH content in the recovered oil increased from 41.8% to 62.6% when using

MEK extraction and from 41.0% to 58.8% when using EA extraction. The improvement

of FIT treatment on TPH content in the recovered oil by CHX extraction was not significant

(p > 0.05). The mechanism for enhanced dewatering by freeze/thaw is the coalescence of

the adjacent water droplets in the W/0 emulsion (Ghosh and Coupland, 2008). During

freezing, the dimension of emulsified water droplets grows as the temperature decreases

and the interfacial film between water and oil becomes thinner. When the temperature

drops below the melting point of water, the ice crystal is formed and it could then break

the interfacial film to form bridge between the adjacent water droplets. During the thawing

process, these connected adjacent water droplets allows coalescence to happen, and thus

an even larger droplet can be formed which can then be readily separated by gravitational

settling (Yang et al., 2009; Lin et al., 2008; Lin et al., 2007). The insignificant dewatering

effect of freeze/thaw on the extractant from CHX extraction might be due to the higher

melting point of this solvent (i.e. 6.5°C) than that of water. The CHX could be frozen

before the emulsified water in the extractant, and this frozen continuous solvent phase

might act as a barrier that hinders the connection of ice crystals from frozen water droplets,

and thus leading to difficulty in the coalescence of water droplets. Lin et al. (2008) also

172

reported that the dewatering effect of freeze/thaw on the emulsion derived from CHX

extraction was not effective until the freezing temperature dropped below -45°C. Figure

5.9b illustrates the effect of FIT cycles, and it was found that the increase of TPH in the

recovered oil by increased number of FIT cycles were not significant (p>0.05) for each

solvent, showing that one FIT cycle would be sufficient for the dewatering of extractant

obtained from the high-moisture petroleum sludge. Ghosh and Rousseau (2009) reported

that approximately 35% of the total water in the emulsion derived from vegetable oil can

be separated after 20 FIT cycles, but the first three FIT cycles showed little effect on water

separation.

173

70 a 60

BEi Without F/T '*- 5o • With F/T ..... @ 40 ..... C:

8 30 :::c ~ 20

10

0

80

70

'eR 60

~ 50 QJ

~ 40 0 u

:::c 30 0.. 1- 20

CHX

b

CHX

MEK EA

@ Cycle 1 • Cycle 2 1§.'1 Cycle 3

MEK

Solvents EA

Figure 5.9 Effect of freeze/thaw (FIT) treatment on (a) TPH content in recovered oil by

solvent extraction and (b) TPH content in recovered oil as a function of FIT

cycle

5.3.5 Characteristics ofrecovered oil and solid residue

As shown in Figure 5.10, the F2, F3, and F4 fractions accounted for 26.5%, 69.1 %,

and 4.3% of TPH in the recovered oil by CHX extraction with freeze/thaw. On the other

hand, the F2 and F3 fractions in the oil recovered by the combination of MEK or EA

extraction with freeze/thaw treatment were slightly higher, while the F4 fraction was lower

than that from CHX extraction. Most volatile PHCs such as the Fl fraction (i.e. C6-C10)

174

can be evaporated during oily sludge storage, leaving the relatively heavy fractions such

as F2 to F4 in oily sludge available for recovery (Heidarzadeh et al. , 2010). Consequently,

the recovered oil normally contains a relatively large amount ofF2 and F3 fractions which

account for >80% ofTPH in the recovered oil (Zhang et al., 2012). In comparison to crude

oil, the recovered oil from the study sludge contained more F3 fraction but less F2 fraction

(Figure 5.10), and it could be suitable for use as a feedstock for heavy fuel oil production

(Smith et al., 2003).

80

70

60 ~ Q) 50 tl.D

~ 40 s:: Q) u <ii 30 Q.

20

10

0 F2 F3

PHCs fractions

~ CHX

• MEK

~ EA

D Fresh crude

F4

Figure 5.10 Distribution of PH Cs fractions in fresh crude oil and the recovered oil by using

the combination of solvent extraction with freeze/thaw

Solid residue is the by-product of oil recovery treatment and it requires proper

treatment before disposal of to avoid causing any secondary pollution. As Figure 5.11

shows, the solid residue were mainly consisted of two layers: a thin non-extractable PHCs

(i.e., asphaltenes and resins) layer on the top and soil particle layer at the bottom. As shown

in Table 5.3, after treatment using the combination of solvent extraction with freeze/thaw,

the TPH content in the solid residue was significantly lower than that in the original oily

175

sludge, with more than half of the TPH in oily sludge being recovered by solvent extraction.

The total carbon (TC) content also declined, while the TPH and TC contents in the solid

residue decreased with the increase of solvent extraction cycles. It was observed that

residues after CHX extraction contained the highest TPH and TC contents as compared

with those after MEK and EA extractions. Although solvent extraction showed obvious

effect on oil recovery from the oily sludge, the TPH content in the solid residue was still

much higher than the clean-up level (i.e. < 10,000 mg/kg of TPH in the soil at industrial

area) recommended by Canadian Council of Ministers of the Environment (CCME, 2008),

indicating that the residue still requires proper management prior to final disposal. The

variation of heavy metal concentrations in the solid residue showed an opposite direction,

while such concentration generally increased with the solvent extraction cycles. The results

revealed that solvent extraction with freeze/thaw treatment had little removal effect on

heavy metals in oily sludge, and most heavy metals were remained and accumulated in the

solid residues. The heavy metal concentration in the solid residue generally increased due

to the mass reduction of sludge after extraction treatment. Although these concentrations

( except Ni) were lower than their regulatory limits in soils set by US EPA (i.e. 7500 mg/kg

for Zn, 420 mg/kg for Pb, 75 mg/kg for Ni, and 3000 mg/kg for Cr) (EPA, 1993), the

increased heavy metal contents in the solid residues from oil recovery treatment by solvent

extraction with freeze/thaw could still be a potential environmental problem which requires

proper management.

176

Figure 5.11 Solid residues from refinery wastewater pond sludge after combined solvent

extraction and freeze/thaw treatment

177

Table 5.3 Properties of the original oily sludge and the solid residues after combined solvent extraction and freeze/thaw treatment

Treatments Extraction TPH (mg/kg) Tea (%) Cr (mg/kg) Ni (mg/kg) Pb (mg/kg) Zn (mg/kg)

cycles

Original sludge 198866.7± 5559.lb 39.5 ± 2.8 70.0 ± 4.9 58.5 ± 3.5 64.7 ± 2.2 2040.0 ± 36.5

CHX extraction I 96400.0 ± 5406.5 26.2± 1.1 144.7 ± 4.6 80.9 ± 3.0 173.5 ± 5.8 2494.4 ± 34.7

2 86566.7 ± 2768.3 20.2 ± 1.0 165.6 ± 2.5 85 .2±3.1 179.3 ± 5.5 2697.3 ± 86.3

3 75333 .3 ±1450.3 16.2 ± 1.2 172.1 ± 4.2 89.6 ± 1.8 188.7 ± 5.0 2829.3 ± 60.7

MEK extraction 1 71933.3 ± 2205.3 15.8 ± 0.2 179.1±6.1 91.9 ± 1.0 197.5± 1.8 2924.7 ± 47.6

2 69033.3 ± 1965.5 14.2 ± 0.1 178.0 ± 3.3 92.7 ± 0.7 202.3 ± 3.0 2922.8 ± 47.6

3 64400.0 ± 1852.0 12.9 ± 1.6 180.8± 1.7 93.0 ± 0.9 204.8 ± 3.6 2967.0 ± 60.6

EA extraction 1 73233.3 ± 4278.2 13.4 ± 0.9 182.1 ± 6.6 92.1 ± 1.2 196.4 ± 6.6 2877.3 ± 46.9

2 66633.3 ± 1692.1 12.0 ± 0.3 182.4 ± 5.9 95.8 ± 2.3 206.3 ± 7.4 2991.6 ± 41.4

3 62233 .3 ± 1616.6 11.6 ± 0.4 184.4 ± 7.3 96.0 ± 1.3 203.4 ± 2.2 2969.7 ± 44.3

a TC: Total carbon content in dry oily sludge and residues ; b standard deviation, n=3

178

5.4 Conclusions

Oil recovery from the high-moisture oily sludge was investigated using solvent

extraction and freeze/thaw treatment. Three solvents including CHX, MEK and EA were

identified as ideal candidates based on their oil recovery and solvent recovery

performances. They can achieve about 40% of oil recovery after 30 min of shaking

extraction at a solvent-to-sludge ratio of 4: 1, while more than 80% of solvent can be

recycled. However, the TPH content in the recovered oil was relatively low (i.e. about 30%

by CHX and 40% by MEK or EA extraction) because of the presence of emulsified water.

It was found that freeze/thaw treatment can significantly increase the TPH content in the

recovered oil from MEK or EA extraction (i.e. from 40% to 60%) by dewatering of the

extractant, but little effect was observed for the combination of CHX extraction with

freeze/thaw. The recovered oil mainly consisted ofF2 and F3 fractions of PH Cs. The TPH

and total carbon contents in the solid residue after solvent extraction with freeze/thaw

treatment can be significantly reduced, but the heavy metal concentrations increased. In

summary, the combination of solvent extraction with freeze/thaw process could represent

an effective and feasible approach for the management of the hazardous high-moisture oily

waste in terms of its satisfactory oil recovery rate and the increased TPH content in the

recovered oil.

179

Chapter 6 Treatment of refinery oily sludge through co-pyrolysis with wood waste

6.1 Introduction

The petroleum industry generates a considerable amount of oily sludge during various

processes such as crude oil exploration, transportation, storage, and refining. An annual

production and total accumulation of oily sludge in the world was estimated to be over 60

million and 1 billion tons, respectively (da Silva et al. , 2012; Hu et al. , 2013). This type of

sludge is a complex and stable emulsion of various petroleum hydrocarbons (PH Cs), water,

solid particles, and metals (Hu et al., 2015). It has been classified as a hazardous waste in

many countries since it may contain various carcinogenic polycyclic aromatic

hydrocarbons (P AHs) such as benzene, phenol, anthracene, and pyrene (Roldan et al. , 2012;

Naik et al. , 2011). Due to its hazardous nature and increasing production quantity, the

effective treatment of oily sludge has become a worldwide pressing need.

Traditional methods for oily sludge treatment such as landfilling have been

challenged by increasingly stringent regulations because of their low efficiency, long

treatment time, and possible secondary pollution (Chirwa et al., 2013). In recent years, the

recycling of energy from such oil-rich waste while minimizing its adverse environmental

impact has received increasing interests, and a variety of physical, chemical, and biological

methods as well as their combinations have been developed (Hu et al. , 2013). Among these

methods, pyrolysis represents an effective thermo-chemical conversion process during

which oily waste is heated in a closed oxygen-free reactor system at moderate operating

temperature (i.e. , usually 200 to 500 °C) (Isahak et al. , 2012; Bridgwater, 2012). This

180

process can convert organic matters into combustible gases, pyrolysis oil, and char. The

combustible gas and pyrolysis oil can be used for energy resource and the char after proper

modification can be used as an adsorbent for pollution control (Bridgwater, 2012; Zhang

et al. , 2013). A few studies reported that pyrolysis can be effective for oily sludge treatment.

For example, Karayidirim et al. (2006) observed that 30.0% of pyrolysis oil and 15.0% of

gas (i.e. , hydrogen, methane, and carbon monoxide) can be produced from the pyrolysis

of oily sludge collected from the primary decanting step following API separation in the

wastewater treatment plant of a refinery. Liu et al. (2009) found that around 80.0% of total

organic carbon can be converted into hydrocarbons gas through pyrolysis when heating

the tank bottom oily sludge sample to a temperature range of 473-723 K. Molt6 et al. (2013)

examined the pyrolysis of two different sludge wastes (i.e. , petrochemical sludge and

biological sludge) and found that the gaseous products can be significantly affected by the

heating rate, oxygen content, contact time, and the nature of sludge. Conesa et al. (2014)

investigated the pyrolysis of petrochemical sludge and observed an increased yield of

liquid oil (i.e. , from 30 to 50%) due to increasing treatment temperature (i .e ., from 350 to

530 °C).

In addition to the application to oily sludge waste treatment, many pyrolysis studies

have focused on pyrolysis oil production from biomasses (i .e. , mainly lignocellulosic

plants) because they are considered as abundant and promising renewable energy sources

(Pootakham and Kumar, 201 O; Butler et al. , 2011 ). Biomass derived pyrolysis oil ( e.g. bio-

oils) possesses several environmental advantages over fossil fuels such as less undesirable

gaseous emission of CO2, SOx, and NOx (Isahak et al., 2012). A variety of biomasses such

as wood and agricultural residues have been used as feedstock for producing bio-oils via

pyrolysis (Ki et al., 2013; Asadullah et al. , 2013; Abnisa and Daud, 2014). In general, the

181

bio-oil yield and properties can be significantly affected by pyrolysis temperature and the

nature offeedstock (Chen et al., 2013). Although showing a great potential as an alternative

energy source, the utilization of bio-oil has been limited due to some drawbacks such as

high oxygen and water content, low heating value, and instability (Isahak et al., 2012). It

is thus of importance to find effective solution for improving the biomass pyrolysis. The

co-pyrolysis of biomass with other organic wastes seems to be a simple and effective way

for such improvement not only in pyrolysis oil quality but also in yield (Abnisa and Daud,

2014). Biomass and organic wastes usually have different chemical and physical properties

such as moisture, volatile matter, ash content, calorific value, porosity, and

oxygen/hydrogen/carbon (0/H/C) molar ratio. The differing properties can change the

reactivity and thermal characteristics of samples and products, and the formation of

synergistic interaction during co-pyrolysis could then result in an improved pyrolysis

product (Kar, 2011 ).

Many efforts have been made to investigate the co-pyrolysis of biomass with other

wastes, such as waste tires (Martinez et al., 2013), plastic wastes (Pinto et al., 2013), and

municipal sewage sludge (Samanya et al. , 2012). For example, Cao et al. (2009) found that

the liquid oil product from the co-pyrolysis of sawdust with waste tires had a higher heating

value (HHV) of 42.44 MJ/kg, which is much higher than that (28 .51 MJ/kg) from the

pyrolysis of sawdust alone. Brebu et al. (2010) examined the co-pyrolysis of pine cone

with synthetic polymers and found that the oil yield increased from 47.5 to 69.7 wt.%.

Onal et al. (2014) investigated oil production by co-pyrolysis of almond shell with high

density polyethylene, and observed that the produced oil via co-pyrolysis had an improved

heating value (i.e., by 38%) but a much decreased oxygen content (i.e., by 86%) than those

from biomass pyrolysis alone. Zuo et al. (2014) also observed that the HHV of oil product

182

can be improved by the co-pyrolysis of sewage sludge with poplar sawdust as compared

to pyrolysis of sewage sludge alone. However, so far there has been limited research

investigating the co-pyrolysis of biomass with refinery oily sludge. As a result, the present

study aims to examine the treatment of refinery oily sludge waste using co-pyrolysis with

a wood waste (saw dust). The synergistic effect of oily sludge addition to sawdust on

pyrolysis oil and char products was investigated. The co-pyrolysis was carried out in a

fixed-bed reactor under different pyrolysis conditions. A number of experimental factors

were examined, including sawdust percentage in the feedstock mixture, temperature, and

heating rate. The main and interaction effects of these factors on the yield of pyrolysis oil

and the solid char were evaluated through a response surface methodology (RSM). The

quality of pyrolysis product was also analyzed. The results would provide valuable

information for developing an effective oily sludge waste management approach in terms

of both resource recycling and minimizing adverse environmental effects.

6.2 Experiment

6.2.1 Materials

Wood waste (Douglas fir sawdust) was collected from a wood industry, and oily

sludge was collected from tank bottom periodical cleaning at an oil refinery in western

Canada (Figure 6.1 ). The sludge and sawdust samples were oven dried at 80 and 105°C

for 8 h to remove moisture, respectively (Liu et al., 2013; Tian et al., 2014). The sawdust

was ground in a high-speed rotary cutting mill and then screened into particles of about 1

mm in diameter. Table 6.1 lists the sample properties and their test methods.

183

Figure 6.1 The feedstock of sawdust and oily sludge (left: before mixed; right: after mixed)

Table 6.1 Properties of sawdust and oily sludge

Analysis Analytical Methods Value

Sawdust (raw) Oily sludge (raw)

Proximate analysis:

Ash content (at 750°C) (wt.%) ASTMD3174 1.0 10.3

Moisture content (wt.%) ASTME871 6.2 22.4

Volatile matter (wt.%) ASTME872 76.5 56.5

Fixed carbon a (wt.%) 16.3 10.8

Ultimate analysis:

Carbon, C (wt.%) ASTMD5373 47.0 60.5

184

Hydrogen, H (wt.%) ASTMD5373 5.8 17.6

Oxygen 3, 0 (wt.%) ASTMD5373 47.2 21.8

Nitrogen, N (wt.%) ASTMD5373 0.01 0.05

Higher heating value (MJ/kg) ASTM D5865/E711 16.5 43.5

a Calculated by mass difference.

6.2.2 Experimental design

When several independent factors and their interaction have influence on the

experimental outcome, the response surface methodology (RSM) is an effective tool for

the optimization of these factors with a minimum number of experimental runs (Ahmadi

et al., 2005). A number of factors such as sawdust percentage in the feedstock, pyrolysis

temperature, and heating rate can affect the pyrolysis products (Abnisa and Daud, 2014).

As a result, these three factors were examined in this study. By using Design Expert® 7.0,

a five level-three variable central composite circumscribed (CCC) experimental design

method was applied for arranging the co-pyrolysis experiments. The CCC had a factorial

design and star points at a distance of ±1.682 from the central point. Table 6.2 lists the

experimental levels (coded as ±1, 0, and ±1.682) and their real values. By using the CCC

design method, a total of 17 experiments (8 factorial points, 3 center points, and 6 star

points) were required. Each experiment arranged at the center of experimental domain (i.e.,

run# 15-17) was repeated for three times in order to estimate the pure error (Katsoura et

al., 2007). The analysis of variance (ANOV A) was used for analyzing the experimental

data. Multiple linear regression analysis was performed to fit a quadratic poly-nominal

model as follows:

185

3 3 2 3

Y = Po + L Pixi + L Puxf + LL PijxixJ Eq. (6.1) i=l i=l i=l J>i

Where y is the response variable (i.e. pyrolysis oil or char yield); x; or XJ is the

independent variable (i.e. the three experimental factors); Po, p;, Pu, and PiJ are the intercept,

linear, quadratic, and interaction coefficients of the model, respectively. RSM was used to

illustrate the main and interaction effects of experimental factors . The optimal co-pyrolysis

condition for maximum oil yield was identified using the numerical optimization function

of Design Expert® 7.0 software (Katsoura et al., 2007).

6.2.3 Pyrolysis procedure

Each pyrolysis experiment was carried out in a fixed-bed tube furnace reactor (quartz

tube length: 600 mm, 0 : 50 mm; MTI Corp.® GSL-1 IOOX) under atmospheric pressure.

(Figure 6.2). The schematic experimental set-up was shown in Figure 6.3. About 20 g of

feedstock (i.e. , sawdust and oily sludge mixture) was put into the sample ark which was

placed in the center section of the tube reactor. The sweeping nitrogen gas was passed

through the tube reactor at a constant flow rate of 100 mL/min. Samples were heated to

the set temperature at a given heating rate as listed in Table 6.2. The hot vapors from

pyrolysis were condensed in three ice-water cooling traps (Chen et al. , 2013). After

pyrolysis, the liquid oil product was collected from the cooling traps and weighed. The

pyrolysis oil yield was then calculated by dividing the mass of collected oil to the mass of

feedstock. Char yield was measured by dividing the mass of char remaining after pyrolysis

to the mass of original feedstock, while the gas yield was determined by overall mass

186

balance (Zuo et al., 2014). The theoretical yield of co-pyrolysis of sawdust with oily sludge

was calculated as follows:

Y product = a X Y product I + B X Y product 2 Eq. (6.2)

Where Y product is the theoretical yield of co-pyrolysis product, Y product I is the yield

from sawdust during co-pyrolysis (%), Y product 2 is the yield from oily sludge during co-

pyrolysis (%), a and B represents the mass percentage (%) of sawdust and oily sludge in

the feedstock mixture, respectively (Song et al., 2014).

Figure 6.2 The fixed-bed reactor designed for the co-pyrolysis of sawdust with oily sludge

187

Thermal couple

Tube furnace reactor

Gases to fume hood

Flow meter

0 0 Cooling trap

Temperature control

Figure 6.3 Schematic diagram of fixed-bed pyrolysis reactor

6.2.4 Sample analysis

Some physical properties of pyrolysis oil were determined by using the following

standard test methods: the specific gravity (ASTM D4052), the higher heating value

(ASTM D240), and water content (ASTM E203). The thermo-gravimetric analysis

(TGA) and differential thermal-gravimetric (DTA) analysis were performed using a TA

Instruments® Q500 thermo-gravimetric analyzer to measure the thermal behavior of

sawdust and oily sludge from room temperature to 1000°C. The analysis was carried out

at 10 mg sample, 100 mL/min nitrogen gas flow with 20°C/min heating rate. The

ultimate analyses of feedstock, pyrolysis oil and char were carried out by a Costech®

ECS 4010 elemental analyzer according to ASTM D5291 method.

188

6.3 Results and discussion

6.3.1 Thermal gravimetric analysis (TGA)

As shown in Figure 6.4a, the decomposition of sawdust can be divided into three

temperature stages: ambient to 110, 110 to 700, and 700 to 900 cc. The first stage was the

result of the loss of moisture and light volatile components. The second stage reveals the

de-volatilization of major organic compounds in sawdust such as cellulose, hemicellulose,

and lignin. It has been reported that hernicellulose and cellulose started to break and release

volatiles in the temperature range of 110 to 400 cc, and the decomposition rate was slow

beyond 400 cc due to the degradation oflignin (Aqsha et al. , 2011). All the volatiles were

evolved at 700 cc, and only the char remained. The horizontal portion of the

thermogravimetric (TG) curve (Figure 6.4a) in the third stage can be attributed to

carbonization (Zuo et al. , 2014, Song et al. , 2014). According to the DTG curve (Figure

6.4a), the maximum weight loss of sawdust was observed at 350 cc. As shown in Figure

6.4b, the thermo-degradation behavior of oily sludge can be divided into four stages: an

initial decrease in the weight of oily sludge from ambient temperature to 110 cc due to

dewatering, a rapid weight loss (i.e. , 43 .2% of weight loss) stage in which the light volatile

petroleum hydrocarbons (PHCs) are volatilized in the temperature range of 110-150 cc, a

third stage (i.e. 42.4% of weight loss) in which the decarboxylation and de-polymerization

of complex PH Cs may occur in the temperature range of 150-550 cc, and the last stage at

temperature above 550 cc during which decomposition of a small fraction of more

complex organic structures and inorganic materials takes place (Wang et al., 2007; Liu et

al. , 2009). The maximum weight loss rate of oily sludge was found at 150 cc (Figure 6.4b)

due to the rapid evolution of small petroleum hydrocarbons. Sawdust and oily sludge both

showed significant weight loss in the temperature range of 110-600 cc, indicating that the

189

co-pyrolysis vapors could possibly have some interaction effects on the quality of pyrolysis

products.

l .E C,

~

~ 0

.E

.S!'

~

100~--,-------------------------,-----------.2.s

80

60

Cellulose Moisture

40

I r, r, 11 ,, 11 IJ r, 11 , , I 1 :--+.___

I I I I I I I I I I

Lignin

a: Sawdust Weight(%) De,ivative (%/'C)

carbonization

2.0

1.5

1.0

0.5

- --- ---------------- - 0.0

04-------~~-~-~----~-~~-~-~----~-~---+~.5 0 200 400 600 800 1000

Temperature (°Cl 100,--"""",---------------- ---------,---------,- 1,2

Light PHCs ~ ,, 11 11

b: Oily Sludge Weight (%) Derivative (%rC)

1.0

80 I Moisture

1:

60

40

20

-- , 1

, I I ,r,_/

+ 1 I I

I -1 Complex PHCs I I I I I I I I I I I I J .,.---- ........ .... ... -, I I l'o.--- -,

0.8 -- Carbonization & Inorganic Materials 0.6

0.4

0.2

\ \

' ------ - - ---------------------- 0~

0-t-----~-.,---~---.,.----~-~--,,--~------~-~---+-~.2 0 200 400 600 800 1000

Temperature (°C)

f ~ .E .S!'

~ > -~ 0

Figure 6.4 TG and DTG curves of (a) sawdust and (b) oily sludge (solid line: TG; dashed line: DTG)

190

6.3 .2 Pyrolysis oil yield

The co-pyrolysis showed a varying product yield and quality. Figure 6.5 presents the

appearance of pyrolysis oil using different feedstock, including saw dust alone, oily sludge

alone, and sawdust/oily sludge mixture. There was only one oil layer observed in the

pyrolysis of sawdust alone and oily sludge alone. However, the oil from co-pyrolysis

presented two distinct layers, with the top layer being dark brown and sticky, and the

bottom layer being translucent red-brown. The top oil layer showed similar appearance

with the oil from the pyrolysis of oily sludge alone, while the bottom layer was more likely

originated from the pyrolysis of sawdust. This observation is in agreement with previous

co-pyrolysis studies of sewage sludge with biomass (Wu et al., 2014; Samanya et al., 2012).

The formation of two distinct oil phases during co-pyrolysis might be due to the fact that

the oil derived from saw dust pyrolysis is composed of a complex mixture of oxygenated

compounds which is not mixable with other hydrocarbon liquids (Bridgewater, 2012).

Table 6.2 2 lists the yields of products (pyrolysis oil, char and gas) obtained from the co-

pyrolysis under different experimental runs. The oil yield from co-pyrolysis was obtained

by the sum of the two oil phases. It is found that the oil yield from the pyrolysis of sawdust

alone was 46.5% (i.e. run #10), which is consistent with the reported yield rates of sawdust

pyrolysis in many literatures (Fei et al., 2012; Liu et al., 2013). Less liquid oil yield (i.e.,

35.3%) was obtained from the pyrolysis of oily sludge alone (i.e. run# 9), and this was

caused by the lower volatile matter content in oily sludge than in sawdust (Table 6.1 ). The

highest oil yield rate was observed as 48 .8% from the co-pyrolysis with a sawdust

191

percentage of 80% in the feedstock (i.e., run #8), indicating a possible synergetic effect

during co-pyrolysis.

Figure 6.5 Liquid oil product from the pyrolysis of sawdust alone (left), mixture of

sawdust (75 wt.%) and oily sludge (25 wt.%), and oily sludge alone (right)

(Heating rate: 20 °C/min; Temperature: 500 °C)

Response surface methodology was employed for modeling the effect of experimental

factors (sawdust percentage A, temperature B, and heating rate C) and examining the

optimal experimental condition for pyrolysis oil yield. The response variable (i.e.,

pyrolysis oil yield) was assessed as a function of these three factors and calculated as the

sum of a constant, three first-order effects (terms in A, B, and C), three interaction effects

(terms in AB, AC, and BC), and three second-order effects (A2, B2, and C2) according to

Eq. (6.1). Only the terms found statistically significant were included in the model. The

interaction effect of AB and second-order effect of C2 were found non-significant by the

F-test at the 5% confidence level, and thus they were dropped from the model through the

"backward" process. The modified regression model for pyrolysis oil yield is:

192

Runs

1

2

3

4

5

Y0 = -102.85 + 0.104A + 0.498 + 2.0SC + O.OlAC - 5.09 x 10-3 BC - 1.46 x

Eq. (6.3)

Where Y0 is the yield of pyrolysis oil (%), A represents the mass percentage of

sawdust in feedstock (%), B is the pyrolysis temperature (0 C), and C is the heating rate

(°C/min).

Table 6.3 presents the statistical parameters obtained from ANOV A for the developed

regression model. The model was significant based on F-test at 5% confidence level (P <

0.05). The adjusted R2 coefficient was well within the acceptable range of R2 > 0.9,

indicating a satisfactory adjustment of the quadratic regression model. The lack of fit was

not significant (P > 0.05), indicating that the fitness of model was robust. High predicted

R2 value means the developed response surface model can be used as a predictive tool over

the whole parameter uncertainty range. The optimal oil yield condition was set as heating

20 g sawdust to 500 °C at a heating rate of 20 °C/min.

Table 6.2 Experimental array of CCC design for co-pyrolysis experiments and product

yield results

Sawdust ratio in Temperature a Heating rate a Pyrolysis oil Char yield Gas yield

feedstock a (wt.%) (QC) (°C/ min) yield(%) (%) (%)

(-1) 20 (-1) 440 (-1) 8 35.8 19.7 44.5

(1) 80 (-1) 440 (-1) 8 40.3 24.2 35.5

(-1) 20 (1) 560 (-1) 8 41.4 15.2 43.4

(1) 80 (1) 560 (-1) 8 46.2 17.6 36.2

(-1) 20 (-1) 440 (1) 17 36.6 19.8 43.6

193

6 (1) 80 (-1) 440 (1) 17 47.7 20.1 32.2

7 (-1) 20 (1) 560 (1) 17 36.2 13.4 50.4

8 (1) 80 (1) 560 (1) 17 48.8 13.4 37.8

9 (-1.682) 0 (0) 500 (0) 12.5 35.3 13.9 50.8

10 (1.682) 100 (0) 500 (0) 12.5 46.5 16.5 37.0

11 (0) 50 (-1.682) 400 (0) 12.5 39.3 26.8 33.9

12 (0) 50 (1.682) 600 (0) 12.5 41.8 13.4 44.8

13 (0) 50 (0) 500 (-1.682) 5 42.5 17.5 40.0

14 (0) 50 (0) 500 (1.682) 20 45.7 13.5 40.8

15 (0) 50 (0) 500 (0) 12.5 45.1 14.6 40.3

16 (0) 50 (0) 500 (0) 12.5 44.5 14.4 41.1

17 (0) 50 (0) 500 (0) 12.5 44.8 14.9 40.3

• Coded levels (in parentheses) and real values of experimental factors

Table 6.3 Statistical parameters obtained from the analysis of variance for the model

Variable (bio oil yield) Value Variable ( char yield) Value

R2 0.98 R2 0.96

R2 adjusted 0.97 R2 adjusted 0.94

R2 predicted 0.91 R2 predicted 0.86

F exp a 482.4 F exp a 1874.6

P > F < 0.001 P > F < 0.001

Lack of fit (P-value) 0.1 > 0.05 Lack of fit (P-value) 0.06 > 0.05

Std. Dev. 0.8 Std. Dev. 1.0

Coefficient of variance(%) 1.9 Coefficient of variance(%) 5.8

194

Press

Adequate precision

28.1

25.1

Press

Adequate precision

• F exp is the ratio of the mean square of the model to mean square of the error.

35.4

23 .8

The simulated main effects of the three experimental factors on pyrolysis oil yield are

illustrated in Figure 6.6. It can be found that the increase of sawdust percentage in

feedstock could have a positive effect on the yield of pyrolysis oil. Generally, a higher

sawdust percentage led to a higher pyrolysis oil yield, but further increase of this

percentage only brought limited increase in oil yield. In fact, the highest oil yield rate of

48.8% during experiments was associated with a sawdust percentage of 80% (Table 6.2).

The temperature illustrated a differing effect from sawdust percentage. As shown in Figure

6.6b, both too high and too low pyrolysis temperature showed a negative effect on oil yield,

with the optimal temperature being in the range of 500-550 °C, while the oil yield declined

by about 3% as the temperature increased from 500 to 600 °C. This might be because of

the fact that a small part of organic matter in sawdust and oily sludge cannot completely

decompose at a temperature lower than 400 °C (Biagini et al. , 2006; Akhtar and Amin,

2012), while a higher temperature (> 600 °C) could result in secondary decomposition

during which the liquid oil is further decomposed into non-condensable gaseous products

(Shen and Gu, 2009). A number of literatures also reported that the maximum liquid oil

yield from biomass pyrolysis often occurred in the temperature range of 500-550 °C

(Akhtar and Amin, 2012). The yield of bio-oil declined about 3% as the temperature

increased from 500 to 600 °C. High temperature (> 600 °C) could result in secondary

decomposition, during which the liquid oil is further decomposed into non-condensable

gas products (Shen and Gu, 2009). The effect of heating rate was positive on the yield of

195

pyrolysis oil. As shown in Figure 6.6c, a higher heating rate was preferred for producing

pyrolysis oil, with oil yield being increased from 41.1 % to 45. 7% as heating rate increased

from 5 to 20 °C/ min. The positive effect of high heating rate is mainly due to the short

time available for secondary decomposition of condensable gases (Akhtar and Amin,

2012). It was found from statistical analysis that all three experimental factors played a

major role in the process of pyrolysis oil production.

The interaction effect of experimental factors on pyrolysis oil yield was shown in

Figure 6.7. The interaction of sawdust percentage and pyrolysis temperature was not

included because this effect was not significant according to statistical analysis. As shown

in Figure 6. 7a, low heating rate and low pyrolysis temperature could result in low pyrolysis

oil yield. Increasing heating rate from 5 to 20 °C/min could enhance the oil yield in

treatments with a temperature lower than 500 °C. However, when the treatment

temperature exceeded 550 °C, high heating rate (i.e. 20 °C/min) could pose a negative

effect on pyrolysis oil production. The interaction effect of sawdust percentage and heating

rate was illustrated in Figure 6.7b. It was found that high heating rate was preferred in the

treatments with a sawdust percentage higher than 50%; however, high heating rate could

compromise the oil yield in the treatments which had a low sawdust percentage (i.e., <

25%). This might be because at high heating rate, light PHCs fractions in oily sludge were

quickly decomposed into organic vapor which cannot be involved in polymerization

reactions to form condensable gaseous products, and thus leading to a low yield of oil

products.

196

55 a -;§!. 0

:2 49 ·~ 'i5

"' 43 'iii .?:-e >, a.. 36

30

0 25 50 75 100 Sawdust percentage (wt.%)

54 b -~ ~ 49 "O al ·;;:. ·o 43 "' 'iii >,

~ 36 6:

30

400 450 500 550 600 Temperature (°C}

55 C -~

'C 49 a:; ·;;;. ·o "' 43 'iii >, e >, 0.. 36

30

5 9 13 16 20 Heating rate (°C/min)

Figure 6.6 The main effects of single factors on pyrolysis oil yield: (a) sawdust

percentage (temperature: 500 °C, heating rate: 12.5 °C/min), (b) temperature

(sawdust percentage: 50 wt.%, heating rate: 12.5 °C/min), and (c) heating rate

(sawdust percentage: 50 wt.%, temperature: 500 °C)

197

60 a -~ 0 _.. 53 "O (I.) ·;;, 45 ·5

"' 38 'ci> >-0 30 ,._ >-a..

600

Heating rate (°C/rnin ) 5 400 Temperature (°C)

60 ........ ~ 0 _..

53 "O (I.) ·;;, 45 ·o "' 38 'ci, >-e >, a.

Heating rate (°C/rn in) 5 0 Sawdust (wt.%)

Figure 6. 7 The interaction effect of experimental factors on pyrolysis oil yield

6.3 .3 Pyrolysis char yield

Solid char is another valuable product from waste pyrolysis treatment. The modified

regression model for char yield (Ye, %) from co-pyrolysis was obtained as follows:

Ye= 178.17 + 0.11A- 0.608 + 0.03C- 6.17 X 10-3AC + 5.43 X 10-4 8 2 (6.4)

198

Table 6.3 lists the statistical parameters obtained from ANOV A for the char yield model.

The model was significant by F-test at the 5% confidence level (P < 0.05). The lack of fit

was not significant (P > 0.05), indicating that the fitness of model is acceptable. The main

effects of experimental factors on char yield are shown in Figure 6.8. Generally, char yield

increased with sawdust percentage because sawdust had higher fixed carbon content than

oily sludge. However, the pyrolysis temperature showed a negative effect on char yield,

and a higher pyrolysis temperature could lead to a lower char yield as a result of better

biomass conversion due to primary reactions of pyrolysis (Uzun et al. , 2006). These

primary reactions include decomposition of cellulose, hemicellulose, and lignin in sawdust

(Xiu and Shahbazi, 2012). Moreover, high temperature could facilitate the secondary

decomposition of char to generate non-condensable gaseous products (Xiu and Shahbazi,

2012; Menendez et al. , 2004; Canel et al. , 2004). Similar to pyrolysis temperature, a higher

heating rate led to a lower char yield. This might be due to the fact that rapid heating can

enhance the abundance of volatiles by fast decomposition of sawdust, resulting in fast

removal of high molecular volatiles and thus leaving low char amounts (Akhtar and Amin,

2012).

Figure 6.9 presents the interaction effect of experimental factors on char yield. Only

the interaction effect of heating rate and sawdust percentage on char yield was found

significant. It can be found that a high heating rate showed little effect on the char

production in the treatments with relatively low sawdust percentage. However, a high

heating rate could pose a significant negative effect on char yield when the sawdust

percentage was high. A higher heating rate could lead to lower oil yield but higher char

yield in the treatments with lower sawdust percentage (i.e. , < 25%), but could result in

199

lower char yield and higher oil yield in the treatments which had a relatively high sawdust

percentage in feedstock.

20 a

-- 18 ~ "O w ·;;:.. 15 .... Ill .r:: (.)

13

10

0 25 50 75 100

Sawdust percentage (wt.%)

28 b ,._ 23 C 32 (I) ·s:. 19 ... C'II .r:: (.)

14

10

400 450 500 550 600 Temperature (°C)

20 C

.-.. ~

18

3:! (I) ·;;:.. 15 ... C'II .r:: u

13

10

5 9 13 16 20 Heating rate (°C/min)

Figure 6.8 The main effects of single factors on char yield: (a) sawdust percentage

(temperature: 500 °C, heating rate: 12.5 °C/min), (b) temperature (sawdust

percentage: 50 wt.%, heating rate: 12.5 °C/min), and (c) heating rate (sawdust

percentage: 50 wt.%, temperature: 500 °C)

200

30

- 25 ~ 0 -"O 20 Q) ·5>-L.

15 cu ..c. (.)

10

0

Heating rate (°C/min) 5 0 Sawdust (wt.%)

Figure 6.9 The interaction effect of experimental factors on char yield

6.3.4 Synergistic effect of co-pyrolysis on oil yield

In order to investigate the synergistic effect of co-pyrolysis, the product yield rate was

further evaluated using various sawdust percentages at a pyrolysis temperature of 500 °C

and a heating rate of 20 °C/min. Table 6.4 lists the observed and predicted pyrolysis

product yields, while the prediction was obtained from Eq. (6.2). The actual oil yield rate

from the pyrolysis of oily sludge alone and sawdust alone were 36.2% and 49.2%,

respectively. Oily sludge generated the highest amount of non-condensable gaseous

products. It can be found that the oil yield was increased by 3-4% as compared to the

predicted yield in co-pyrolysis treatments with a sawdust percentage higher than 40%,

suggesting the presence of synergetic effect during co-pyrolysis (Onal et al. , 2014). For

example, the actual oil yield rate of co-pyrolysis treatment was 48.9% when sawdust /oily

201

sludge mass ratio was 3: 1, which is higher than the predicted yield (i.e., 46.0% ). At a mass

ratio of 1: 1, the actual oil yield reached 46.4%, which is also higher than the predicted

yield rate of 42. 7%. However, the synergetic effect was not obvious in co-pyrolysis

treatments with a sawdust percentage of lower than 50%, with the difference between

observed and predicted oil yield being less than 1 %. The synergistic effect of co-pyrolysis

on char yield was not obvious, but less gaseous products were obtained in co-pyrolysis

treatments at a sawdust percentage of higher than 50% as compared to the predicted value,

indicating that the increase in oil yield might be at the cost of gas yield. It can be found

from Table 6.4 that the enhancement in total oil yield from co-pyrolysis was mainly

attributed to the increase of bottom layer oil, indicating that adding oily sludge into sawdust

could result in better biomass conversion. Zuo et al. (2014) studied the co-pyrolysis of

sawdust and sewage sludge, and they also observed an increase in total oil yield which was

mainly contributed by the increase of bottom oil. Song et al. (2014) investigated the co-

pyrolysis of pine sawdust and lignite, and found that the experimental oil yield rate was

higher than that of calculated value in the treatment of a mixed feedstock (i.e., 20% of

lignite and 80% of sawdust) at 600 °C. They suggested that the synergetic effect can be

attributed to the different H/C and 0 /C molar ratios in two materials and the catalytic effect

of some alkali and alkaline metallic species in lignite (Song et al., 2014). High H/C and

0 /C molar ratios can facilitate the generation of radicals such as hydroxyl which can act

as hydrogen donors, promoting the cracking of heavy weight aromatic and aliphatic

compounds in oily sludge (Conesa et al., 2014; Song et al., 2014; Chen et al., 2014). Some

radicals could combine with others to form aldehydes, carboxylic acids, and ketones, and

thus increase the yield of bottom oil (Zuo et al., 2014).

202

Table 6.4 Observed product yields from co-pyrolysis experiments as compared with the

predicted values (temperature: 500 °C, heating rate: 20 °C/min)

Sawdust Top layer oil(%) Bottom layer oil (%) Total oil(%) Char(%) Gas(%)

percentage (%) Obs./Pre./Dif. * Obs./Pre./Dif. Obs./Pre./Dif. Obs./Pre./Dif. Obs./Pre./Dif.

0

25

40

50

60

75

100

36.2/36.2/0 36.2/36.2/0 13.7/13.7/0 50.1/50.1/0

24.6/27.2/-2.6 15.6/12.3/3.3 40.2/39.5/0.7 14.8/14.3/-0.5 45.0/46.3/-1.3

20.5/21.7/-1.2 20.3/19.7/0.6 40.8/41.4/-0.6 15.2/14.7/0.5 44.0/43.9/0.1

13.2/18.1/-4.9 33.2/24.6/8.6 46.4/42. 7 /3. 7 16.1 /15.0/1.1 37.5/42.3/-4.8

14.3/14.5/-0.2 33.7/29.5/4.2 48.2/44.0/4.0 15.7/15.2/0.5 36.3/40.8/-4.5

5.8/9.1 /-3 .3 43 .1/36.9/6.2 48.9/46.0/2.9 15.8/15 .6/0.2 35.3/38.5/-3.2

49.2/49.2/0 49.2/49.2/0 16.2/16.2/0 34.6/34.6/0

*Obs.: Observed yield; Pre.: Predicted yield; Dif.: Difference

6.3.5 Pyrolysis product characterization

The ultimate analysis of pyrolysis products and their properties are listed in Table 6.5.

The liquid oil from the pyrolysis of oily sludge alone contained the lowest moisture (i.e.,

1.2%) and the highest HHV (i.e., 46.7 MJ/kg), illustrating a similar characteristics of heavy

petroleum fuel oil in terms of elemental contents and HHV (Xiu and Shahbazi, 2012). This

suggests that the pyrolysis oil from oily sludge alone has a great potential to be used as a

fuel source. In contrast, the pyrolysis oil from sawdust treatment alone was associated with

the highest moisture (i.e., 23.5%) and the lowest HHV (i.e., 19.5 MJ/kg) among all the

treatments. It has been reported that typical wood-derived pyrolysis oil has a HHV of about

17 MJ/kg and a moisture of about 25 wt.% (Bridgwater, 2012). According to the ultimate

analysis, the oxygen content in pyrolysis oil from sawdust alone was high (i.e. 41.7%) as

203

compared to the oil derived from co-pyrolysis treatments, but its hydrogen content was

lower (i.e. 6.3%). The co-pyrolysis treatment of sawdust with oily sludge can improve the

quality of bottom layer oil derived from sawdust. For example, as a result of increasing

oily sludge percentage from O to 25%, the moisture content in the bottom layer oil

decreased by about 2% and the hydrogen content increased by about 2.5%. The highest

H/C ratio of the bottom layer oil was observed in the co-pyrolysis treatment with an oily

sludge percentage of25%. As a result of reduced moisture content and increased H/C ratio,

the HHV was increased from 19.5 to 24.5 MJ/kg, indicating that the co-pyrolysis treatment

had a synergistic effect on oil quality. The similar synergistic effect on pyrolysis oil quality

was observed in the treatment with an oily sludge percentage of 50%. An increment of 4.3

MJ/kg in HHV was obtained in the treatment with 50% of oily sludge. Generally, pyrolysis

oil from treatments with oily sludge addition have a higher HHV and lower moisture

content as compared to the pyrolysis oil from sawdust treatment alone. Among all

treatments, the co-pyrolysis with a sawdust/sludge ratio of 3: 1 can bring the most

significant improvement in oil quality, and this is consistent with the oil yield results that

the highest oil yield was obtained when the mixture of sawdust and oily sludge at a mass

ratio of 3: 1 was pyrolyzed at the optimum condition.

Figure 6: 10 shows the solid product char from the co-pyrolysis of oily sludge with

sawdust. As shown in Table 6.5, the carbon and hydrogen contents of the solid char residue

from co-pyrolysis were higher than those from the pyrolysis of oily sludge alone,

indicating that most PHCs contents were removed from oily sludge during pyrolysis

treatment. The results from this study have proven that pyrolysis is a highly effective

method for oily sludge treatment, by which the negative environmental impact can be

greatly reduced. A low carbon and hydrogen content were observed in solid residue as a

204

high oily sludge percentage was applied in the co-pyrolysis feedstock, but the oxygen

content showed an opposite trend. The char from the pyrolysis of sawdust alone had the

highest carbon and hydrogen content, and its HHV was 27.0 MJ/kg, indicating that it has

the value to be applied as a potential energy source (Akkaya, 2009; Chen et al., 2014).

Figure 6.10 The solid product from co-pyrolysis of sawdust with oily sludge at a mass ratio

of3:l (500 °C, 20 °C/min)

Table 6.5 Properties of pyrolysis products from co-pyrolysis experiments (heating rate:

20 °C /min, temperature: 500 °C)

Properties

Saw dust percentage (wt.%)

Pyrolysis oil

Moisture content (wt.%)

Specific gravity (g/mL)

Elemental composition (wt.%)

1.2

0.9

25 C

22.7

1.2

205

Values

50 C

22.3

1.1

75 C

21.6

1.1

100 C

23.5

1.2

C 82.0 52.3 51.8 52.1 51.5

H 13 .5 7.0 7.8 8.8 6.3

o · 4.5 40.6 39.9 38.5 41.7

N 0.1 0.5 0.6 0.5

H/C ratio 2.0 1.6 1.8 2.0 1.5

HHV (MJ/kg) 46.7 21.6 23.8 24.5 19.5

Char

Elemental composition (wt.%)

C 11.5 22.5 37.2 57.9 78.6

H 0.3 0.5 1.2 1.9 2.5

o · 88.0 76.7 61.2 39.9 18.3

N 0.2 0.3 0.2 0.3 0.6

H/C ratio 0.3 0.3 0.4 0.2 0.4

• By difference

6.4 Conclusions

The synergetic effect in co-pyrolysis of sawdust with oily sludge was investigated in

a fixed-bed reactor. The impact of three experimental factors including sawdust percentage,

temperature, and heating rate on the yield of pyrolysis oils and chars were examined using

RSM. The pyrolysis oil yield increased as a result of increasing sawdust percentage and

heating rate, and the optimum temperature for oil production was identified as 500 °C.

Char yield was enhanced by increasing sawdust percentage, but was compromised by high

heating rate and temperature. The interaction effect of heating rate and temperature, as well

206

as heating rate and sawdust percentage, were significant on pyrolysis oil yield. The

synergetic effect can bring about 4% increase in yield of bottom layer oil derived from co-

pyrolysis treatment as compared to the predicted value, but no obvious effect on char yield

was observed. Co-pyrolysis of sawdust with oily sludge at a sawdust percentage higher

than 40% can improve the quality of bottom layer oil by increasing its H/C ratio and HHV.

Pyrolysis oil from individual oily sludge treatment had the highest HHV and the lowest

moisture. High sawdust percentage in the mixture feedstock also can improve the quality

of char product by increasing its carbon content. The results confirmed that using oily

sludge as an additive in sawdust pyrolysis could improve pyrolysis oil production, and thus

the co-pyrolysis treatment could be a promising way for the disposal of hazardous oily

sludge.

207

Chapter 7 Conclusions and future research

7.1 Thesis conclusions

The purpose of this thesis was to develop combined oil recovery techniques for the

treatment of various refinery oily sludges. In doing so, four oil recovery approaches

including ultrasonic irradiation, solvent extraction, freeze/thaw, and pyrolysis were

selected. The oil recovery performance of different approaches was improved by

combining single treatment approach with each other for increasing the oil recovery rate,

reducing the treatment duration, and improving the quality of recovered oil as value-added

products.

The results of this study have demonstrated that using novel oil recovery techniques

can effectively handle complex oily sludge from different sources in petroleum refining

industries. These combined treatment methods have great potential to be commercialized

and applied in practical oily wastes treatment in order to reduce the environmental risks

posed by hazardous oily sludge. The followings are the major conclusions of this thesis.

7.1.1 The oil recovery and desalting effect of ultrasonic irradiation on oily sludge

In Chapter 3, the oil recovery and desalting effect of ultrasonic irradiation on refinery

tank bottom sludge was investigated. The impact of influential factors including ultrasonic

power, treatment duration, sludge-to-water (S/W) ratio, and sludge-water slurry

temperature on the treatment performance were studied. It can be concluded that ultrasonic

irradiation showed effective oil recovery and desalting effect on refinery tank bottom

sludge. During the treatment process, the temperature of the oily sludge-water mixture

208

increased as a result of ultrasonic irradiation. At an ultrasonic power of 75 W, the slurry

temperature increased from 25 to 77 °C within 6 min of irradiation. In comparison, the

slurry temperature increased from 25 to 61 °C within 6 min as an ultrasonic power of 21

W was applied.

Direct heating was able to recover part of the oil from oily sludge and also reduce the

salt content in recovered oil. When the bulk temperature of oily sludge-water slurry

mixture rose to 80 °C by direct heating, an oil recovery rate of 32.9% and a salt content of

6.0 mg/L in recovered oil were obtained. Under the same conditions, the lowest TPH

concentration and highest salt content were found to be 533 .3 mg/Land 159.8 mg/Lin the

separated wastewater as by-product.

Low ultrasonic power (i.e. , 21 W) and short irradiation duration (2 min) were not

effective for oil recovery from oily sludge. Under an ultrasonic power of 48 W, PHC

recovery significantly increased from 35.7% to 61.2% as the irradiation duration extended

from 2 to 8 min. The highest PHC recovery was 62.2% in the treatment with an ultrasonic

power of 7 5 W and treatment duration of 6 min. Further extended the duration could cause

decrease in PHC recovery.

The optimal sludge-to-water ratio was found to be 1 :4, yet too high or too low water

content in the sludge slurry system led to lower oil recovery rate. It was found that the salt

content in recovered oil was decreased to less than 5 mg/L after 6 min of ultrasonic

irradiation at a power of 75 W, which can meet the salt content requirement for feedstock

oil in most refineries. The separated wastewater from ultrasonic irradiation contained

relatively high concentrations of TPH (i.e., 1044 mg/L) and salt (i.e. , 588 mg/L) which

require proper treatment.

209

The recovered oil contained about 80.4% TPH which is higher than in crude oil (i.e.,

71.7%). Its TPH contained similar percentages of F2, F3 and F4 fractions as the original

sludge, which implies that ultrasonic irradiation did not cause the degradation of PHCs

fractions in oily sludge.

7 .1.2 The combined effect of ultrasonic irradiation and solvent extraction on oil recovery

In Chapter 4, the oil recovery effect of two UAE systems including ultrasonic assisted

extraction probe (UAEP) system and ultrasonic assisted extraction bath (UAEB) system

on tank bottom sludge was examined as compared to that of mechanical shaking extraction

(MSE) treatment.

Results showed that solvent type, SIS ratio, extraction duration, and shaking speed

have had significant influence on oil recovery in MSE treatment. The highest oil recovery

of CHX, EA, and MEK were 63 .7%, 35.2%, and 34.8%, respectively when the SIS ratio

was 4: 1, extraction duration was 60 min, and shaking speed was 250 rpm. Solvent type has

the most influence on oil recovery, followed by shaking speed, extraction duration, and

SIS ratio. More than 70% of solvent can be recovered after MSE treatment, and the highest

solvent recovery of 83.3% was observed in MEK. Using a higher SIS ratio also results in

a higher solvent recovery rate. As the increase of SIS ratio from 2:1 to 4:1, the solvent

recovery rate increased from 72.8% to 83.1 % accordingly.

In UAEP treatment, SIS ratio, treatment duration and ultrasonic power have had

significant effects on oil recovery. The highest oil recovery obtained by CHX was 59.8%

under the conditions of ultrasonic power of 21 W, SIS ratio of 4: 1, and treatment duration

of 21 s, which was higher than MSE. The extraction using MEK also can be improved by

UAEP with the highest oil recovery of38.3% as compared to 33.3% ofMSE. Among four

factors, ultrasonic power showed insignificant influence on oil recovery while the effects

210

of other three factors were significant. The effects of treatment duration and ultrasonic

irradiation power on solvent recovery were not significant, indicating that UAEP could not

cause negative influence on solvent recovery because solvent recovery of UAEP was

generally higher than that of MSE treatment.

In UAEB treatment, the highest oil recovery of 62.6% can be found in the treatment

using CHX. The influence of four factors on oil recovery were in an order of solvent type >

SIS ratio > treatment duration > bath temperature. The optimum condition for oil recovery

was using solvent of CHX, ultrasonic duration of 15 min, and S/S ratio of 4: 1. Among four

experimental factors, bath temperature showed insignificant effect on oil recovery.

However, bath temperature showed a significant negative effect on the solvent recovery.

The results showed that two extractions of the three treatment processes could be

sufficient to achieve a satisfactory oil recovery. Among the three solvents, CHX was

associated with the highest TPH recovery for each extraction process. The TPH in

recovered oil from UAEP treatment were generally higher than these of MSE treatment.

However, the analysis on the properties of recovered oil showed that the TPH content of

recovered oil using CHX was the lowest. Recovered oil from three treatments mainly

consisted of F2 and F3 fractions, while there was also significant amount ofF4 fraction in

the recovered oil by CHX. Results also showed that no degradation effect on PHCs by

UAE process since the distribution of PH Cs in recovered oil by UAE process was similar

to that of original oily sludge.

Recovered oil by MEK and EA presented higher quality than that by CHX. Recovered

oil also showed higher density, carbon residue, and viscosity as compared to crude oil. The

calorific value of recovered oil by MEK and EA was also higher than that of crude oil. The

asphaltene content in recovered oil by MEK and EA were lower than crude oil, which is

211

more preferable for downstream refining process. Results showed that UAE could be a fast

and effective oil recovery method for refinery oily sludge treatment.

7.1.3 The combined effect of solvent extraction and freeze/thaw on oil recovery

In Chapter 5, a combinational solution of solvent extraction and freeze/thaw was

developed for the oil recovery from high moisture dredged sludge from refinery

wastewater pond. The oil recovery rate, solvent recovery rate, and the waste reduction rate

of five solvents including CHX, EA, MEK, DCM, and 2-Pro were examined. The

performance of freeze thaw treatment on improving the quality of recovered oil in terms

of its TPH content was evaluated. The properties of solid residue as the by-product of the

treatment process were also investigated.

It was found that 2-Pro was associated with a much lower oil recovery than the other

four solvents, while CHX was associated with the highest oil recovery rate. The oil

recovery rate of CHX, DCM, MEK, EA, and 2-Pro reached up to 42.4%, 37.7%, 37.2%,

37.6%, and 17.1 %, respectively when the solvent-to-sludge ratio was 4: 1. For all solvents

except DCM, the oil recovery rate increased with the extraction duration until reaching

equilibrium at the extraction duration of 30 min.

The solvent recovery rate after oil extraction generally increased with the solvent-to-

sludge ratio. The CHX extraction treatment was associated with a considerable increase of

solvent recovery rate from 51.2% to 81.3% when the ratio increased from 1:1 to 4:1. All

the solvents except DCM showed satisfactory solvent recovery rate, with the highest

solvent recovery rate of only 48.5% for DCM but higher than 80% for the other four

solvents. The effect of extraction duration on the solvent recovery was not significant.

More than 50% of original oily sludge mass was reduced in all the treatments. It is

noted that the waste reduction rate for DCM extraction was the highest (i.e. 80%) among

212

all the treatments. The waste reduction rate increased from 50. 7% to 64.1 % (CHX), 56.3%

to 61.5% (MEK), 53 .2% to 66.4% (EA), and 49.3% to 57.2% (2-Pro) as the solvent-to-

sludge ratio increased from 1: 1 to 4: 1, respectively. The extraction duration illustrated little

effect on the variation of waste reduction rate when using CHX, MEK, EA, and 2-Pro

extraction. For DCM extraction, the waste reduction increased from 62.6% to 77.7% as

the extraction duration increased from 5 min to 30 min when the equilibrium was reached.

It was also found that under the optimum extraction conditions (SIS ratio of 4: 1 and

duration of 30 min), the oil recovery rate generally increased with the solvent extraction

cycles. After one solvent extraction cycle, CHX achieved a higher oil recovery (i.e., 41.0%)

as compared to that of MEK (i.e. 35.4%) and EA (i.e. 35.8%). After the second extraction,

the oil recovery rate increased to 56.0% (CHX), 40.4% (MEK), and 38.3% (EA),

respectively. There was only trivial increment in oil recovery as the third extraction applied,

indicating that only one or two extractions cycles could be enough for the study sludge.

Individual FIT treatment was not effective for oil recovery from refinery wastewater

pond sludge. However, FIT treatment can remove the emulsified water from the extractant

using MEK and EA as extraction solvents. The combined solvent extraction with FIT

treatment can significantly increase the TPH content in the recovered oil for both MEK

and EA extractions. With FIT treatment, the TPH content in the recovered oil increased

from 41.8% to 62.6% when using MEK extraction and from 41.0% to 58.8% when using

EA extraction. The improvement of FIT treatment on TPH content in the recovered oil by

CHX extraction was not significant. The increase ofTPH in the recovered oil by increased

number of FIT cycles were not significant for each solvent, showing that one FIT cycle

would be sufficient for the dewatering of extractant obtained from the high-moisture

petroleum sludge.

213

The F2, F3, and F4 fractions accounted for 26.5%, 69.1%, and 4.3% ofTPH in the

recovered oil by CHX extraction with freeze/thaw. On the other hand, the F2 and F3

fractions in the oil recovered by the combination of MEK or EA extraction with

freeze/thaw treatment were slightly higher, while the F4 fraction was lower than that from

CHX extraction. In comparison to crude oil, the recovered oil from the study sludge

contained more F3 fraction but less F2 fraction.

Although solvent extraction showed obvious effect on oil recovery from the oily

sludge, the TPH content in the solid residue was still much higher than the clean-up level.

Solvent extraction with FIT treatment had little removal effect on heavy metals in oily

sludge, and most heavy metals were remained and accumulated in the solid residues,

indicating that the solid residues from this treatment require proper management.

7.1.4 The co-pyrolysis of sawdust with oily sludge for pyrolysis oil production

In Chapter 6, the synergetic effect in co-pyrolysis of sawdust with oily sludge was

investigated in a fixed-bed reactor. The impact of three experimental factors including

sawdust percentage, temperature, and heating rate on the yield of pyrolysis oils and chars

were examined using RSM. The characteristics of products from the co-pyrolysis were

determined to evaluate their possibility of being a potential energy source and

petrochemical feedstock.

According to the TGA, the decomposition behavior of sawdust can be divided into

three stages: ambient to 110, 110 to 700, and 700 to 900 °C, and the decomposition of oily

sludge can be divided into four stages: ambient to 110, 110-150, 150-550, and 550 to

900 °C. The maximum weight loss of sawdust and oily sludge was observed at 350 °C and

150 °C, respectively.

214

The oil product from the co-pyrolysis of sawdust with oily sludge had two layers. The

top layer was mainly from the pyrolysis of oily sludge alone and the bottom layer was

derived from sawdust pyrolysis alone. The oil yield of sawdust pyrolysis alone was 46.5%,

while the pyrolysis of oily sludge alone can generate 35. 7% of liquid oil. The highest oil

yield rate was 48.8% from the co-pyrolysis of sawdust with oily sludge with a sawdust

percentage of 80%, which was higher than the pyrolysis of sawdust or oily sludge alone.

Two quadratic models were developed for investigating the main and interaction

effect of experimental factors on oil and char yield in co-pyrolysis treatment. The model

for predicting pyrolysis oil production was: Oil yield(%)= -102.85 + 0.104A + 0.49B +

2.0SC + O.OlAC - 5.09 X 10-3 BC - 1.46 X 10-3 A2 - 4.06 X 10-4 0 2 and for predicting char

yield was: Char yield (%) = 178.17 + O.llA- 0.60B + 0.03C- 6.17 x 10-3AC + 5.43 x

10-4 0 2 . The A, B, and C represent sawdust percentage (wt.%), temperature (cC), and

heating rate (cC/min). Both models were robust and showed good fit to experimental

results.

The main effect of experimental factors showed that sawdust percentage had a

positive effect on pyrolysis oil yield. Generally, pyrolysis oil yield increased with the

increase of sawdust percentage in feedstock. Both too high and too low temperature

showed a negative effect on oil yield and the optimal temperature for pyrolysis oil

production was 500 cc. A higher heating rate was also preferred for pyrolysis oil

production.

The interaction effect of heating rate and temperature showed that low heating rate

and low pyrolysis temperature could result in the low oil yield. Increasing heating rate

from 5 to 20 cc/min could enhance the oil yield in treatments with a temperature lower

than 500 cc. When the treatment temperature exceeded 550 cc, a high heating rate (i.e.

215

20 °C/min) could pose a negative effect on pyrolysis oil production. The interaction effect

of sawdust amount in mixture and heating rate showed that high heating rate was preferred

in the treatment with a sawdust percentage higher than 50%; however, high heating rate

could compromise the oil yield in the treatments which had a low sawdust percentage (i.e.,

< 25%). The optimum condition for pyrolysis oil yield was heating 20 g sawdust to 500 °C

at a heating rate of 20 °C/min.

Char yield increased as the result of a higher sawdust percentage in feedstock.

Moreover, higher pyrolysis temperature and heating rate could lead to a lower char yield.

The interaction effect of heating rate and sawdust percentage on char yield was significant.

High heating rate had little effect on char production in the treatments with relatively low

sawdust percentage; however, high heating rate could have a negative effect on char yield

when sawdust percentage in feedstock was high.

Under the optimum condition, the actual oil yield for the pyrolysis of oily sludge

alone and individual sawdust alone were 36.2% and 49.2%, respectively. As compared to

the predicted yield, about an increase of 3-4% in actual oil yield was observed in co-

pyrolysis treatment with a sawdust percentage higher than 40%, suggesting the presence

of synergetic effect during the co-pyrolysis process. The effect of co-pyrolysis on char

yield was not obvious, but less gases products were obtained from co-pyrolysis as

compared to the predicted value, and thus the increase in oil yield might be at the cost of

gas yield. The enhancement in oil yield can mainly be attributed to the increase of bottom

layer oil, indicating adding oily sludge into sawdust could result in better biomass

convers10n.

Product characterization showed that pyrolysis oil from the pyrolysis of oily sludge

alone contained the lowest moisture (1.2%) and the highest HHV (46.7 MJ/kg), which

216

showed a similar property to heavy petroleum fuel. The pyrolysis oil from pyrolysis of

sawdust alone (bottom layer oil) had the highest moisture (23.5%) and the lowest HHV

(19.5 MJ/kg). Co-pyrolysis of sawdust with oily sludge can improve the quality of bottom

layer oil. As a result of increasing oily sludge percentage from O to 25%, the moisture

content of bottom phase oil decreased by about 5% and the hydrogen content increased by

about 2.5%. The highest H/C ratio of bottom layer oil was observed in the treatment with

an oily sludge percentage of 25%. As a result of reduced moisture content and increased

H/C ratio, the HHV increased from 19.5 to 24.2 MJ/kg, indicating that co-pyrolysis had a

synergistic effect on the quality of oil derived from sawdust pyrolysis.

The carbon and hydrogen content in solid residue from oily sludge pyrolysis alone

was low, suggesting most petroleum hydrocarbons were removed from oily sludge after

pyrolysis treatment. The carbon and hydrogen content of char increased with sawdust

percentage in feedstock. The char from pyrolysis of sawdust alone had the highest carbon

and hydrogen content, and its HHV was 27.0 MJ/kg, proving it has the value to be applied

as a potential energy source.

7.2 Research achievements

In this dissertation research, four novel oil recovery methods were developed for the

treatment of refinery oily sludges. The results of this research indicate that the combined

oil recovery methods have the potential to be applied for the treatment of different complex

oily wastes in petroleum refining industries to meet sustainable development principles.

These combined oil recovery methods offer several advantages over conventional

treatment approaches. First, these novel methods can handle oily sludges generated from

major sources in a typical petroleum refinery such as tank bottom sediments, API separator

217

sludge, and high moisture wastewater pond sludge; Second, these methods can

successfully recycle valuable energy content in a form of recovered oil from wastes,

significantly reduce the volume and hazardous level of oily wastes, and effectively curtail

the adverse impact on the environment resulting from the disposal of such wastes; Third,

these methods can reduce the time burdens for the remediation of recalcitrant oily sludges

in cold environments such as the vast area of Canada; and fourth, the recovered oil showed

good properties as compared to crude oil and can be further processed for the production

of various valuable petrochemical products. The achievements of this dissertation research

will provide useful solutions for the management of oily sludge in petroleum refining

industry.

7.3 Future research

Although this study has developed several novel oil recovery methods for the

effective treatment of various refinery oily sludge, there are still rooms for further

improving these methods. Recommendations regarding possible future work in areas

related to this study are suggested as follows:

1. The properties of solid residues from oil recovery treatment need to be evaluated as a

by-product from treatment process. Since most PHCs was recovered as oil, there was

a small amount of unrecoverable PHCs fractions left in solid residues. As compared

with recoverable PHCs, these leftover PHCs in solid residues are relatively heavy

fractions associated with larger molecules, lower bio-availability, and are more

recalcitrant to conventional decontamination processes such as biodegradation and

thermo-desorption. Therefore, the solid residues could cause environmental pollution if

they are not properly handled. Moreover, heavy metals concentrations in solid residues

218

might be higher than the original oily sludge due to limited removal effect on heavy

metals by oil recovery treatment. Therefore, it is important to investigate the

characteristics of contaminants such as leftover PH Cs and heavy metals in solid residues

from oil recovery treatment and develop cost-effective decontamination techniques for

the treatment of solid residues.

2. In UAE treatment using MEK and EA as the solvents, there were significant amount of

asphaltene compounds left in solid residues. During the extraction, insoluble

asphaltenes in extractant aggregated and settled down under the effect of gravity. These

asphaltene fractions settled down slower than the solid particles due to the fact that

asphaltenes have lower density, and thus they were accumulated on the top of solid

residues after gravitational settling. The asphaltenes can be separated from solid

residues readily and have the potential to be reused as hydrophobic painting or

pavement materials. So it is encouraged to investigate the property of this non-

extractable layer to further improve the overall waste recycling and reutilization.

3. In the co-pyrolysis of sawdust with oily sludge, the mechanisms of the composition

conversion and synergistic reactions were still unclear. It is important to further

investigate the reaction pathways of different organic vapors generated during the co-

pyrolysis treatment. GC-MS, GC-TCD, and FT-IR analytical methods should be used

to investigate the composition change in pyrolysis oils, gases products, and chars,

respectively.

4. The biomass-derived oil from co-pyrolysis treatment still needs to be improved. It is of

great importance to find an effective solution for the upgrading of pyrolysis oil from

biomass. Upgrading of such oil products can be achieved by reducing moisture

andoxygen content and increase H/C ratio in oil products. Catalytic cracking and

219

polymerization processes such as the Fischer-Tropsch process can be used for the

conversion of synthesis gas and production of synthetic fuels, can be used for further

upgrading of the pyrolysis oils. The char from the biomass pyrolysis can not only be

used as an energy source, but also has the potential to be modified into adsorbents for

pollution control. It has been reported that the char products from oily wastes pyrolysis

can be modified for the removal of toxic heavy metals in wastewater. If the modified

char can adsorb and stabilize the heavy metals in oily sludge, the co-pyrolysis treatment

could be a more appealing waste treatment method. Moreover, the gaseous products

from the co-pyrolysis of sawdust with oily sludge were unknown in this study. The

gaseous products can be served as an energy source to provide supplementary power

for pyrolysis process, and thus it is worthy to investigate the characteristics of gas

products and evaluate their heating values. Also, it is of interest to evaluate the

utilization of pyrolysis gaseous and chars as heat sources for reducing the energy

consumption of co-pyrolysis treatment.

References

220

Abalos, A., Pinaso, A., Infante, M.R., Casals, M., Garcia, F., and Manresa, A., 2001.

Physicochemical and antimicrobial properties of new rhamnolipids by

Pseudomonas aeruginosa ATlO from soybean oil refinery wastes. Langmuir, 17,

1367-1371.

Abdel Azim, A.A., Abdul-Raheim, A.M., Kamel, R.K., and Abdel-Raouf, M.E., 2011.

Demulsifier systems applied to breakdown petroleum sludge. Journal of

Petroleum Science and Engineering, 78, 364-370.

Abdul, W.S ., Elkamel, A., Madhuranthakam, C.R., and Al-Otaibi, M.B., 2006. Building

inferential estimators for modeling product quality in a crude oil desalting and

dehydration process. Chemical Engineering Process, 45, 568-577.

Abdulbari, H.A., Abdurahman, N.H., Rosli, Y.M., Mahmood, W.K. , and Azhari, H.N.,

2011. Demulsification of petroleum emulsions using microwave separation

method. International Journal of Physical Sciences, 6, 5376-5382.

Abnisa, F. and Daud, W.M.A.W., 2014. A review on co-pyrolysis of biomass: An optional

technique to obtain a high-grade pyrolysis oil. Energy Conversion and

Management, 87, 71-85.

Abnisa, F., Arami-Niya, A. , Wan Daud, W.M.A. , Sahu, J.N. , and Noor, I.M., 2013.

Utilization of oil palm tree residues to produce bio-oil and bio-char via pyrolysis.

Energy Conversion and Management, 76, 1073-1082.

Adewuyi, Y.G., 2001. Sonochemistry: Environmental science and engineering

applications. Industrial & Engineering Chemistry Research, 40, 4681-4 715.

Ahmadi, M., Vahabzadeh, F., Bonakdarpour, B., Mofarrah, E., and Mehranian, M., 2005.

Application of the central composite design and response surface methodology to

the advanced treatment of olive oil processing wastewater using Fenton's

peroxidation. Journal of Hazardous Materials, B123, 187-195.

221

Akhtar, J. and Amin, N.S., 2012. A review on operating parameters for optimum liquid oil

yield in biomass pyrolysis. Renewable & Sustainable Energy Reviews, 16, 5101-

5109.

Akkaya, A.V., 2009. Proximate analysis based multiple regression models for higher

heating value estimation of low rank coals. Fuel Processing Technology, 90 (2),

165-170.

Alexander, M., 2000. Aging, bioavailability, and overestimation of risk from

environmental pollutants. Environmental Science & Technology, 34, 4259-4265.

Al-Futaisi, A., Jamrah, A. , Yaghi, B. , and Taha, R. , 2007. Assessment of alternative

management techniques of tank bottom petroleum sludge in Oman. Journal of

Hazardous Materials, 141, 557-564.

Ali, M.F. and Alqam, M.H., 2000. The role of asphaltenes, resins and other solids in the

stabilization of water in oil emulsions and its effects on oil production in Saudi oil

fields. Fuel, 79, 1309-1316.

Al-Mutairi, N., Bufarsan, A., and Al-Rukaibi, F., 2008. Ecorisk evaluation and

treatability potential of soils contaminated with petroleum hydrocarbon-based

fuels. Chemosphere, 74, 142-148.

Al-Otoom, A., Allawzi, M., Al-Omari, N., and Al-Hsienat, E., 2010. Bitumen recovery

from Jordanian oil sand by froth flotation using petroleum cycles oil cuts. Energy,

35, 4217-4225.

Al-Shamrani, A.A., James, A., and Xiao, H., 2002. Separation of oil from water by

dissolved air flotation. Colloids and Surfaces A: Physicochemical and

Engineering, 209, 15-26.

Al-Zahrani, S.M. and Putra, M.D., 2013 . Used lubricating oil regeneration by various

solvent extraction techniques. Journal of Industrial and Engineering Chemistry,

9, 536-539.

222

Anna, L.M.S., Soriano, A.U., Gomes, A.C., Menezes, E.P., Gutarra, M.L.E., Freire,

D.M.G., and Pereira, N., 2007. Use ofbiosurfactant in the removal of oil from

contaminated sandy soil. Journal of Chemical Technology and Biotechnology, 82,

687-691.

API, 1989. API Environmental Guidance Document: Onshore solid waste management

in exploration and production operations, in: American Petroleum Institute (API),

Washington DC.

API, 1992. Technical Data Book-Petroleum Refining, 5th Edition. in: American

Petroleum Institute (API), Washington DC.

API, 2010. Category Assessment Document for Reclaimed petroleum hydrocarbons:

Residual hydrocarbon wastes from petroleum refining. U.S. EPA HPV Challenge

Program, in: American Petroleum Institute (API), Washington DC.

Appleton, T.J., Colder, R.I., Kingman, S.W., Lowndes, I.S., and Read, A.G., 2005.

Microwave technology for energy-efficient processing of waste. Applied Energy,

81, 85-113.

Aqsha, A. , Mahinpey, N., Mani, T., Salak, F., and Murugan, P., 2011. Study of sawdust

pyrolysis and its devolatilisation kinetics. The Canadian Journal of Chemical

Engineering, 89, 1451-1457.

Asadullah, M., Ab Rasid, N.S., Kadir, S.A.S.A., and Azdarpour, A., 2013. Production and

detailed characterization of bio-oil from fast pyrolysis of palm kernel shell.

Biomass &Bioenergy, 59, 316-324.

ASTM D3230, 2009. Standard test method for salts in crude oil (electrometric method).

Pennsylvania, United States, pp.1-6.

ASTM, 2007. ASTM D3279, Standard test method for n-heptane insoluble. American

Society of Testing Materials, Philadelphia, PA.

223

ASTM, 2009. ASTM D5185, Standard test method for determination of additive

elements, wear metals, and contaminants in used lubricating oils and

determination of selected elements in base oils by Inductively Coupled Plasma

Atomic Emission Spectrometry (ICP-AES). American Society of Testing

Materials, Philadelphia, PA.

ASTM, 2013 . ASTM D2974-00 Method A, Standard test method for moisture, ash, and

organic matter of peat and other organic soils. American Society of Testing

Materials, Philadelphia, PA.

Avila-Chavez, M.A., Eustaquio-Rinc6n, R. , Reza, J. , and Trejo, A. , 2007. Extraction of

hydrocarbons from crude oil tank bottom sludges using supercritical ethane.

Separation Science and Technology, 42, 2327-2345.

Aydin, M.E., Ozcan, S. , and Tor, A. , 2007. Ultrasonic solvent extraction of persistent

organic pollutants from airborne particles. Clean, 35(6), 660-668 .

Ayotamuno, M.J. , Okparanma, R.N. , Nweneka, E.K., Ogaji, S.O.T. , and Probert, S.D.,

2007. Bio-remediation of a sludge containing hydrocarbons. Applied Energy, 84,

936-943 .

Ball, A.S., Stewart, R.J., and Schliephake, K., 2012. A review of the current options for

the treatment and safe disposal of drill cuttings. Waste Management Research, 30,

457-473 .

Banjoo, D.R. and Nelson, P.K., 2005. Improved ultrasonic extraction procedure for the

determination of polycyclic aromatic hydrocarbons in sediments. Journal of

Chromatography A , 1066, 9-18.

Bassam, M. and Mohammed, N.B., 2005 . Biodegradation of total organic carbons (TOC)

in Jordanian petroleum sludge. Journal of Hazardous Materials, Bl 20, 127-134.

Bendicho, C., de la Calle, I., Pena, F., Costas, M., Cabaleiro, N. , and Lavilla, I., 2012.

Ultrasound-assisted pretreatment of solid samples in the context of green

analytical chemistry. Trends in Analy tical Chemistry, 31 , 50-60.

224

Bernardo, M., Lapa, N., Gonc;alves, M. , Mendes, B., Pinto, F. , and Fonseca, I. , 2012.

Physicochemical properties of chars obtained in the co-pyrolysis of waste

mixtures. Journal of Hazardous Materials, 219, 196-202.

Bhattacharyya, J.K. and Shekdar, A.V., 2003. Treatment and disposal ofrefinery

sludges: Indian scenario. Waste Management & Research, 21, 249-261.

Biagini, E., Barontini, F., and Tognotti, L., 2006. Devolatilization of biomass fuels and

biomass components studied by TG/FTIR technique. Industrial & Engineering

Chemistry Research, 45, 4486-4493.

Biswal, B.K., Tiwari, S.N., and Mukherji, S., 2009. Biodegradation of oil in oily sludges

from steel mills. Bioresource Technology, 100, 1700-1703 .

Bossio, J.P., Harry, J., and Kinney, C.A., 2008. Application of ultrasonic assisted

extraction of chemically diverse organic compounds from soils and sediments.

Chemosphere, 70, 858-864.

Bossio, J.P. , Harry, J. , and Kinney, C.A. , 2008. Application of ultrasonic assisted

extraction of chemically diverse organic compounds from soils and sediments.

Chemosphere, 70, 858-864.

BP, 2012. British petroleum (BP) statistical review of world energy, London, 2012, pp.

16.

Brebu, M. and Spiridon, I. , 2012. Co-pyrolysis of LignoBoost® lignin with synthetic

polymers. Polymer Degradation and Stability, 97, 2104-2109.

Bridgwater, A.V., 2012. Review of fast pyrolysis of biomass and product upgrading.

Biomass & Bioenergy, 38, 68-94.

Bridgwater, A.V., Meier, D., and Radlein, D., 1999. An overview of fast pyrolysis of

biomass. Organic Geochemistry, 30, 1479-1493.

225

Bridle, T. and Unkovich, I., 2002. Critical factors for sludge pyrolysis in Australia.

Water(Australia), 29, 43-48.

Bridle, T.R. and Pritchard, D., 2004. Energy and nutrient recovery from sewage sludge

via pyrolysis. Water Science and Technology, 50, 169-175.

Butler, E., Devlin, G., Meier, D., and McDonnell, K., 2011. A review of recent laboratory

research and commercial developments in fast pyrolysis and upgrading. Renewable

& Sustainable Energy Reviews, 15 (8), 4171-4186.

Butt, T.E., Lockley, E., and Oduyemi, K.O.K., 2008. Risk assessment of landfill disposal

sites-State of the art. Waste Management, 28, 952-964.

Calvo, C., Manzanera, M., Silva-Castro, G.A., Uad, I., and Gonzalez-Lopez, J., 2009.

Application of bioemulsifiers in soil oil bioremediation processes. Science of the

Total Environment, 407, 3634-3640.

Cambiella, A., Benito, J.M., Pazos, C., and Coca, J., 2006. Centrifufal separation

efficiency in the treatment of waste emulsified oils. Chemical Engineering

Research and Design, 84, 69-76.

Cameotra, S.S. and Singh, P., 2008. Bioremediation of oil sludge using crude

biosurfactants. International Biodeterioration & Biodegradation, 62, 274-280.

Canel, M., Misirlioglu, Z., and Sinag, A., 2004. Influence of experimental conditions on

product distribution during pyrolysis of tuncbilek lignite (turkey) at low heating

rates. Energy Sources, 26, 1265-1276.

Canselier, J.P., Delmas, H., Wilhelm, AM., and Abismai1, B., 2007. Ultrasound

emulsification-an overview. Journal Of Dispersion Science And Technology, 23,

333-349.

Cao, Q., Jin, L.E., Bao, W., and Lv, Y., 2009. Investigations into the characteristics of oils

produced from co-pyrolysis of biomass and tire. Fuel Processing Technology, 90,

337-342.

226

Capelo, J.L. and Mota, A.M., 2005 . Ultrasonication for analytical chemistry. Current

Analy tical Chemistry, 1, 193-201.

CCME., 2001. Reference method for the canada-wide standard for petroleum

hydrocarbons in soil-tier 1 method. Canadian Council of Minister of the

Environment (CCME), Winnipeg, Manitoba, pp. 1310.

Certini, G., 2005. Effects of fire on properties of forest soils: A review. Oecologia,

143,1-10.

Chan, C.C. and Chen, Y.C. , 2002. Demulsification ofW/0 emulsions by microwave

radiation. Separation Science and Technology, 37, 3407-3420.

Chang, C.Y. , Shie, J.L. , Lin, J.P., Wu, C.H. , Lee, D.J. , and Chang, C.F., 2000. Major

products obtained from the pyrolysis of oil sludge. Energy & Fuels, 14, 1176-

1183.

Chemat, F., Huma, Z., Khan, M.K. , 2011. Applications of ultrasound in food technology:

Processing, preservation and extraction. Ultrasonics Sonochemistry, 18, 813-835.

Chen, G. , Liu, C. , Ma, W. , Zhang, X. , Li, Y. , Yan, B., and Zhou, W., 2014. Co-pyrolysis

of com cob and waste cooking oil in a fixed bed. Bioresource Technology, 166,

500-507.

Chen, G.H. and He, G.H., 2003. Separation of water and oil from water-in-oil emulsion

by freeze/thaw method. Separation and Purification Technology, 31 , 83-89.

Cheremisinoff, N.P. and Rosenfeld, P. , 2009. Chapter 1 - The petroleum industry, in:

Handbook of Pollution Prevention and Cleaner Production - Best Practices in The

Petroleum Industry, William Andrew Publishing, Oxford, pp. 1-97.

Chiaramonti, D., Oasmaa, A. , and Solantausta, Y. , 2007. Power generation using fast

pyrolysis liquids from biomass. Renewable & Sustainable Energy Reviews, 11 ,

1056-1086.

227

Chin, C.L., Yi, C.H. , Yu, H.W., and Jo, S.C., 2009. Biosurfactant-enhanced removal of

total petroleum hydrocarbons from contaminated soil. Journal of Hazardous

Materials, 167, 609-614.

Chirwa, E .M.N. , Mampholo, T. , and Fayemiwo, 0 ., 2013 . Biosurfactants as demulsifying

agents for oil recovery from oily sludge-performance evaluation. Water Science

and Technology, 61(12), 2875-2881.

Christodoulatos, C. and Koutsospyros, A., 1998. Bioslurry reactors, in: G.A.

Lewandowski, L.J. Defilippi (Eds.) Biological treatment of hazardous waste,

Wiley, New York, pp. 69-101.

Christofi, N . and Ivshina, I.B., 2002. A review: Microbial surfactants and their use in

field studies of soil remediation. Journal of Applied Microbiology, 93, 915-929.

Chu, C.P. , Chang, B.V. , Liao, G.S. , Jean, D.S., and Lee, D.J. , 2001. Observations on

changes in ultrasonically treated waste-activated sludge. Water Research, 35,

1038-1046.

Chun, K.L. , Chen, G.H., and Lo, M.C., 2004. Salinity effect on freeze/thaw conditioning

of activated sludge with and without chemical addition. Separation and

Purification Technology, 34, 155-164.

Chung, H.I. and Kamon, M. , 2005. Ultrasonically enhanced electrokinetic remediation

for removal of Pb and phenanthrene in contamianted soils. Engineering Geology,

77, 233-242.

Collings, A.D., Farmer, A.D., Gwan, P.B. , Sosa Pintos, A.P. , and Leo, C.J., 2006.

Processing contaminated soils and sediments by high power ultrasound. Minerals

Engineering, 19, 450-453 .

Conaway, L.M. , 1999. Method for processing oil refining waste, in: Continuum

Environmental, Inc., United States.

228

Conesa, J.A., Molt6, J., Ariza, J., Ariza, M., and Bameto, A.G., 2014. Study of the thermal

decomposition of petrochemical sludge in a pilot plant reactor. Journal of

Analytical and Applied Pyrolysis, 107, 101-106.

Conner, J.R. and Hoeffner, S.L., 1999. A critical review of stabilisation/solidification

technology. Critical Reviews in Environmental Science and Technology, 28, 397-

462.

Cort, T.L. , Song, M.S., and Bielefeldt, A.R., 2002. Non ionic surfactant effects on

pentachlorophenol degradation. Water Research, 36, 1253-1261.

Cui, B., Cui, F., Jing, G., Xu, S., Huo, W., and Liu, S., 2009. Oxidation of oily sludge in

supercritical water. Journal of Hazardous Materials, 165, 511-517.

Cuypers, C., Pancras, T., Grotenhuis, T., and Rulkens, W., 2002. The estimation of PAH

bioavailability in contaminated sediments using hydroxypropyl-beta-cyclodextrin

and Triton X-100 extraction techniques. Chemosphere, 46, 1235-1245.

Czernik, S. and Bridgwater, A.V., 2004. Overview of applications of biomass fast

pyrolysis oil. Energy & Fuels, 18, 590-598.

da Rocha, O.R.S. , Dantas, R.F., Duarte, M.M.M.B., Duarte, M.M.L., and da Silva, V.L.,

2010. Oil sludge treatment by photocatalysis applying black and white light.

Chemical Engineering Journal, 157, 80-85.

da Silva, L.J., Alves, F.C., and de Fran9a, F.P., 2012. A review of the technological

solutions for the treatment of oily sludges from petroleum refineries. Waste

Management & Research, 30, 1016-1030.

Dara, O.R. and Sarah, C. , 2003. Just oil? The distribution of environmental and social

impacts of oil production and consumption. Annual Review of Environment and

Resources, 28, 587-617.

229

Delille, D. , Pelletier, E., and Coulon, F., 2007. The influence of temperature on bacterial

assemblages during bioremediation of a diesel fuel contaminated subantarctic

soil. Cold Regions Science and Technology, 48, 74-83.

Dewil, R., Baeyens, J., and Goutvrind, R., 2006. Ultrasonic treatment of waste activated

sludge. Environmental Progress & Sustainable Energy, 25, 121-128.

Dewulf, J. and Langenhove, H.V., 2001. Ultrasonic degradation of trichloroethylene and

chlorobenzene at micromolar concentrations: kinetics and modeling. Ultrasonics

Sonochemistry, 8, 143-150.

Dominguez, A., Menendez, J.A., Inguanzo, M., and Pis, J.J., 2005. Investigations into the

characteristics of oils produced from microwave pyrolysis of sewage sludge. Fuel

Processing Technology, 86, 1007-1020.

Edwards, K.R., Lepo, J.E., and Lewis, M.A., 2003.Toxicity comparison ofbiosurfactants

and synthetic surfactants used in oil spill remediation to two estuarine species.

Marine Pollution Bulletin, 46, 1309-1316.

El Naggar, A.Y., Saad, E.A., Kandil, AT., and Elmoher, H.O., 2010. Petroleum cuts as

solvent extractor for oil recovery from petroleum sludge. Journal of Petroleum

Technology and Alternative Fuels, 1, 10-19.

Electorowicz, M. and Hatim, J., 2000. Application of surfactant enhanced elektrokinetics

for hydrocarbon contaminated soils, in: 53th Canadian Geotechnical Conference,

Montreal, pp. 617-624.

Elektorowicz, M. and Habibi, S., 2005. Sustaibable waste management: Recovery of

fuels from petroleum sludge. Canadian Journal of Civil Engineering, 2005, 32,

164-169.

Elektorowicz, M., Habibi, S., and Chifrina, R., 2006. Effect of electrical potential on the

electro-demulsification of oily sludge. Journal of Colloid and Interface Science,

295, 535-541.

230

EPA, 1980. Resource Conservation and Recovery Act: Hazardous Waste Regulations:

Identification and Listing of Hazardous Waste, in: U.S. Environmental Protection

Agency (EPA), Washington DC.

EPA, 1991. Safe, environmentally acceptable resources recovery from oil refinery

sludge, U.S. Environmental Protection Agency (EPA), Washington DC.

EPA, 1993. Clean Water Act, 58. United States Environmental Protection Agency (US

EPA), Washington DC, Section503, No.32.

EPA, 1994. Test methods of evaluating solid waste, physical/chemical methods (SW-

846). United States Environmental Protection Agency (US EPA), Cincinnati, OH.

EPA, 2007. Method 8015C, Nonhalogenated organics by gas chromatography, Revision

3. United States Environmental Protection Agency, Cincinnati, OH.

EPA, 2008. Hazardous waste listings, a user-friendly reference document, U.S.

Environmental Protection Agency (EPA), Washington DC.

Fakhru'l-Razi, A., Pendashteh, A., Abdullah, L.C., Biak, D.R.A., Madaeni, S.S., and

Abidin, Z.Z., 2009. Review of technologies for oil and gas produced water

treatment. Journal of Hazardous Materials, 170, 530-551.

Faksness, L.G., Grini, P.G., and Daling, P.S., 2004. Partitioning of semi-soluble organic

compounds between the water phase and oil drophlets in produced water. Marine

Pollution Bulletin, 48, 731-742.

Fang, C.S. and Lai, P.M.C., 1995. Microwave heating and separation of water-in-oil

emulsion. The Journal of Microwave Power & Electromagnetic Energy, 30, 46-

57.

Fei, J., Zhang, J., Wang, F., and Wang, J., 2012. Synergistic effects on co-pyrolysis of

lignite and high-sulfur swelling coal. Journal of Analytical and Applied Pyrolysis,

95, 61-67.

231

Feng, D. and Aldrich, C., 2000. Sonochemical treatment of simulated soil contaminated

with diesel. Advances in Environmental Research, 4, 103-112.

Fernandez, L.F., Valenzuela, E.C., Marsch, R., Martinez, S.C., Vazquez, N.E., and

Dendooven, L., 2011. Microbial communities to mitigate contamination of P AHs

in soil - possibilities and challenges: a review. Environmental Science and

Pollution Research, 18, 12-30.

Ferrarese, E., Andreottola, G., and Oprea, I.A., 2008. Remediation of PAR-contaminated

sediments by chemical oxidation. Journal of Hazardous Materials, 152, 128-139.

Fisher, J.A., Scarlett, M.J., and Stott, A.D., 1997. Accelerated solvent extraction: an

evaluation for screening of soils for selected U.S . EPA semivolatile organic

priority pollutants. Environmental Science & Technology, 31, 1120-1127.

Flores, R., Blass, G., and Dominguez, V., 2007. Soil remediation by an advanced

oxidative method assisted with ultrasonic energy. Journal of Hazardous

Materials, 140, 399-402.

Fonts, I., Gea, G., Azuara, M., Abrego, J., and Arauzo, J., 2012. Sewage sludge pyrolysis

for liquid production: a review. Renewable & Sustainable Energy Reviews, 16,

2781-2805.

Fortuny, M., Oliveira, C.B.Z., Melo, R.L.F.V., Nele, M., Coutinho, R.C.C., and Santo,

A.F., 2007. Effect of salinity, temperature, water content, and pH on the

microwave demulsification of crude oil emulsion. Energy & Fuels, 21, 13 5 8-

1364.

Frank, J. and Castaldi, P.E., 2003. Tank-based bioremediation of petroleum waste

sludges. Environmental Progress, 22, 25-36.

Fujii, T., Hayashi, R., Kawasaki, S., Suzuki, A., and Oshima, Y., 2011. Water density

effects on methanol oxidation in supercritical water at high pressure up to 100

MPa. The Journal of Supercritical Fluids, 58, 142-149.

232

Gazineu, M.H.P., de Araujo, A.A., Brandao, Y.B., Hazin, C.A., and Godoy, J.M., 2005.

Radioactivity concentration in liquid and solid phases of scale and sludge

generated in the petroleum industry. Journal of Environmental Radioactivity, 81,

47-54.

Gholam, R.C. and Dariush, M., 2013. Theoretical and experimental investigation of

desalting and dehydration of crude oil by assistance of ultrasonic irradiation.

Ultrasonics Sonochemistry, 20, 378-385.

Ghosh, S. and Coupland, J.N., 2008. Factors affecting the freeze-thaw stability of

emulsions. Food Hydrocolloids, 22, 105-111.

Ghosh, S. and Rousseau, D., 2009. Freeze-thaw stability of water-in-oil emulsions. Journal

of Colloid and Interface Science, 339, 91-102.

Gorn;alves, C. and Alpenurada, M.F., 2005. Assessment of pesticide contamination in

soil samples from an intensive horticulture area, using ultrasonic extraction and

gas chromatography-mass spectrometry. Talanta, 65, 1179-1189.

Gonder, Z.B., Kaya, Y., Vergili, I., and Barlas, H., 2010. Optimization of filtration

conditions for CIP wastewater treatment by nanofiltration process using Taguchi

approach. Separation and Purification Technology, 70(3), 265-273.

Grasso, D., Subramaniam, K., Pignatello, J.J., Yang, Y., and Ratte, D., 2001. Micellar

desorption of polynuclear aromatic hydrocarbons from contaminated soil.

Colloids and Surfaces A: Physicochemical and Engineering Aspects, 194, 65-74.

Greg, M.H., Robert, A.H., and Zdenek, D., 2004. Paraffinic sludge reduction in crude oil

storage tanks through the use of shearing and resuspension. Acta Montanistica

Slovaca, 9, 184-188.

Guo, S.H., Li, G., Qu, J.H., and Liu, X.L., 2011. Improvement of acidification on

dewaterability of oily sludge from flotation. Chemical Engineering Journal, 168,

746-751.

233

Gussoni, M., Greco, F., Bonazzi, F., Vezzoli, A., Botta, D., Dotelli, G., Sora, I.N.,

Pelosato, R.,and Zetta, L., 2004. 1H NMR spin-spin relaxation and imaging in

porous system: an application to the morphological study of white Portland

cement during hydration in the presence of organics. Magnetic Resonance

Imaging, 22, 877-889.

Hahn, W.J., 1994. High-temperature reprocessing of petroleum oily sludges. SPE

Production & Facilities, 9, 179-182.

Hamad, A., Al-Zubaidy, E., and Fayed, M.E., 2005. Used lubricating oil recycling using

hydrocarbon solvents. Journal of Environmental Management, 74, 153-159.

Haritash, A.K. and Kaushik, C.P., 2009. Biodegradation aspects of polycyclic aromatic

hydrocarbons (PAHs): a review. Journal of Hazardous Materials, 169, 1-15.

He, G. and Chen, G., 2002. Lubricating oil sludge and its demulsification. Drying

Technology, 20, 1009-1018.

Heidarzadeh, N., Gitipour, S., and Abdoli, M.A., 2010. Characterization of oily sludge

from a Tehran oil refinery. Waste Management & Research, 28, 921-927.

Hejazi, R.F. and Husain, T., 2004. Landfarm performance under arid conditions. 2.

Evaluation of parameters. Environmental Science & Technology, 38, 2457-2469.

Hejazi, R.F. and Husain, T., 2004. Landfarm performance under arid conditions. 1.

Conceptual framework. Environmental Science & Technology, 38, 2449-2456.

Hejazi, R.F., Husain, T., and Khan, F.I., 2003. Landfarming operation of oily sludge in

arid region - human health risk assessment. Journal of Hazardous Materials, B99,

287-302.

Hu, G., Li, J., and Hou, H.B., 2015. A combination of solvent extraction and freeze thaw

for oil recovery from petroleum refinery wastewater treatment pond sludge.

Journal of Hazardous Materials, 283, 832-840.

234

Hu, G.J., Li, J.B., and Zeng, G.M., 2013. Recent development in the treatment of oily

sludge from petroleum industry: A review. Journal of Hazardous Materials, 261,

470-490.

Huang, Q., Han, X., Mao, F., Chi, Y., and Yan, J., 2014. A model for predicting solid

particle behavior in petroleum sludge during centrifugation. Fuel, 117, 95-102.

Isahak:, W.N.R.W., Hisham, M.W.M., Yarmo, M.A., and Hin, T.Y., 2012. A review on

bio-oil production from biomass by using pyrolysis method. Renewable and

Sustainable Energy Reviews, 16, 5910-5923.

Janbandhu, A. and Fulekar, M.H., 2011. Biodegradation of phenanthrene using adapted

microbial consortium isolated from petrochemical contaminated environment.

Journal of Hazardous Materials, 187, 333-340.

Jean, D.S. and Lee, D.J., 1999. Expression deliquoring of oily sludge from a petroleum

refinery plant. Waste Management, 19, 349-354.

Jean, D.S., Lee, D.J., and Wu, J.C.S., 1999. Separation of oil from oily sludge by

freezing and thawing. Water Research, 33, 1756-1759.

Jin, Y.Q., Zheng, X.Y., Chu, X.L., Chi, Y.J., Yan, H., and Cen, K.F., 2012. Oil recovery

from oil sludge through combined ultrasound and thermochemical cleaning

treatment. Industrial & Engineering Chemistry Research, 51, 9213-9217.

Jing, G., Luan, M., Han, C., Chen, T., and Wang, H., 2012. An effective process for

removing organic compounds from oily sludge using soluble metallic salt.

Journal of Industrial and Engineering Chemistry, 18, 1446-1449.

Jing, G.L. , Luan, M.M., Chen, T.T. , and Han, C.J., 2011. An effective process for

removing organic compounds from oily sludge. Journal of the Korean Chemical

Society, 55, 842-845.

John, F.S. and Joseph, F.R.J., 2008. On-column precipitation and re-dissolution of

asphaltenes in petroleum residua. Fuel, 87, 165-176.

235

Kar, Y., 2011. Co-pyrolysis of walnut shell and tar sand in a fixed-bed reactor.

Bioresource Technologies, 102, 9800-9805.

Karamalidis, A.K. and Voudrias, E.A., 2007. Cement-based stabilization/solidification of

oil refinery sludge: Leaching behavior of alkanes and P AHs. Journal of

Hazardous Materials, 148, 122-135.

Karamalidis, A.K. and Voudrias, E.A., 2007. Release of Zn, Ni, Cu, soi- and cr0i- as

a function of pH from cement-based stabilized/solidified refinery oily sludge and

ash from incineration of oily sludge. Journal of Hazardous Materials, 141, 591-

606.

Karayildirim, T. Yanik, J., Yuksel, M., and Bockhorn, H., 2006. Characterisation of

products from pyrolysis of waste sludges. Fuel, 85, 1498-1508.

Karayildirim, T., Yanik, J., Yuksel, M., and Bockhorn, H., 2006. Characterisation of

products from pyrolysis of waste sludges. Fuel, 85, 1498-1508.

Katsoura, M.H., Polydera, A.C., Katapodis, P., Kolisis, F.N., and Stamatis, H., 2007.

Effect of different reaction parameters on the lipase-catalyzed selective acylation

of polyhydroxylated natural compounds in ionic liquids. Process Biochemistry, 42,

1326-1334.

Khan, F.I., Husain, T., and Hejazi, R., 2004. An overview and analysis of site

remediation technologies. Journal of Environmental Management, 71, 95-122.

Kheireddine, H.A., El-Halwagi, M.M., and Elbashir, N.O., 2013. A property-integration

approach to solvent screening and conceptual design of solvent-extraction systems

for recycling used lubricating oils. Clean Technologies and Environmental Policy,

15, 35-44.

Ki, O.L., Kurniawan, A., Lin, C.X., Ju, Y.-H., and Ismadji, S., 2013. Bio-oil from cassava

peel: a potential renewable energy source. Bioresource Technology, 145, 157-161.

236

Kim, Y. and Parker, W., 2008. A technical and economic evaluation of the pyrolysis of

sewage sludge for the production ofbio-oil. Bioresouce Technology, 99, 1409-

1416.

Kim, Y.U. and Wang, M.C. , 2003 . Effect of ultrasound on oil removal from soils.

Ultrasonics Sonochemistry, 41, 539-542.

Kralova, I. , Sjoblom, J., 0ye, G., Simon, S., Grimes, B.A., Paso, K., 2011. Heavy crude

oils/particle stabilized emulsions. Advances in Colloid and Interface Science, 169,

106-127.

Kriipsalu, M., Marques, M. , and Maastik, A., 2008. Characterization of oily sludge from

a wastewater treatment plant flocculation-flotation unit in a petroleum refinery

and its treatment implications. Journal of Material Cycles and Waste

Management, 10, 79-86.

Kriipsalu, M. , Marques, M., Nammari, D.R., and Hogland, W., 2007. Biotreatrnent of

oily sludge: The contribution of amendment material to the content of target

contaminants, and the biodegradation dynamics. Journal of Hazardous Materials,

148, 616-622.

Kuo, C.H. and Lee, C.L., 2010. Treatment of oil/water emulsions using seawater-assisted

microwave irradiation. Separation Science and Technology, 74, 288-293.

Kuyukina, M.S., Ivshina, L.B. , Makarov, S.O., Litvinenko, L.V. , Cunningham, C.J., and

Philp, J.C., 2005. Effect ofbiosurfactants on crude oil desorption and

mobilization in a soil system. Environment International, 31 , 155-161.

Kwah, T.H., Maken, S., Lee, S. , Park, J.W., Min, B.R., and Yoo, Y.D., 2006.

Environmental aspects of gasification of Korean municipal solid waste in a pilot

plant. Fuel, 85, 2012-2017.

Lazar, I. , Dobrota, S., Voicu, A., Stefanescu, M., Sandulescu, L., and Petrisor, I.G., 1999.

Microbial degradation of waste hydrocarbons in oily sludge from some Romanian

oil fields. Journal of Petroleum Science and Engineering, 22, 151-160.

237

Lehmann, J. and Rondon, M., 2006.Bio-char soil management on highly weathered soils

in the humid tropics, in: N. Uphoff, A.S. Ball, C. Palm, E. Fernandes, J. Pretty, H.

Herren, P. Sanchez, 0. Husson, N. Sanginga, M. Laing, J. Thies (Eds.) Biological

Approaches to Sustainable Soil Systems, CRC Press, Taylor & Francis Group,

Boca Raton, pp. 518-519.

Leonard, S.A. and Stegemann, J.A., 2010. Stabilization/solidification of petroleum drill

cuttings. Journal of Hazardous Materials, 174, 463-472.

Li, C.T., Lee, W.J., Mi, H.H., and Su, C.C., 1995. PAH emission from the incineration of

waste oily sludge and PE plastic mixtures. Science of the Total Environment, 170,

171-183.

Li, J.B., Song, X.Y., Hu, G.J., and Thring, R.W., 2013. Ultrasonic desorption of

petroleum hydrocarbons from crude oil contaminated soils. Journal of

Environmental Science and Health, Part A: Toxic/Hazardous Substances and

Environmental Engineering, 48 (11), 1378-1389.

Li, X., He, L., Wu, G., Sun, W., Li, H., and Sui, H., 2012. Operational parameters,

evaluation methods, and fundamental mechanisms: Aspects of nonaqueous

extraction of bitumen from oil sands. Energy & Fuels, 26, 3553-3563.

Liang, J., Zhao, L., Du, N., Li, H., and Hou, W., 2014. Solid effect in solvent extraction

treatment of pre-treated oily sludge. Separation and Purification Technology,

130, 28-33

Lim, M.H., Kim, S.H., Kim, Y.U., and Khim, J., 2007. Sonolysis of chlorinated

compounds in aqueous solutions. Ultrasonics Sonochemistry, 14, 93-98.

Lima, T.M.S., Fonseca, A.F., Leiio, B.A., Mounteer, A.H., T6tola, M.R., and Borges,

A.C., 2011. Oil recovery from fuel oil storage tank sludge using biosurfactants.

Journal of Bioremediation & Biodegradation, 2, 1-5.

238

Lima, T.M.S., Procopio, L.C., Brandao, F.D., Leao, B.A., T6tola, M.R. , and Borges,

AC. , 2011. Evaluation of bacterial surfactant toxicity towards petroleum

degrading microorganisms. Bioresource Technology, 102, 2957-2964.

Lin, C. , He, G. , Dong, C., Liu, H., Xiao, G., and Liu, Y., 2008. Effect of oil phase

transition on freeze/thaw-induced demulsification of water-in-oil emulsions.

Langmuir, 24, 5291-5298.

Lin, G. , He, G., Li, X. , Peng, L. , Dong, S., Gu, S., and Xiao, G. , 2007. Freeze/thaw

induced demulsification of water-in-oil emulsions with loosely packed droplets.

Separation and Purification Technology, 56, 175-183.

Liu, J., Jiang, X. , and Han, X. , 2011. Devolatilization of oil sludge in a lab-scale

bubbling fluidized bed. Journal of Hazardous Materials , 185, 1205-1213.

Liu, J., Jiang, X., Zhou, L., Han, X. , and Cui, Z., 2009. Pyrolysis treatment of oil sludge

and model-free kinetics analysis. Journal of Hazardous Materials, 161 , 1208-

1215.

Liu, J., Jiang, X. , Zhou, L. , Wang, H. , and Han, X., 2009. Co-firing of oil sludge with

coal-water slurry in an industrial internal circulating fluidized bed boiler. Journal

of Hazardous Materials, 167, 817-823.

Liu, W., Tian, K., Jiang, H., Zhang, X. , and Yang, G. , 2013. Preparation ofliquid chemical

feedstocks by co-pyrolysis of electronic waste and biomass without formation of

polybrominated dibenzo-p-dioxins. Bioresource Technology, 128, 1-7.

Liu, W.X., Luo, Y.M. , Teng, Y., Li, Z.G. , and Ma, L.Q. , 2010. Bioremediation of oily

sludge-contaminated soil by stimulating indigenous microbes. Environmental

Geochemistry and Health, 32, 23-29.

Liu, W.X. , Wang, X.B. , Wu, L.H. , Chen, M.F. , Tu, C., Luo, Y.M., and Christie, P. , 2012.

Isolation, identification and characterization of Bacillus amyloliquefaciens BZ-6,

a baterial isolate for enhancing oil recovery from oily sludge. Chemosphere, 87,

1105-1110.

239

Long, X., Zhang, G., Shen, C., Sun, G., Wang, R., Yin, L., and Meng, Q., 2013.

Application of rhamnolipid as a novel biodemulsifier for destabilizing waste

crude oil. Bioresource Technology, 131, 1-5.

Machin-Ramirez, C., Okoh, A.I., Morales, D., Mayolo-Deloisa, K., Quintero, R., Trejo-

Hernandez, M.R., 2008. Slurry-phase biodegradation of weathered oily sludge

waste. Chemosphere, 70, 737-744.

Maga, S., Goetz, F., and Durlak, E., 2003. Operational test report (OTR): On-site

degradation of oily sludge in a ten-thousand gallon sequencing batch reactor at

Navsta Pearl Harbor, HI, in: Naval Facilities Engineering Command, Port

Hueneme, California.

Mahdi, K., Gheshlaghi, R., Zahedi, G., and Lohi, A., 2008. Characterization and

modeling of a crude oil desalting plant by a statistically designed approach.

Journal of Petroleum Science and Engineering, 61, 116-123.

Majumdar, S., Guha, AK., and Sirkar, K.K., 2002. Fuel oil desalting by hydrogel hollow

fiber membrane. Journal of Membrane Science, 202, 253-256.

Makkar, R.S. and Rockne, K.J., 2003. Comparison of synthetic surfactants and

biosurfactants fin enhancing biodegradation of polycyclic aromatic hydrocarbons.

Environmental Toxicology and Chemistry, 22, 2280-2292.

Malviya, R. and Chaudhary, R., 2006. Factors affecting hazardous waste

solidification/stabilization: a review. Journal of Hazardous Materials, 13 7, 267-

276.

Marin, J.A., Hernandez, T., and Garcia, C., 2005. Bioremediation of oil refinery sludge

by landfarming in semiarid conditions: Influence on soil microbial activity.

Environmental Research, 98, 185-195.

Marin, J.A., Moreno, J.L., Hernandez, T., and Garcia, C., 2006. Bioremediation by

composting of heavy oil refinery sludge in semiarid conditions. Biodegradation,

17, 251-261.

240

Martinez, J.D. , Yeses, A. , Mastral, A.M., Murillo, R. , Navarro, M.V. , and Puy, N ., 2014.

Copyrolysis of biomass with waste tyres: upgrading of liquid bio-fuel. Fuel

Processing Technology, 119, 263-271.

Mason, T.J. , 2007. Sonochemistry and environment-Providing a "green" link between

chemistry, physics and engineering. Ultrasonics Sonochemistry, 14, 476-483.

Mater, L. , Sperb, R.M. , Madureira, L. , Rosin, A. , Correa, A. , and Radetski , C.M. , 2006.

Proposal of a sequential treatment methodology for the safe reuse of oil sludge-

contaminated soil. Journal of Hazardous Materials , B136, 967-971.

Mazlova, E.A. and Meshcheryakov, S.V., 1999. Ecological characteristics of oil sludges.

Chemistry and Technology of Fuels and Oils, 35, 49-53.

Mazzarino, I. and Piccinini, P. , 1999. Photocatalytic oxidation of organic acids in

aqueous media by a supported catalyst. Chemical Engineering Science, 54, 3107-

3111.

Menendez, J.A. , Dominguez, A., Inguanzo, M. , and Pis, J.J., 2004. Microwave pyrolysis

of sewage sludge: analysis of the gas fraction. Journal of Analytical and Applied

Pyrolysis, 71 , 657-667.

Meyer, D.S., Brons, G.B., Perry, R. , Wildemeersch, S.L.A., and Kennedy, R.J., 2006. Oil

tank sludge removal method, United States Patent, US 2006/0042661 Al.

Mishra, S., Jyot, J. , Kuhad, R.C. , and Lal, B., 2001. In situ bioremediation potential of an

oily sludge-degrading bacterial consortium. Current Microbiology, 43 , 328-335 .

Mishra, S. , Lal, B., Jyot, J., Rajan, S., Khanna, S., and Kuhad, R., 1999. Field study: In

situ bioremediation of oily sludge contaminated land using "oil zapper". in:

Hazardous and Industrial Wastes: proceedings of mid-atlantics industrial waste

conference, pp. 177-186.

Mohan, D., Pittman Jr, C.U., Bricka, M. , Smith, F. , Yancey, B. , Mohammad, J. , Steele,

P.H. , Alexandre-Franco, M.F. , Gomez-Serrano, V., and Gong, H., 2007. Sorption

241

of arsenic, cadmium, and lead by chars produced from fast pyrolysis of wood and

bark during bio-oil production. Journal of Colloid and Interface Science, 310, 57-

73.

Mohan, S.V. and Chandrasekhar, K., 2011 . Self-induced bio-potential and graphite

electron accepting conditions enhances petroleum sludge degradation in bio-

electrochemical system with simultaneous power generation. Bioresouce

Technology, 102, 9532-9541.

Molt6, J. , Barneto, A.G. , Ariza, J., and Conesa, J.A. , 2013. Gas production during the

pyrolysis and gasification of biological and physico-chemical sludges from oil

refinery. Journal of Analy tical and Applied Pyrolysis , 103, 167-172.

Moosai, R. and Dawe, R.A. , 2003 . Gas attachment of oil drophlets for gas flotation for

oily wastewater cleanup. Separation and Purification Technology, 33 , 303-314.

Moses, J.R. , Menaka, P., and Tapasya, M.N. , 2003. Sustainable landfilling of oily sludge,

in: Proceedings of the Workshop on Sustainable Landfill Management, Chennai,

India, pp. 225-233.

Mrayyan, B. and Battikhi, M.N., 2005 . Biodegradation of total organic carbon (TOC) in

Jordanian petroleum sludge. Journal of Hazardous Materials, 120, 127-134.

Mukred, A.M., Hamid, A.A., Hamzah, A. , and Yusoff, W.M.W., 2008 . Development of

three bacterial consortium for the bioremediation of crude petroleum-oil in

contaminated water. International Journal of Biological Sciences, 8, 73-79.

Mulligan, C.N. , 2005. Environmental applications for biosurfactants. Environmental

Pollution, 133, 183-198.

Mulligan, C.N., 2009. Recent advances in the environmental applications of

biosurfactants. Current Opinion in Colloid & Interface Science, 14, 372-378.

Mulligan, C.N. , Yong, R.N. , and Gibbs, B.F., 2001. Surfactant-enhanced remediation of

contaminated soil: A review. Engineering Geology, 60, 371-380.

242

Na, S., Park, Y., Hwang, A., Ha, J., Kim, Y., and Khim, J., 2007. Effect of ultrasound on

surfactant-aided soil washing, Japanese Journal of Applied Physics, 46, 4775-

4778.

Nahmad, D.G., 2012. Method to recover crude oil from sludge or emulsion, United

States Patent, US 8,197,667 B2.

Naik, B.S., Mishra, I.M., and Bhattacharya, S.D., 2011. Biodegradation of total petroleum

hydrocarbons from oily sludge. Bioremediation Journal, 15(3), 140-147.

Neuma de Castro Dantas, T., Avelino Dantas Neto, A., and Ferreira Moura, E., 2001.

Microemulsion systems applied to breakdown petroleum emulsions. Journal of

Petroleum Science and Engineering, 32, 145-149.

Nii, S., Kikumoto, S., and Tokuyama, H., 2009. Quantitative approach to ultrasonic

emulsion separation. Ultrasonics Sonochemistry, 16, 145-149.

Noordman, W.H., Wachter, J.J.J., Boer, G.J.D., and Janssen, D.B., 2002. The

enhancement by biosurfactants of hexadecane degradation by Pseudomonas

aeruginosa varies with substrate availability. Journal of Biotechnology, 94, 195-

212.

Oh, K. and Deo, M.D., 2011.Yield behavior of gelled waxy oil in water-in-oil emulsion

at temperatures below ice formation. Fuel, 90, 2113-2117.

Olivera, N.L., Commendatore, M.G., Moran, A.C., and Esteves, J.L., 2000.

Biosurfactant-enhanced degradation of residual hydrocarbons from ship bilge

wastes. Journal of Industrial Microbiology and Biotechnology, 25, 70-73.

Onal, E. , Uzun, B.B., and Piitiin, A.E., 2014. Bio-oil production via co-pyrolysis of

almond shell as biomass and high density polyethylene. Energy Conversion

Management, 78, 704-710.

243

Otidene, R.S.D.R., Renato, F.D., Marta, M.M.B.D., Marcia, M.L.D., and Valdinete,

L.D.S., 2010. Oil sludge treatment by photocatalysis applying black and white

light. Chemical Engineering Journal, 157, 80-85.

Ouyang, W., Liu, H., Murygina, V. , Yu, Y.Y., Xiu, Z.D., and Kalyuzhnyi, S., 2005.

Comparison of bio-augmentation and composting for remediation of oily sludge:

a field-scale in China. Process Biochemistry, 40, 3763-3768.

Patel, N.K. and Salanki, S.S., 1999. Hazardous, toxic and mixed chemical waste's

disposal by incineration, in: the 15th International Conference on Solid Waste

Technology and Management, Phildelphia, USA.

Pekdemir, T., Copur, M., and Urum, K., 2005 . Emulsification of crude oil-water system

using biosurfactant. Process Safety and Environmental Protection, 83, 38-46.

Peng, W., Jiao, H., Shi, H. , and Xu, C. , 2012. The application of emulsion liquid

membrane process and heat-induced demulsification for removal of pyridine from

aqueous solutions. Desalination, 286, 372-378.

Peter, D., 2001. Sonolytic degradation of volatile pollutants in natural ground water:

conclusions from a model study. Ultrasonics Sonochemistry, 8, 221-226.

Pilli, S., Bhunia, P., Yan, S., LeBlanc, R.J., Tyagi, R.D., and Surampalli, R.Y., 2011.

Ultrasonic pretreatment of sludge: a review. Ultrasonics Sonochemistry, 18, 1-18.

Pinheiro, B.C.A. and Rolanda J.N.F., 2009. Processing of red ceramics incorporated with

encapsulated petroleum waste. Journal of Materials Processing Technology, 209,

5606-5610.

Pinto, F., Paradela, F., Gulyurtlu, I., and Ramos, A.M., 2013. Prediction of liquid yields

from the pyrolysis of waste mixtures using response surface methodology. Fuel

Processing Technology, 116, 271-283.

Pootakham, T. and Kumar, A. , 2010. Bio-oil transport by pipeline: a techno-economic

assessment. Bioresource Technology, 101 (18), 7137-7143.

244

Powell, S.M., Paul, M., Harvey, P.M., Stark, S.J. , Snipe, I., and Riddle, J.M., 2007.

Biodegradation of petroleum products in experimental plots in Antarctic marine

sediments is location dependent. Marine Pollution Bulletin, 54, 434-440.

Prak, D.J.L. and Pritchard, P.H., 2002. Degradation of polycyclic aromatic hydrocarbons

disolved in Tween 80 surfactant solutions by Sphingomonas paucimobilis EPA

505. Canadian Journal of Microbiology, 48, 151-158.

Punnaruttanakun, P., Meeyoo, V., Kalambaheti, C., Rangsunvigit, P., Rirksomboon, T.,

and Kitiyanan, B., 2003. Pyrolysis of API separator sludge. Journal of Analytical

and Applied Pyrolysis, 68-69, 547-560.

Rahman, K.S.M., Banat, I.M., Rahman, T.J., Thayumanavan, T., and

Lakshmanaperumalsamy, P., 2002. Bioremediation of gasoline contaminated soil

by a bacterial consortium amended with poultry litter, coir pith and rhamnolipid

biosurfactant. Bioresource Technology, 81, 25-32.

Rahman, K.S.M., Rahman, T.J., Kourkoutas, Y., Petsas, I., Marchant, R., and Banat,

I.M., 2003. Enhanced bioremediation of n-alkane in petroleum sludge using

bacterial consortium amended with rhamnolipid and micronutrients. Bioresource

Technology, 90, 159-168.

Rajakovic, V. and Skala, D., 2006. Separation of water-in-oil emulsions by freeze/thaw

method and microwave radiation. Separation and Purification Technology, 49,

192-196.

Ramaswamy, D., Kar, D.D., and De, S., 2007. A study on recovery of oil from sludge

containing oil using froth flotation. Journal of Environmental Management, 85,

150-154.

Reddy, M.V., Devi, M.P., Chandrasekhar, K., Goud, R.K., and Mohan, S.V., 2011.

Aerobic remediation of petroleum sludge through soil supplementation: Microbial

community analysis. Journal of Hazardous Materials, 197, 80-87.

245

Rincon, J., Canizares, P., and Garcia, M.T., 2005. Regeneration of used lubricant oil by

polar solvent extraction. Industrial & Engineering Chemistry Research, 44, 4343-

4379.

Rivas, F.J., 2006. Polycyclic aromatic hydrocarbons sorbed on soils: A short view of

chemical oxidation based treatments. Journal of Hazardous Materials, B 138,

234-251.

Robertson, S.J., McGill, W.B., Massicotte, H.B., and Rutherford, P.M., 2007. Petroleum

hydrocarbon contamination in boreal forest soils: a mycorrhizal ecosystems

perspective.

Robinson, J.P., Snape, C.E., Kingman, S.W., and Shang, H., 2008. Thermal desorption

and pyrolysis of oil contaminated drill cuttings by microwave heating. Journal of

Analytical and Applied Pyrolysis, 81, 27-32.

Roldan, C.T., Castorena, C.G., Zapata, P.I., Reyes, A.J., and Olguin, L.P., 2012. Aerobic

biodegradation of sludge with high hydrocarbon content generated by a Mexican

natural gas processing facility. Journal of Environmental Management, 95, 93-98.

Roldan-Carrillo, T., Castorena-Cortes, G., Zapata-Pefiasco, I., Reyes-Avila, J., Olguin-

Lora, P ., 2012. Aerobic biodegradation of sludge with high hydrocarbon content

generated by a Mexican natural gas processing facility. Journal of Environmental

Management, 95, S93-S98.

Ron, E.Z. and Rosenberg, E., 2002. Biosurfactants and oil bioremediation. Current

Opinion in Biotechnology, 13, 249-252.

Rondon, M., Bouriat, P., and Lachaise, J., 2006. Breaking of water-in-oil emulsions. 1.

Physicochemical phenomenology of demulsifier action. Energy and Fuels, 20,

1600-1604.

Saari, E., Periimaki, P., and Jalonen, J. , 2007. A comparative study of solvent extraction

of total petroleum hydrocarbons in soil. Microchimica Acta, 158, 261-268.

246

Samanya, J. , Hornung, A., Apfelbacher, A., and Vale, P., 2012. Characteristics of the upper

phase of bio-oil obtained from co-pyrolysis of sewage sludge with wood, rapeseed

and straw. Journal of Analytical and Applied Pyrolysis, 94, 120-125.

Sankaran, S., Pandey, S., and Sumathy, K., 1998. Experimental investigation on waste

heat recovery by refinery oil sludge incineration using fluidised-bed technique.

Journal of Environmental Science and Health , Part A, 33, 829-845.

Santiago, B., Susannah, F., Kebba, J., Nathaniel, L., Jose, M., Elise, P ., Jennifer, R.V.,

David, R. , Jesse, S., Samir, S., and Pratikshe, S., 2002. Oil: A life cycle analysis

of its health and environmental impacts, in: R.E. Paul, S. Jesse (Eds.), The Center

for Health and the Global Environment, Harvard Medical School, Boston, pp. 27-

29.

Scala, F . and Chirone, R., 2004. Fluidized bed combustion of alternative solid fuels.

Experimental Thermal and Fluid Science, 28, 691-699.

Schmidt, H. and Kaminsky, W., 2001 . Pyrolysis of oil sludge in a fluidised bed reactor.

Chemosphere, 45, 285-290.

Schwab, A.P ., Su, J ., Wetzel, S., Pekarek, S., and Banks, M.K. , 1999. Extraction of

petroleum hydrocarbons from soil by mechnical shaking. Environmental Science

& Technology, 33, 1940-1945.

Shang, H., Snape, C.E., Kingman, S.W., and Robinson, J.P. , 2006. Microwave treatment

of oil-contaminated North Sea drill cuttings in a high power multimode cavity.

Separation Science and Technology, 49, 84-90.

Shen, D.K. and Gu, S., 2009. The mechanism for thermal decomposition of cellulose and

its main products. Bioresource Technology, 100, 649~504.

Shen, L. and Zhang, D.K. , 2003 . An experimental study of oil recovery from sewage

sludge by low-temperature pyrolysis in a fluidised-bed. Fuel, 82, 465-472.

247

Shie, J. , Chang, C., Lin, J., Lee, D. , and Wu, C. , 2002. Use of inexpensive additives in

pyrolysis of oil sludge. Energy & Fuels, 16, 102-108.

Shie, J.L., Lin, J.P., Chang, C.Y., Lee, D.J., and Wu, C.H., 2003 . Pyrolysis of oil sludge

with additives of sodium and potassium compounds. Resources, Conservation

and Recycling, 39, 51-64.

Shie, J .L. , Lin, J .P., Chang, C.Y. , Shih, S.M., Lee, D.J., and Wu, C.H. , 2004. Pyrolysis of

oil sludge with additives of catalytic solid wastes. Journal of Analy tical and

Applied Pyrolysis, 71, 695-707.

Shie, J .L., Lin, J.P., Chang, C.Y. , Wu, C.H. , Lee, D.J., Chang, C.F. , and Chen, Y.H. ,

2004. Oxidative thermal treatment of oil sludge at low heating rates. Energy &

Fuels, 18, 1272-1281.

Shiva, H., 2004. A new electrokinetic technology for revitalization of oily sludge, in:

The Department of Building, Civil, and Environmental Engineering, Concordia

University, Montreal, Quebec, Canada, pp. 29.

Shuchi, V., Renu, B., and Vikas, P., 2006. Oily sludge degradation by bacteria from

Ankleshwar, India. International Biodeterioration & Biodegradation , 57, 207-

213 .

Smadar, A., Michal, G. , and Yoram, A. , 2001. Biodegradation kinetics of hydrocarbons

in soil during land treatment of oily sludge. Bioremediation Journal, 5,193-209.

Smith, P., Noonan, J. , and Thouin, G. , 2003 . Characteristics of spilled oils, fuels, and

petroleum products: 1. Composition and properties of selected oils, US EPA,

EPA/600/r-03/072, July 2003.

Song, W., Li, J ., Zhang, W., Hu, X., and L. Wang, 2012. An experimental study on the

remediation of phenanthrene in soil using ultrasound and soil washing.

Environmental Earth Sciences, 66, 1487-1496.

248

Song, Y., Tahmasebi, A., and Yu, J., 2014. Co-pyrolysis of pine sawdust and lignite in a

thermogravimetric analyzer and a fixed-bed reactor. Bioresource Technology, 174,

204-211.

Spiecker, P.M., Gawrys, K.L., and Kilpatrck, P.K., 2003. Aggregation and solubility

behavior of asphaltenes and their subfractions. Journal of Colloid and Interface

Science, 267, 178-193.

Stasiuk, E.N. and Schramm, L.L., 2001. The influence of solvent and demulsifier

additions on nascent froth formation during flotation recovery of Bitumen from

Athabasca oil sands. Fuel Processing Technology, 73, 95-110.

Straube, W.L., Nestler, C.C., Hansen, L.D., Ringleberg, D., Pritchard, P.J., and Jones-

Meehan, J., 2003. Remediation ofpolyaromatic hydrocarbons (PAHs) through

landfarming with biostimulation and bioaugmentation. Acta Biotechnologica, 2-3,

179-196.

Suleimanov, R.R., Gabbasova, I.M., and Sitdikov, R.N., 2005. Changes in the properties

of oily gray forest soil during biological reclamation. The Biological Bulletin, 32,

109-115.

Swamy, K.M. and Narayana, K.L., 2001 . Intensification ofleaching process by dual-

frequency ultrasound. Ultrasonics Sonochemistry, 8, 341-346.

Syal, S. and Ramamurthy, V., 2003. Characterization ofbiosurfactant synthesis in a

hydrocarbon utilizing bacterial isolate. Indian Journal of Microbiology, 43, 175-

180.

Tahhan, R.A., Ammari, T.G., Goussous, S.J. , Al-Shdaifat, H.I., 2011. Enhancing the

biodegradation of total petroleum hydrocarbons in oily sludge by a modified

bioaugmentation strategy. International Biodeterioration & Biodegradation, 65,

130-134.

Taiwo, E.A. and Otolorin, J.A., 2009. Oil recovery from petroleum sludge by solvent

extraction. Petroleum Science and Technology, 27, 836-844.

249

Tan, W., Yang, X., and Tan, X., 2007. Study on demulsification of crude oil emulsions

by microwave chemical method. Separation Science and Technology, 42, 1367-

1377.

Tang, J., Lu, X., Sun, Q., and Zhu, W., 2012. Aging effect of petroleum hydrocarbons in

soil under different attenuation conditions. Agriculture, Ecosystems &

Environment, 149, 109-117.

Tavassoli, T., Mousavi, S.M., Shojaosadati, S.A., and Salehizadeh, H., 2012. Asphaltene

biodegradation using microorganisms isolated from oil samples. Fuel, 93, 142-

148.

Tian, K., Liu, W., Qian, T., Jiang, H., and Yu, H., 2014. Investigation on the evolution of

N-containing organic compounds during pyrolysis of sewage sludge.

Environmental Science & Technology, 48, 10888-10896.

Trofimov, S.Y. and Rozanova, M.S., 2003. Transformation of soil properties under the

impact of oil pollution. Eurasian Soil Science, 36, S82-S87.

Urbina, R.H., 2003 . Recent developments and advances in formulations and applications

of chemical reagents. Mineral Processing and Extractive Metallurgy Review, 24,

139-182.

Urum, K. and Pekdemir, T., 2004. Evaluation ofbiosurfactants for crude oil

contaminated soil washing. Chemosphere, 57, 1139-1150.

Uzun, B.B., Piitiln, A.E., and Piitiln, E., 2006. Fast pyrolysis of soybean cake: product

yields and compositions. Bioresource Technology, 97, 569-576.

van Hamme, J.D., Odumeru, J.A., and Ward, O.P., 2000. Community dynamics of a

mixed-baterial culture growing on petroleum hydrocarbons in batch culture.

Canadian Journal of Microbiology, 46, 441-450.

van Oudenhoven, J.A.C.M., Cooper, G.R., Cricchi, G., Gineste, J., Potzl, P., and Martin,

D.E., 1995. Oil refinery waste, disposal methods and costs 1993 survey, in:

250

Conservation of Clean Air and Water in Europe (CONCA WE), Brussels, pp. 1-

39.

Vanessa, S.C., Emanuel, B.H., Franciele, M., Marilene, H.V., Flavio, A.O.C., Maria,

D.C.R.P., and Fatima, M.B., 2011. Biodegradation potential of oily sludge by

pure and mixed bacterial cultures. Bioresource Technology, 102, 1003-1010.

Vasudevan, N . and Rajaram, P., 2001. Bioremediation of oil sludge-contaminated soil.

Environment International, 26, 409-411.

Virkutyte, J., Sillanpaa, M., and Latostenmaa, P., 2002. Electrokinetic soil remediation -

critical overview. Science of The Total Environment, 289, 97-121.

Wang, R., Liu, J., Gao, F., Zhou, J., and Cen, K., 2012. The slurrying properties of slurry

fuels made of petroleum coke and petrochemical sludge. Fuel Processing

Technology, 104 (2012) 57-66.

Wang, X., Wang, Q.H., Wang, S.J., Li, F.S., and Guo, G.L., 2012. Effect of

biostimulation on community level physiological profiles of microorganisms in

field-scale biopiles composed of aged oil sludge. Bioresource Technology, 111,

308-315.

Wang, Z., Guo, Q., Liu, X., and Cao, C., 2007. Low temperature pyrolysis characteristics

of oil sludge under various heating conditions. Energy & Fuels, 21, 957-962.

Ward, 0., Singh, A., and van Hamme, J ., 2003. Accelerated biodegradation of petroleum

hydrocarbon waste. Journal of Industrial Microbiology & Biotechnology, 30,

260-270.

Ward, O.P. and Singh, A., 2003. Biodegradation of oil sludge, United States Patent, No.

6,652,752.

Weber W.J. and Kim, H.S., 2005. Optimizing contaminant desorption and bioavailability

in dense slurry systems. 1. Rheology, mechanical mixing, and P AH desorption.

Environmental Science & Technology, 39, 2267-2273.

251

Whang, L.M., Liu, P.W., Ma, C.C., and Cheng, S.S., 2008. Application ofbiosurfactants,

rhamnolipid, and surfactin, for enhanced biodegradation of diesel-contaminated

water and soil. Journal of Hazardous Materials, 151, 155-163.

Woo, S.H. and Park, J.M., 1999. Evaluation of drum bioreactor performance used for

decontamination of soil polluted with polycyclic aromatic hydrocarbons. Journal

of Chemical Technology and Biotechnology, 74, 937-944.

Xia, L.X., Lu, S.W., and Cao, G.Y., 2003. Demulsification of emulsions exploited by

enhanced oil recovery system. Separation Science and Technology, 38, 4079-

4094.

Xia, L.X., Lu, S.W., and Cao, G.Y., 2004. Salt-assisted microwave demulsification.

Chemical Engineering Communications, 191, 1053-1063.

Xiu, S. and Shahbazi, A., 2012. Bio-oil production and upgrading research: A review.

Renewable & Sustainable Energy Reviews, 16, 4406-4414.

Xu, N., Wang, W., Han, P., and Lu, X., 2009. Effects of ultrasound on oily sludge

deoiling. Journal of Hazardous Materials, 171, 914-917.

Yan, P., Lu, M., Yang, Q., Zhang, H.L., Zhang, Z.Z. , and Chen, R., 2012. Oil recovery

from refinery oily sludge using a rhamnolipid biosurfactant-producing

Pseudomonas. Bioresource Technology, 116, 24-28.

Yang, L., Nakhla, G., and Bassi, A., 2005. Electro-kinetic dewatering of oily sludges.

Journal of Hazardous Materials, 125, 130-140.

Yang, S.Z., Jin, H.J., Wei, Z., He, R.X. , Ji, Y.J., Li, X.M., and Shao, S.P., 2009.

Bioremediation of oil spills in cold environments: A review. Pedosphere, 19(3),

371-381.

Yang, X., Tan, W., and Bu, Y., 2009. Demulsification of asphaltenes and resins

stabilized emulsions via the freeze/thaw method. Energy & Fuels, 23, 481-486.

252

Yang, X., Tan, W., and Tan, X. , 2009. Demulsification of crude oil emulsion via

ultrasonic chemical method. Petroleum Science and Technology, 27, 2010-2020.

Ye, G., Lu, X., Han, P., Peng, F., Wang, Y., and Shen, X., 2008. Application of

ultrasound on crude oil pretreatment. Chemical Engineering and Processing:

Process Intensification, 47, 2346-2350.

Yerushalmi, L. , Rocheleau, S., Cimpoia, R. , Sarrazin, M., Sunahara, G., and Peisajovich,

A. , 2007. Enhanced biodegradation of petroleum hydrocarbon in contaminated

soil. Bioremediation Journal, 7, 37-51.

Zhang, J. , Li, J.B., Thing, R.W. , and Hu, G.J. , 2013 . Investigation of impact factors on

the treatment of oily sludge using a hybrid ultrasonic and Fenton ' s reaction

process, In: Proceedings of the International Conference on Environmental

Pollution and Remediation, Toronto, Ontario, Canada, 2013, pp. 76.

Zhang, J., Li, J.B., Thring, R.W. , Hu, X. , and Song, X.Y., 2012. Oil recovery from

refinery oily sludge via ultrasound and freeze/thaw. Journal of Hazardous

Materials , 203, 195-203.

Zhang, S.P. , Yan, Y.J. , Li, Y.C., and Ren, Z.W ., 2005. Upgrading of liquid fuel from the

pyrolysis of biomass. Bioresouce Technology, 96, 545-550.

Zhang, W., Li, J., Huang, G. , Song, W., and Huang, Y. , 2011. An experimental study on

the bio-surfactant-assisted remediation of crude oil and salt contaminated soils.

Journal of Environmental Science and Health , Part A, 46, 306-313.

Zhang, X. , Li, J., Thring, R. , and Huang, Y. , 2010. Surfactant enhanced biodegradation

of petroleum hydrocarbons in oil refinery tank bottom sludge. Journal of

Canadian Petroleum Technology, 49, 34-39.

Zhang, X. , Wang, H. , He, L. , Lu, K. , Sarmah, A. , Li J. , Bolan, N., Pe, J., and Huang, H. ,

2013. Using biochar for remediation of soils contaminated with heavy metals and

organic pollutants. Environmental Science and Pollution Research, 20, 84 72-8483 .

253

Zhao, D., Xue, J., Li, S., Sun, H., and Zhang, Q., 2011. Theoretical analyses of thermal

and economical aspects of multi-effect distillation desalination dealing with high-

salinity wastewater. Desalination, 273, 292-298.

Zhou, L., Jiang, X., and Liu, J., 2009. Characteristics of oily sludge combustion in

circulating fluidized beds. Journal of Hazardous Materials, 170, 175-179.

Zubaidy, E.A.H. and Abouelnasr, D.M., 2010. Fuel recovery from waste oily sludge

using solvent extraction. Process Safety and Environmental Protection, 88, 318-

326.

Zuo, W., Jin, B., Huang, Y., and Sun, Y., 2014. Characterization of top phase oil obtained

from co-pyrolysis of sewage sludge and poplar sawdust. Environmental Science

and Pollution Research, 21 , 9717-9726.

254

APPENDIX Appendix I

.-~..----.---,--------------------------------~1 1.00

0.75 J!l 5' 0.50

0.25

0.00 -11---_J

1.50

1.25

Io~ 100lJ 0~ 0~ \ 000 1.50

~ 1.00 >

0.50

0 OOJ ....... J/-,.,-1'--2.00

1.50

~ 1 00 >

0.50

0 00

a

b

·~~-C

LlmulwllilLLLLLw~~----·---··-·-·····--··-d

' 15- 10 15 20 25 30 35 Minutes

GC profiles of different hydrocarbon samples, (a) original oily sludge, (b) recovered oil, (c) fresh crude oil, (d) diesel.

255