192
Downstream processing of lipids and lipases from thraustochytrids by Avinesh Reddy Byreddy (M.Sc., M. Phil) Submitted in fulfilment of the requirements for the degree of Doctor of Philosophy Deakin University September, 2015

Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Downstream processing of lipids and

lipases from thraustochytrids

by

Avinesh Reddy Byreddy (M.Sc., M. Phil)

Submitted in fulfilment of the requirements for the degree of

Doctor of Philosophy

Deakin University

September, 2015

Page 2: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),
sfol
Retracted Stamp
Page 3: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),
sfol
Retracted Stamp
Page 4: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Acknowledgements

I would like to express my immense gratitude to Principal supervisor Professor Colin

James Barrow for providing me with this exciting opportunity to undertake the most

rewarding and challenging research under your supervision. Thank you very much for

the belief you had in me and the encouragement you have given me over the last few

years will stay with me.

I am grateful to my Principal co-supervisor, Associate Professor Munish Puri for

providing direction, designing experiments, valuable discussions surrounding

achieved results, and getting publications from the carried research work. I also thank

him for his scientific advices, his constructive feedback on my writing and

encouragement continually given over the last few years.

I would also like to thank Dr Madhusudan Rao Nalam (Senior Scientist, Centre for

Cellular and Molecular Biology, Hyderabad, India) for his insightful advices in

phospholipase research.

I would like to acknowledge an award of Deakin university postgraduate scholarship

(DUPRS), Deakin University, Australia for supporting my PhD research.

The journey made in the last 4 years would not have been possible without the support

of many kind hearted and good witted people around me. To my dear, “Super chooks”,

for your care and emotional support in all the good and bad times. I wish to express

my thanks to my fellow students and other postdocs from Barrow group. I am thankful

to the Technical staff at LES and Centre for Chemistry and Biotechnology for their

support and help.

Last but not least, I am extremely grateful to my dear wife Suchitra and my parents

Rama Mohan Reddy and Jayalakshmi and lovable sister Sowjanya for continues

support and encouragement throughout this study.

i

Page 5: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

List of publications

1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016), Bead milling for

lipid recovery from thraustochytrid cells and selective hydrolysis of

Schizochytrium DT3 oil using lipase. Bioresource Technology. 200, 464-469.

2. Avinesh R. Byreddy, Adarsha Gupta, Colin Barrow, Munish Puri (2015),

Comparison of cell disruption methods for improving lipid extraction from

thraustochytrid strains. Marine Drugs, 13, 5111-5127

3. Adarsha Gupta, Dilip Singh, Avinesh R. Byreddy, Tamilsevi Thyagrajan,

Shailedra Sonkar, Anshu S Mathur, Deepak K Tuli, Colin J Barrow, Munish

Puri (2015), Comparison of Australian and Indian thraustochytrids for omega-

3 fatty acids, enzymes and biodiesel production, Biotechnology Journal.

Accepted, 14th Oct.

Manuscripts Communicated/prepared

1. Avinesh R. Byreddy, Adarsha Gupta, Colin Barrow, Munish Puri (2015), A

quick colorimetric method for total lipid quantification in thraustochytrids.

New Biotechnology Journal.

2. Tushar R M, Avinesh R. Byreddy, Colin Barrow, NM Rao (2015),

Application of phospholipase A1 for concentration of polyunsaturated fatty

acids from fish oil.

Conference Proceedings

1. Avinesh Byreddy, Colin J Barrow and Munish Puri, Evaluating suitability of

bead milling on lipid recovery from marine microalgae for omega- 3 fatty acid

concentration. AAOCS meeting: 9-11 September 2015.

2. Avinesh Byreddy, Munish Puri, NM Rao, Colin Barrow, Application of

phospholipases for selective concentration of omega-3 fatty acids from fish oil.

Presented in 4th international conference on Novel Enzymes Organized by

Ghent University, Belgium 2014.

3. Munish Puri, Adarsha Gupta, Tamilselvi Thyagrajan, Avinesh Byreddy, Colin

J Barrow, Optimising lipid yields in microalgae screened from Australian

ii

Page 6: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

marine environment for producing next generation biofuels. SMBB

conference, Mexico 2014

4. Avinesh Byreddy, Munish Puri, Colin Barrow, Production and extraction of

lipases from the Schizochytrium sp. S31. AISRF conference, March 12-14,

2014

5. Avinesh Byreddy, Munish Puri, Colin Barrow, Screening and comparison of

cell disruption methods for the extraction of lipases from marine microbes.

Organised by IFM, Deakin University, Nov 2012.

iii

Page 7: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Sequences deposited in GenBank database

1. DT1 Ulkenia sp. KF682136

2. DT2 Ulkenia sp. KF682137

iv

Page 8: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Table of contents

Acknowledgements ......................................................................................................i

List of publications .....................................................................................................ii

Sequences deposited in GenBank database.............................................................iv

Table of contents.........................................................................................................v

List of figures .............................................................................................................ix

List of tables ..............................................................................................................xii

List of abbreviations................................................................................................xiv

Abstract ....................................................................................................................xviChapter 1: Introduction and Literature review ......................................................1

1.1. Introduction....................................................................................................2

1.2. Omega-3 FAs.....................................................................................................3

1.2.1. Health benefits of omega-3 FAs .................................................................4

1.2.2. Dietary sources of omega-3 FAs.................................................................5

1.2.3. Techniques used for the concentration of omega-3 FAs.............................7

1.2.4. Microalgae as an alternative source for omega-3 fatty acids......................9

1.3. Thraustochytrids ..............................................................................................10

1.3.1. Potential of Thraustochytrids ....................................................................12

1.3.2. Strain selection and sustainable lipid production......................................12

1.3.3. Estimation of lipid content and fatty acid profile .....................................13

1.4. Extraction of lipids from dry cell biomass ......................................................15

1.4.1. Selection of suitable extraction solvents ...................................................15

1.4.2. Lipid extraction from wet biomass ...........................................................19

1.4.3. Microalgae cell disruption for lipid extraction .........................................20

1.4.3. Effect of cell disruption methods on lipid extraction................................21

1.5. Lipases .............................................................................................................25

1.5.1. Marine lipases ...........................................................................................25

1.5.2. Isolation of lipase producing microbes and screening methods ...............26

1.5.3. Extraction of intracellular lipases .............................................................28

1.5.2. Lipase based methods for concentration of omega-3 fatty acids ..............28

1.6. Research aim and objectives............................................................................33

v

Page 9: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 2: A quick colorimetric method for total lipid quantification in thraustochytrids........................................................................................................34

2.1. Introduction..................................................................................................35

2.2. Materials and Methods ................................................................................36

2.2.1. Chemicals used .........................................................................................36

2.2.2. Cultivation of thraustochytrids..................................................................37

2.2.3. Standard curve preparation .......................................................................37

2.2.4. Quantification of thraustochytrid lipids by SPV.......................................37

2.2.5. Lipid extraction .........................................................................................38

2.3. Results and discussion .....................................................................................39

2.3.1. Biomass and lipid production ...................................................................39

2.3.2 Microalgae oil quantification by SPV method...........................................40

2.3.3. Lipid quantification in thraustochytrid whole cells via SPV method .......42

2.3.4. Validating the accuracy of the SPV method .............................................44

2.4. Conclusion .......................................................................................................46

Chapter 3: Comparison of cell disruption methods for improving lipid extraction from thraustochytrid strains.................................................................47

3.1. Introduction..................................................................................................48

3.2. Experimental Section.......................................................................................49

3.2.1. Chemicals..................................................................................................49

3.2.2. Strain selection and biomass production...................................................50

3.2.3. Lipid extraction from thraustochytrids by organic solvents .....................50

3.2.4. Cell disruption for lipid extraction............................................................51

3.5. Fatty acids methyl esters (FAMEs) analysis ...................................................53

3.3. Results and Discussion ....................................................................................53

3.3.1. Biomass and lipid production ...................................................................53

3.3.2. Lipid extraction from thraustochytrid by organic solvents .......................54

3.3.3. Comparison of lipid extraction methods...................................................56

3.3.4. Fatty acid composition of extracted lipid..................................................60

3.3.5. Prediction of biodiesel properties .............................................................63

3.3.6. Energy analysis .........................................................................................65

3.4. Conclusion .......................................................................................................66

Chapter 4: Optimised cell disruption methods for lipase extraction from Australian thraustochytrids ....................................................................................67

4.1. Introduction......................................................................................................68

4.2. Materials and methods.....................................................................................69

4.2.1. Chemicals..................................................................................................69vi

Page 10: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

4.2.2. Strain selection and cultivation .................................................................69

4.2.3. Enzymatic profiling of isolates .................................................................70

4.2.4. Lipase production......................................................................................70

4.2.5. Cell disruption methods for lipase release ................................................70

4.2.6. Microscopic observation of the native and disrupted thraustochytrid cells.............................................................................................................................71

4.2.7. Lipase and protein assay ...........................................................................71

4.3. Results and discussion .....................................................................................72

4.3.1. Cell disruption for lipase extraction..........................................................76

4.3.2. Comparison of different cell disruption methods for lipase extraction ....82

4.4. Conclusion .......................................................................................................83

Chapter 5: Tween 80 influences the production of intracellular lipase by Schizochytrium S31 in a stirred tank reactor .........................................................84

5.1. Introduction......................................................................................................85

5.2. Materials and methods.....................................................................................86

5.2.1. Microorganism and maintenance of culture..............................................86

5.2.2. Schizochytrium S31 growth and lipase production ...................................86

5.2.3. Estimation of Schizochytrium S31 growth................................................87

5.2.4. Optimisation of cell disruption through sonication...................................87

5.2.5. Lipase assay and protein estimation..........................................................88

5.2.6. Lipid extraction .........................................................................................88

5.3. Results and discussion .....................................................................................88

5.3.1. Optimisation of culture conditions for lipase production .........................88

5.3.2. Production of lipase in stirred tank reactor ...............................................90

5.3.3. Disruption of Schizochytrium S31 using sonication .................................92

5.3.4. Hydrolysis of p-Nitrophenyl esters ...........................................................95

5.4. Conclusion .......................................................................................................96

Chapter 6: Bead milling for lipid recovery from thraustochytrids cells.............97

6.1. Introduction......................................................................................................98

6.2. Materials and Methods ....................................................................................99

6.2.1. Chemicals..................................................................................................99

6.2.2. Thraustochytrids biomass cultivation .....................................................100

6.2.3. Cell disruption by bead milling...............................................................100

6.2.4. Microscopic observation .........................................................................101

6.2.5. Lipid extraction .......................................................................................101

6.2.6. Lipase catalysed hydrolysis of thraustochytrid oil..................................101

vii

Page 11: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

6.2.7. Thin Layer Chromatography for lipid class analysis ..............................102

6.2.8. Gas Chromatography analysis for fatty acid composition ......................102

6.3. Results and discussion ...................................................................................102

6.3.1. Thraustochytrid cultivation .....................................................................102

6.3.2. Cell disruption by bead mill....................................................................103

6.3.3. Lipid profile analysis ..............................................................................107

6.3.4. Selective hydrolysis of Schizochytrium DT3 oil using Candida rugosalipase .................................................................................................................108

6.4. Conclusion .....................................................................................................111

Chapter 7: Phospholipase A1 as a catalyst to enrich omega-3 fatty acids in anchovy and algal oil ..............................................................................................113

7.1. Introduction....................................................................................................114

7.2. Material and Methods ....................................................................................116

7.2.1. Chemicals................................................................................................116

7.2.2. Hydrolysis of anchovy oil .......................................................................116

7.2.3. Thraustochytrid oil hydrolysis by phospholipase A1 .............................117

7.2.4. IATROSCAN analysis ............................................................................117

7.2.5. Methylation and GC analysis ..................................................................117

7.2.6. Analysis of positional distribution of fatty acid by NMR.......................118

7.3. Results and discussion ...................................................................................118

7.3.1. Lipase activity of Phospholipase A1.......................................................118

7.3.2. Hydrolysis of Anchovy oil by PLA1 ......................................................119

7.3.3. 13C NMR studies of the anchovy oil hydrolysis by PLA1 and BSL.......123

7.3.4. Thraustochytrid oil hydrolysis by phospholipase A1 .............................128

7.4. Conclusion .....................................................................................................130

Chapter 8. Summary and future aspects..............................................................131

viii

Page 12: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

List of figures

Chapter 1:

Figure. 1.1 Chemical structures of omega-3 FAs, ALA; C18:3n-3 (cis-9,cis-12-cis-

15-octadecatrienoic acid, EPA; C20:5n- -eicosapentaenoic acid)and

DHA; C22:6n- -docosahexaenoic acid).........................................3

Figure. 1.2. Microscopic images of thraustochytrid cells; (A) Schizochytrium S31,

(B) Schizochytrium DT3 and (C) Thraustochytrium AMCQS5-5. ............................11

Figure. 1.3. Classification of cell disruption methods................................................21

Chapter 2:

Figure. 2.1. Schematic diagram of SPV method. .......................................................38

Figure. 2.2. Standard curve representing relationship between Schizochytrium oil

concentration and increase in the absorbance. ...........................................................40

Figure. 2.3. Comparison of various lipids and other compounds with sulfo-phospho-

vanillin reaction. .........................................................................................................41

Figure. 2.4. Profile of colorimetric detection of thraustochytrid whole cell lipids. ...43

Figure. 2.5. Change of lipid colour after reaction with sulfo-phospho-vanillin

reagent. .......................................................................................................................44

Chapter 3:

Figure. 3.1. Lipid extraction percentage from dry biomass using various solvents

from Schizochytrium sp. S31......................................................................................56

Figure. 3.2. Effect of different cell disruption methods for lipid extraction from

thraustochytrids. .........................................................................................................57

Figure. 3.3. Effect of cell disruption on thraustochytrids cells. Thraustochytrid cells

before (A) and after (B) cell disruption (scale bar 20 μm).........................................58

Figure. 3.4a. Effect of different cell disruption methods on fatty acid profiles of

Schizochytrium sp. S31...............................................................................................61

Figure. 3.4b. Effect different cell disruption methods on SFA, MUFAs and PUFAs of

Schizochytrium sp. S31...............................................................................................62

Figure. 3.5a. Effect of different cell disruption methods on fatty acid profiles of

Thraustochytrium sp. AMCQS5-5. ............................................................................62

Figure. 3.5b. Effect of different cell disruption methods on SFA, MUFAs and PUFAs

of Thraustochytrium sp. AMCQS5-5.........................................................................63 ix

Page 13: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4:

Figure. 4.1. Cluster analysis of thraustochytrids based on various enzyme activities.

....................................................................................................................................75

Figure. 4.2. Microscopic observation of thraustochytrid after cell disruption. ..........77

Figure. 4.3. Effect of bead vortexing on thraustochytrid cell disruption and lipase

activity. .......................................................................................................................78

Figure. 4.4. Effect of grinding on thraustochytrid cell disruption and lipase activity.

....................................................................................................................................79

Figure. 4.5. Effect of sonication on thraustochytrid cell disruption and lipase activity.

....................................................................................................................................81

Figure. 4.6. Effect of homogenisation on thraustochytrid cell disruption and lipase

activity. .......................................................................................................................82

Chapter 5:

Figure. 5.1. Effect of Tween 80 concentration on lipase production by

Schizochytrium S31 at shake flask, pH 6.0 and incubation temperature 25 oC..........89

Figure. 5.2. Fermentation profile for the production of lipase by Schizochytrium S31

in a stirred tank reactor……………………………………………………………...92

Figure. 5.3. Biomass, lipid and DHA production reached a maximum after five days

of incubation...............................................................................................................91

Figure. 5.4. Fatty acid composition of Schizochytrium S31.......................................92

Figure. 5.5. Cell disruption optimisation of Schizochytrium S31 a) as a function of

sonication power, b) time duration.............................................................................95

Figure. 5.6. Substrate specificity of lipase from Schizochytrium S31........................96

Chapter 6:

Figure. 6.1. Schematic diagram of experimental process followed from

thraustochytrid cell disruption to enzymatic omega-3 fatty acids concentration. ....104

Figure. 6.2. Thraustochytrid cell disruption (40 X magnification) employing bead

mill............................................................................................................................104

Figure. 6.3. Optimisation of lipid extraction by bead mill as function of biomass

concentration.. ..........................................................................................................105

x

Page 14: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Figure. 6.4. Effect of bead mill agitator speed (rpm) on lipid yield from

thraustochytrid..........................................................................................................106

Figure. 6.5. Fatty acid composition of lipid extracted from thraustochytrids S31 and

DT3...........................................................................................................................108

Figure. 6.6. Capillary chromatography (iatrascan) profile of thraustochytrid oil ....109

Figure. 6.7. Optimisation of enzyme concetrtation. .................................................110

Figure. 6.8. Lipid class analysis of thraustochytrid oil after hydrolysis with Candida

rugosa lipase at different time intervals. ..................................................................110

Figure. 6.9. AG and FFA are separated hydrolysed oil and analysed by gas

chromatography........................................................................................................111

Chapter 7:

Figure.7.1 Phospholipase A1 activity towards p-NP esters with different acyl chains.

..................................................................................................................................119

Figure. 7.2. Time course of hydrolysis of anchovy fish oil by PLA1 and BSL A ...120

Figure. 7.3. Time course of hydrolysis of anchovy oil and fatty acid distribution in

the hydrolysate. ........................................................................................................122

Figure. 7.4. Percent hydrolysis of various fatty acids (from Fig.7.3) at 32%

hydrolysis of anchovy oil by PLA1 (light) and BSL (dark).....................................124

Figure. 7.5. 13C NMR spectra of the unhydrolysed oil fraction before (blue) and

after (red) 30% hydrolysis of anchovy oil by PLA1 (A) and BSL (B). ...................125

Figure. 7.6. Percent accumulation of different fatty acid classes in TG fraction after

30% hydrolysis of anchovy oil by PLA1 (dark) and BSL (light). ...........................126

Figure. 7.7. Lipase hydrolysis of structured triglyceride1, 3-dioeoyl-2-myristic

triglyceride................................................................................................................127

Figure. 7.8. Lipid class analysis of thraustochytrid oil after hydrolysis with

phospholipase at different time intervals..................................................................129

Figure. 7.9. Fatty acid profiles of hydrolysed and unhydrolysed DT3 oil. ..............130

xi

Page 15: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

List of tables

Chapter 1:

Table 1.1. Contents of EPA and DHA in some fish species. .......................................6

Table 1.2. Dietary recommendation of omega-3 FAs by various scientific bodies. ....7

Table 1.3. Recent studies investigating the use of organic solvent in the extraction of

microalgal lipids. ........................................................................................................16

Table 1.4. Extraction of lipids improved by cell disruption methods. .......................24

Table 1.5. Main patents relating production of omega-3 fatty acids processing using

lipases from the last 10 years. ....................................................................................29

Chapter 2:

Table 2.1. Biomass and lipid production in glycerol medium by thraustochytrids....39

Table 2.2. Comparison of thraustochytrid lipid content by SPV, gravimetric and GC

methods. .....................................................................................................................45

Chapter 3:

Table 3.1. Biomass, lipid contents and productivity of thraustochytrids. ..................54

Table 3.2. Comparison of cell disruption methods employed to extract total lipids

from different microalgae...........................................................................................58

Table 3.3. Advantages and disadvantages of the investigated cell disruption methods.

....................................................................................................................................60

Table 3.4. Biodiesel properties of the given thraustochytrids strains. .......................64

Table 3.5. Energy consumption comparison. .............................................................66

Chapter 4:

Table 4.1. Lipase activity of thraustochytrids new isolates and ATCC cultures.

Lipase activity was determined by p-NPP assay using whole cell as catalyst. ..........76

Table 4.2. Disruption of thraustochytrid cells for lipase extraction. ..........................83

Chapter 5:

Table 5.1. Effect of sonication on protein release and lipase activity from

Schizochytrium S31. ...................................................................................................93

xii

Page 16: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6:

Table 6.1. Energy consumption analysis for different mass concentrations and lipid

yields. .......................................................................................................................106

Chapter 7:

Table 7.1. Positional distribution of various fatty acids in anchovy oil. ..................123

xiii

Page 17: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

List of abbreviations

°C - Degree Celsius

% - Percentage

- Micrometre

nm - nanometre

mg - Milligram

- Microgram

ml - Millilitre

- Micro litre

M - Micro mole

nM - Nano mole

L - Litre

g L-1 - Grams per litre

mg L-1 - Milligrams per litreL-1 - Micrograms per litre

Kg - Kilo gram

h - Hours

min - Minutes

sec - Seconds

- Alpha

- Gamma

- Omega

DHA - Docosahexaenoic acid

EPA - Eicosapentaenoic acid

AA - Arachidonic acid

OA - Oleic acid

GLA - Gamma-linolenic acid

DGLA - Di-homo gamma-linolenic acid

ALA - Alpha-linolenic acid

LA - Linoleic acid

SFAs - Saturated fatty acids

MUFAs - Monounsaturated fatty acids

PUFAs - Polyunsaturated fatty acids

FAs - Fatty acidsxiv

Page 18: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

TFA - Total fatty acids

TAGs - Triacylglycerols

DAGs - Diacylglycerol

MAGs - Monoacylglycerol

PLA1 - Phospholipase A1

BSL - Bacillus subtilis lipase

DU - Degree of unsaturation

SV - Saponification value

IV - Iodine value

CN - Cetane number

LCDF - Long chain saturated factor

CFPP - Cold filter plugging point

CP - Cloud point

APE - Allylic position equivalents

BAPE - Bis-Allylic position equivalents

OS - Oxidation stability

HHV - Higher heating value

KV - Kinematic Viscosity

MJ - Mega joules

BHT - Butylated hydroxytoluene

GYP - Glucose, Yeast extract, Peptone

sp. - Species

GC - Gas Chromatography

FID - Flame Ionisation Detector

DCW - Dry cell weight

v/v - Volume by volume

w/v - Weight by volume

w/w - weight by weight

SD - Standard deviation

OD - Optical density

FAMEs - Fatty acid methyl esters

ATCC - American Type Culture Collection

xv

Page 19: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Abstract

Marine microalgae are a renewable alternative source to fish oil for the

sustainable production of omega-3 fatty acids. The potential use of marine microalgae

for nutraceutical and biofuel production is of global interest owing to increasing

concerns with sustainable production and environment impact. Thraustochytrid are

common marine heterotrophs, capable of producing large amounts of lipid,

particularly omega-3 fatty acids. These organisms are also potential sources of

bioactive proteins, enzymes, pigments and exopolysaccharides. The development

of rapid lipid quantification method for the isolation of new strains from diverse

marine habitats is important for future cost-effective production of these bioactive

metabolites. Cell disruption methods for improving lipid and lipase extraction has the

potential use in economising large scale algal processing for biofuel production.

Further, development of enzymatic methods for concentrating omega-3 fatty acids

from fish and algal oils is desirable because of their beneficial health effects.

Isolation and screening of suitable microalgae with high lipid productivity are

important for achieving sustainable commercial production of omega-3 oils and

biofuel. Current methods of lipid quantification are time consuming and costly.

Quantification of microalgal lipids using the sulpho-phospho-vanillin (SPV) reaction

is a rapid colorimetric method. This method was successfully tested on in-house

thraustochytrid strains. Conventional gravimetric analysis confirmed the accuracy of

the colorimetric method with R2 = 0.99. The total lipid contents of thraustochytrids

were determined gravimetrically was DT3 33.3%, S31 25.8%, AMCQS5-5 22.8% and

PRA 296 18.0%, respectively. Carbohydrates, proteins and glycerol did not interfere

with the SPV reaction.

Lipid extraction is an integral part of biodiesel production, as it facilitates the

release of fatty acids from algal cells. We evaluated the extraction efficiency of various

solvents and solvent combinations for lipid extraction from Schizochytrium sp. S31

and Thraustochytrium sp. AMCQS5-5. The maximum lipid extraction yield was 22%

using a chloroform:methanol ratio of 2:1. The highest lipid extraction yields were

obtained using osmotic shock method with 48.7% lipid yield from Schizochytrium sp.

S31 and 29.1% from Thraustochytrium sp. AMCQS5-5. Saturated and

monounsaturated fatty acid contents were more than 60% in Schizochytrium sp. S31

which suggests their suitability for biodiesel production.xvi

Page 20: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Marine microorganisms are a potential source of enzymes with structural

stability, high activity at low temperature and unique substrate selectivity. We

screened for a range of enzymatic activities from the seven thraustochytrid strains. The

relationship between isolate(s) and enzyme activity was investigated using cluster

analysis. The two strains AMCQ-4b27 (34 IU/g) and AMCQS1-9 (36 IU/g) exhibited

the highest lipase activity of the strains investigated. Four different cell disruption

methods, bead vortexing, grinding, sonication and homogenization, were evaluated for

their impact on lipase extraction yields. Sonication was found to be the best method

for enhancing lipase extraction yields from Thraustochytrium sp AMCQ-4b27 (0.903

IU/μg) and AMCQS1-9 (1.44 IU/μg).

The effect of Tween 80 as a carbon source was investigated with regard to

biomass, lipase and lipid productivity in Schizochytrium sp. S31. Tween 80 (1%) and

120 h of incubation was the optimum period for biomass, lipid and lipase productivity

in a stirred tank reactor. The yields obtained were 0.9 g L-1 of biomass, 300 mg g-1 of

lipid and 39 U g-1 of lipase activity. Sonication was optimised in terms of time and

acoustic power to maximise the yield of extracted lipase. The extracted lipase from

Schizochytrium S31 had unusual fatty acid selectivity in that it preferentially

hydrolysed DHA over EPA.

A method for lipid extraction from marine thraustochytrids using a bead mill

and enzymatic concentration of omega-3 fatty acids from the thraustochytrid oil was

investigated. The optimised lipid extraction conditions were, bead size 0.4-0.6 μm,

4500 rpm, 4 min of processing time at 5% biomass concentration. The maximum lipid

yield achieved at optimum conditions were 40.5% for Schizochytrium sp. S31 (ATCC)

and 49.4% for Schizochytrium sp. DT3 (in-house isolate). DT3 oil contained 39.8%

DHA as a percentage of lipid, a higher DHA percentage than S31. Partial hydrolysis

of DT3 oil using Candida rugosa lipase was performed to enrich omega-3

polyunsaturated fatty acids (PUFAs) in the glyceride portion. Total omega-3 fatty acid

content was increased to 88.7%.

For enrichment of omega-3 fatty acids in fish and algal oils, we have explored

the possibility of using phospholipase A1 (PLA1), a sn-1 position specific

phospholipid hydrolysing enzyme that also demonstrates lipase activity. PLA1 from

Thermomyces species was initially tested for its substrate specificity with various p-xvii

Page 21: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

nitrophenyl esters. In this study, lipase from Bacillus susbtilis (BSL) was taken as a

control. The pH stat based hydrolysis followed by gas chromatographic

characterisation of the hydrolysate of Anchovy fish oil indicated a selective retention

of omega-3 fatty acids in the triglyceride fraction by PLA1, whereas such

discrimination was not observed with BSL. Selective accumulation of omega-3 fatty

acids was also confirmed by 13C NMR spectra. Position-wise analysis of fatty acids in

the unhydrolysed fractions suggests that PLA1 preferably retains docosahexaenoic

acid. Phospholipase A1 was used to catalyse the hydrolysis of thraustochytrid DT3 oil

to enrich DHA in glyceride portion, with short chain fatty acids being preferentially

cleaved from the triacylglyceride. This results show that phospholipase A1 can be used

to concentrate DHA from triacylglycerides.

xviii

Page 22: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1

Chapter 1: Introduction and Literature review

1

Page 23: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 1.1. Introduction

Omega-3 fatty acids (FAs) are essential FAs, meaning that the human body

cannot synthesise them, so they must be obtained from the diet. Omega-3 FAs are well

known for their health benefits (Swanson et al., 2012). For example, high consumption

of omega-3 fatty acids in their diet decrease the risks of sudden death particularly in

those who already have had a heart attack (Blondeau, 2015; Hoen et al., 2013). DHA

is added to the infant formulas, as it is necessary for eye and brain development. These

fatty acids may also decrease the risks of as depression, schizophrenia, Alzheimer’s

and Parkinson’s diseases (Nichols et al., 2014; Tai et al., 2013). Multiple health

benefits associated with omega-3 FAs have increased the demand for these ingredients

in nutraceuticals and functional food industries. It has been forecast that the

consumption of omega-3 FAs in 2014 was 134.7 thousand metric tons valued at US$

2.5 billion. By 2020, it is been projected that the demand further will increase

significantly to 241 thousand metric tons with a value of US$ 4.96 billion (Market

watch, 2014). Most of the health benefits associated with omega-3 fatty acids are

attributed to the long-chain FAs eicosapentaenoic acid (EPA, C20:5-3) and

docosahexaenoic acid (DHA, C22:6-3). The most widely and naturally available

sources of EPA and DHA are fish oils. Concentration of EPA and DHA from fish oils

is normally achieved via chemical methods, which are not environmentally friendly,

also result in partial degradation of long chain FAs and production of toxic waste. This

has led researchers to seek an alternative, environmentally friendly method (Namal

Senanayake, 2013; Shahidi & Wanasundara, 1998). Enzymatic concentration of EPA

and DHA from fish oil using selective lipases is ecofriendly and causes less damage

to chemically sensitive long chain FAs (Shahidi & Wanasundara, 1998).

Production of EPA and DHA from fish oils has resulted in increased pressure

on fish species and number. Moreover, it has been recognised that mass fishing is not

a sustainable approach for EPA and DHA production (Lenihan-Geels et al., 2013). In

addition, these fish oils can be contaminated with heavy metal (methyl mercury) that

are hazardous to human health, particular to pregnant and lactating women’s (Lin et

al., 2014; Mahaffey et al., 2011). A viable alternative source of omega-3 FAs are

heterotrophic microalgae (Adarme-Vega et al., 2012). Commercial scale production

has been performed by using Crypthecodinium cohnii and Schizochytrium sp. as a

source for DHA (Infant formulas) and EPA (Food and beverages) (Kuratko et al.,

2

Page 24: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 2013; Kuratko & Salem, 2013). These algae can be grown in suitably sterile conditions

and oil extracted in a manner that is suitable for use in infant formula and dietary

supplements. Biomass harvesting and lipid extraction are expensive steps in the

production of oil from algal biomass (Kim et al., 2013). Effective cell disruption

methods and efficient solvent extraction methods are critical for high lipid yield and

cost effective production (Ryckebosch et al., 2014).

1.2. Omega-3 FAs

The term ‘Omega-3’ refers that the first double bond between the third and

fourth carbon atoms from methyl end, typically in cis configuration, hence called

omega-3 FAs. In cis configuration the hydrogen atoms are attached to double bonds at

the same side (Figure 1.1). Natural PUFAs consist of an even number of carbon atoms

with various unsaturation points. The most common omega-3 FAs are DHA C22:6,

EPA C20:5 and -linolenic acid (ALA) C18:3 (Poudyal et al., 2011; Ratnayake &

Galli, 2009). ALA is a precursor molecule in the biosynthesis of omega-3 FAs.

Figure. 1.1 Chemical structures of omega-3 FAs, ALA; C18:3n-3 (cis-9,cis-12-cis-15-

octadecatrienoic acid, EPA; C20:5n- -eicosapentaenoic acid)and

DHA; C22:6n-3 -docosahexaenoic acid). Modified from (Akanbi et

al., 2014b).

Generally omega-3 FAs are stored in adipose tissue in triacylglyceride (TAG)

form and present in all cellular membrane in phospholipid and related forms. Cell

membrane phospholipid composition determines the physiological characteristics of

the cell membrane, including changes in response to the external environment and

Alpha-linolenic acidALA (C18:3n-3)

Eicosapentaenoic acidEPA (C20:5n-3)

Docosahexaenoic acidDHA (C22:6n-3)

3

Page 25: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 functional activities of membrane-bound proteins (Murphy, 1990). The proportion of

EPA and DHA to all other FAs present in cell membranes differs and impacts the

function of cells, particularly storage and transport (Browning et al., 2012). DHA is

predominantly present in specific regions such as brain and eye. So for example, DHA

constitutes 18% of FA in adult human brain grey matter (Innis, 2007; Skinner et al.,

1993). With demonstrated physiological roles and established clinical benefits, the link

between human health and EPA and DHA is relatively well established.

1.2.1. Health benefits of omega-3 FAs

In the beginning of the 20th century dietary fats were recognised as a good

source of energy and fat soluble vitamins. This view later changed following various

studies, including a study conducted on rats reported that ‘dietary fatty acids were

required to prevent deficiency disease’ (Burr & Burr, 1929). The health benefits of

omega-3 FAs have been evaluated by a number of clinical studies. There is evidence

for the efficacy of omega-3 FAs in the prevention of sudden death from cardiovascular

disease and in ameliorating rheumatologic conditions (Barrow, 2010; Bhangle &

Kolasinski, 2011; Cicero et al., 2015). An early study on fat consumption and health

with Greenland Inuit indicated the cardiovascular benefits of omega-3 FAs. Further

studies produced more evidence for the cardiovascular benefits associated with

omega-3 FAs, including the reduction of the risk of arrhythmias (Asif, 2014),

decreasing platelet aggregation (Gao et al., 2013), lowering plasma triglycerides (Qi

et al., 2008), decreasing blood pressure (Cabo et al., 2012).

The consumption of dietary omega-3 FAs by ulcerative colitis and Crohn’s

disease patients has been shown to produce beneficial weight gain and significant

improvement in disease activity (Papadia et al., 2010). Omega-3 FAs have been

recently shown to have anti-cancer activity, particularly against colorectal cancer

(Cockbain et al., 2012). Many research investigations have been conducted to

understand the positive effects of omega-3 FAs on growth, learning and behavioural

problems of young children. Supplementation of omega-3 FAs during pregnancy

resulted in an increase in birth size (Makrides et al., 2011). Brain and eye development

in infants is particularly influenced by DHA, which has led to the addition of DHA to

infant formulae. Learning capacity of school going children is increased by taking

DHA enriched food supplements (Birch et al., 2007; Nunes et al., 2014; Richardson &

Montgomery, 2005). A number of studies have provided evidence for the anti-

4

Page 26: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 inflammatory effects of omega-3 FAs. Omega-3 FAs reduce the production of pro-

inflammatory molecules (eicosanoids) and increase the production of anti-

inflammatory molecules such as resolvins and protectins (Calder, 2012). The

combined effect of these multiple health benefits has been a dramatic increase in the

market demand for omega-3 FAs.

1.2.2. Dietary sources of omega-3 FAs

The main traditional sources of omega-3 FAs are marine fish. The marine

ecosystem is a rich source of omega-3 FAs, which are naturally produced by

microalgae and accumulated in fish and higher animal via the food chain (Lenihan-

Geels et al., 2013). Some other common sources include crustaceans, such as lobsters,

krill and molluscs, such as oysters and squid. Some land animals and plant oils also

contain omega-3 FAs in small quantities. Plant oils, such as canola, walnut oil, olive

oil and flaxseed oils are rich source of the omega-3 fatty acid ALA, a precursor for the

biosynthesis of EPA and DHA. To meet current requirements for men and women, the

recommended consumption of olive oil and flaxseed oil varies from 158–229 ml/d

(olive oil) to 2.2–3.2 ml/d (flaxseed oil) (USDA, 2006). The recommended intake of

omega-3 FAs from seeds and nuts to meet ALA recommended consumption for

women and men varies, from 12.2–17.6 g/d of English walnuts, 4.8–7.0 g/d of

flaxseeds, 612–890 g/d of pumpkin seeds (USDA, 2006). Many studies indicate that

the conversion of ALA to long chain FAs EPA and DHA is very low (Burdge &

Calder, 2005), because of this, diet rich in preformed EPA and DHA are required to

achieve the associated health benefits.

Marine derived fish oils are an excellent source for EPA and DHA. The

percentages of EPA and DHA in total fish oils vary depending on the fish species.

Farmed fish have more fat than wild fish, but the level of omega-3 in that fat varies

depending on fish feed use. Farm fish often has lower omega-3 levels than wild fish

even, although in one study farm-raised Atlantic salmon has 2.128 g SFA/85 g and

1.825 g EPA+DHA/85 g; whereas wild Atlantic salmon has 1.068 g SFA/85 g and

1.564 g EPA+DHA/85 g, respectively (Gebauer et al., 2006). Some of popular wild

caught fish and their EPA and DHA contents listed in table 1.1.

5

Page 27: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 Table 1.1. Contents of EPA and DHA in some fish species.

Common name Scientific name g/100g of food stuff portion

EPA DHA

Mackerel Scomber scombrus 1.10 2.56

Red mullet Mullus surmuletus 0.91 1.66

Sardine Sardina pilchardus 0.62 1.12

Salmon Salmo salar 0.50 1.00

Ton Thunnus thinnus 0.24 0.98

Fresh anchovy Engraulis

encrasicolus

0.14 0.80

Sea bream Pagellus bogaraveo 0.12 0.61

Cod Gadus morrhua 0.23 0.47

Hake Merluccius

merluccius

0.10 0.54

Conger eel Conger conger 0.15 0.43

Swordfish Luvarus imperialis 0.15 0.30

Dogfish Galeorhinus geleus 0.04 0

This table has been reproduced and modified from Rubio et al., (2010)

Dietary recommendation for omega-3 FAs advised by various scientific bodies

are given in table 1.1. The Dietary Guidelines for Americans 2005 report states,

‘‘Evidence suggests consuming approximately two servings of fish per week

(approximately 227 g; 8 ounces total) may reduce the risk of mortality from CHD

(Coronary heart disease) and that consuming EPA and DHA may reduce the risk of

mortality from cardiovascular diseases in people who have already experienced a

cardiac event’’. Another important challenge is meeting the recommended

consumption of omega-3 FAs in vegetarians and people who do not enjoy the taste of

fish. Alternative ways of getting omega-3 FAs for vegetarian is consuming fish oil

supplements and plant oils with significant omega-3 fatty acid content (Gebauer et al.,

2006; Maurer et al., 2012).

6

Page 28: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 Table 1.2. Dietary recommendation of omega-3 FAs by various scientific bodies.

S.

No

Professional body (year) Recommendation

1 UK Scientific Advisory

Committee on Nutrition (2004)

Minimum of 2 portions of fish/week that

is equivalent to 450 mg EPA + DHA

2 World Health Organization

(2003)

Up to two fish meals/week (400–1000 mg

EPA+DHA)

3 US National Academies of

Science, Institute of Medicine

(2002)

-linolenic acid/day (140 mg

EPA+DHA)

4 American Heart Association

(2002)

1g/day omega-3 FA for secondary

prevention of

CHD

This table has been reproduced from Yashodra et al., (2009)

1.2.3. Techniques used for the concentration of omega-3 FAs

Fish is a proven natural source of omega-3 FAs, but its consumption may not

meet recommended levels. A prior concentration and purification step is valuable,

therefore, to convert fish oils into a suitable chemical form which can be easily

metabolised in the human body, with the added benefit of improved oxidative stability.

Some studies suggest that the acylglyceride form of omega-3 FAs are better absorbed

than as ethyl esters and free fatty acid forms. In addition to that, some studies reported

that their oxidative stability has been increased in acylglyceride form (Dyerberg et al.,

2010). However, eating large amount of fish is also associated with the risk of heavy

metal – including mercury absorption, which may discourage people from eating fish.

(Lin et al., 2014). To achieve highly purified omega-3 fatty acids concentrate, physical,

chemical and enzymatic methods may be employed. The common methods employed

for the concentration and purification of omega-3 fatty acids from marine oils include

urea adduction, chromatographic methods, low-temperature fractional crystallization,

7

Page 29: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 supercritical fluid extraction, distillation, and enzymatic and integrated methods

(Namal Senanayake, 2013).

Urea adduction is simple and efficient technique for the concentration of

omega-3 fatty acids from marine fish oils. The principle behind this method is to

hydrolyse fish oil, using ethanolic KOH or NaOH, then mixing free FAs with urea

solution to form a complex. Saturated and monounsaturated fatty acids form complex

structures with urea and PUFAs include omega-3 FAs remain in the liquid portion,

which is highly enriched with omega-3 fatty acids. These two portions can be separated

by filtration and concentrated omega-3 FAs portion can be collected (Shahidi and

Senanayake, 2006). The major disadvantage of this method is that it produces large

amounts of urea waste, the disposal of which is expensive (Zuta et al., 2003).

Another important method for the concentration of omega-3 fatty acids from

fish oils is chromatographic techniques. This method involves saponification, followed

by treatment with urea solution. The non-complexed free fatty acids are converted into

fatty acid methyl esters and separated by chromatography (Namal Senanayake, 2013).

Guil-Guerrero and Belarbi used silver nitrate-impregnated silica gel column to purify

EPA and DHA from cod liver oil. Multiple washings with mobile phases led to the

recovery of DHA and EPA, 64% and 29.6%, with resulting purities of 100% and

90.6%, respectively (Guil-Guerrero & Belarbi, 2001).

Low-temperature fractional crystallisation is a simple method that depends on

differences in melting points of different fatty acids. The melting point depends on the

degree of unsaturation of each fatty acid. In general, the more saturate the fatty acids

the higher the melting point. The omega-3 FAs EPA and DHA melting points were -

54 and -44.5 oC, compared with 18:1 (Oleic acid) and 18:0 (Steric acid) of 13.4 and

69.6 oC, respectively (Namal Senanayake, 2013). As temperature decreases for a

mixture of saturated and unsaturated fatty acids, saturated fatty acids first crystalise

and the liquid portion enriched with unsaturated fatty acids (Shahidi and Senanayake,

2006).

Fractional distillation is a straight forward and widely accepted method for the

concentration of omega-3 FAs from a mixture under ultralow pressure. Based on their

boiling point and molecular weight, fatty acids are separated at reduced pressure and

elevated temperature of about 250 oC. This method allows highly selective separation

to concentrate selective fatty acids, but its high operating temperatures results in

8

Page 30: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 increased operating costs and damage to the oxidatively sensitive omega-3 FAs

(Breivik, 2007).

Large scale concentration of omega-3 FAs by physical or chemical methods is

associated with some drawbacks in terms of low yield, use of large volumes of

chemical, production of large volumes of waste effluents, high operation costs and

damage to the resulting fatty acids (Senanayake, 2000). Enzymatic methods such as

the use of lipase in concentrating omga-3 FAs is an alternative green method. Lipase

work under mild operating conditions, which make them suitable for the concentration

of omega-3 FAs from fish oils. Lipases have been used for many years in lipid

modification and food processing. Lipase based concentration involves either cleaving

short chain fatty acids from glyceride or selectively hydrolysing EPA and DHA from

glyceride. Thraustochytrid represent a group of marine microalgae and a source from

which to isolate lipases with novel activities. A novel lipase with selectivity towards

EPA and DHA would be useful and no such lipase has been characterised from any

source. Thraustochytrid are a potential source of omega-3 specific lipases. A detailed

review is given in (Section 1.3.6) on lipase mediated concentration of omega-3 fatty

acids.

1.2.4. Microalgae as an alternative source for omega-3 fatty acids

The current commercial source of omega-3 fatty acids for human consumption

is marine fish. Although most fish oil comes from anchovy, which is considered a

sustainable source, there are still concerns regarding overfishing, particularly for larger

fish such as tuna. There are also concerns with the presence of contaminants such as

mercury and polychlorinated biphenyls (PCBs), which are harmful to consumers

(Seabert et al., 2014). Furthermore, the fish oils are not suitable for vegetarians due to

their fish origins. New alternative sources must be found to supply the growing global

omega-3 market (Ratledge, 2013). Various alternative sources such as bacteria, fungi,

plants and microalgae have been explored for commercial production of single cell

oils (SCOs) (Adarme-Vega et al., 2012; Delarue & Guriec, 2014).

Marine microalgae are the primary producer of EPA and DHA in the marine

ecosystem. These naturally grow under a variety of culture conditions including

autotrophic, mixotrophic and heterotrophic (Buono et al., 2014). Use of microalgae as

a source of omega-3 FAs production has several advantages over traditional sources.

They can be grown in a controlled environment with readily available substrates in the 9

Page 31: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 fermentation process. They do not compete for land and fresh water and also eliminate

the risk of chemical pollutant contamination (Ryckebosch et al., 2012a). There are a

number of algal species with higher EPA and DHA contents. Thraustochytrids species

produce elevated quantities of omega-3 FAs with potential commercial application in

infant formulas, food, cosmetic and pharmaceutical products (Sijtsma & de Swaaf,

2004).

1.3. Thraustochytrids

Thraustochytrids are marine heterotrophic fungi like microorganisms, widely

distributed in a variety of habitats such as mangrove detritus, rocky shores, coral reefs,

salt marshes, sandy sediment, coastal waters, and deep sea (Liu et al., 2014a;

Raghukumar & Damare, 2011). The lipid content of thraustochytrids ranges from 50–

77% on a dry weight basis with more than 90% as triacylglycerides. Thraustochytrids

have recently emerged as a potential alternative source for the production of omega-3

fatty acids, particularly for DHA (Lee Chang et al., 2014). For example, Lee chang et

al, 2013, reported that the DHA percentage of total fatty acid content was more than

40% (w/w) in Aurantiochytrium sp. TC20 (Lee Chang et al., 2013). Another study

conducted by Ryu et al, (2013), observed that the DHA content of Aurantiochytrium

sp. KRS101 was 38 % (w/w) of total fatty acids (Ryu et al., 2013). These high DHA

levels indicate that thraustochytrids are potential sources for commercial DHA

production. Some of the important properties of thraustochytrids which could

influence successfully commercial scale production of omega-3 fatty acids include

high growth rates, a high proportion of DHA/EPA as percentage of total lipids and

unaffected growth at low salinity conditions (William et al., 2005).

Thraustochytrids are fungi like microorganisms, but they are classified under

the Labyrinthulomycetes, because ultrastructurally, they are closely related to

heterokont algae (Honda et al., 1999). This phylum contains 8 genera:

Thraustochytrium, Schizochytrium, Ulkenia, Labyrinthuloides, Labyrinthula,

Japonochytrium, Aplanochytriun and Althornia. Thraustochytrium and

Schizochytrium have a rapid growth rate and capable of producing large amount of

DHA (Azevedo & Corral, 1997; Wong et al., 2008). Microscopic image of

thraustochytrids used in this study are presented in Fig. 1.2.

10

Page 32: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1

Figure. 1.2. Microscopic images of thraustochytrid cells; (A) Schizochytrium S31, (B)

Schizochytrium DT3 and (C) Thraustochytrium AMCQS5-5.

Many strains available today for commercial production of EPA and DHA are

the result of intensive procedures such as collection, isolation and screening. Extensive

screening and characterisation of thraustochytrid from a wide variety of samples led

to identification of thraustochytrids that can accumulate about 50% of total lipid, with

more than 25% of DHA (Raghukumar, 2008). Thraustochytrids are typically abundant

in marine sediments where large amounts of decaying organic matter accumulate,

resulting in enzymatic degradation (Lucia & Fernando, 2002). Many studies have

focused on the isolation and characterisation of thraustochytrids from various marine

habitats such as oceanic and coastal marine environments (Burja et al., 2006; Yang et

11

Page 33: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 al., 2010) as well as low temperature environments (Zhou et al., 2010) and the colonial

tunicate Botryllus schlosseri (Rabinowitz et al., 2006). There are two important

techniques used for the isolation of thraustochytrids from marine samples, direct

plating and pollen baiting. Direct plating techniques involve the spreading of samples

on agar plates under sterilised conditions (Bremer, 2000). Pollen baiting is the most

common method for isolation of thraustochytrids by which, sterilised pollen grains are

spread in marine samples with anti-bacterial and fungal agents and incubated for 1-2

weeks. Thraustochytrids cell growth can be observed on the surface of the pollen grain,

after which follows isolation of pure cultures on selective agar medium (Wilkens &

Maas, 2012).

1.3.1. Potential of Thraustochytrids

Currently the most common microalgae used for commercial production of

omega-3 fatty acids are marine derived, particularly from family members of

Thraustochytriaceae and Crypthecodiniaceae. Thraustochytrids are known to produce

several commercially interesting biotechnological compounds such as omega-3 fatty

acids, carotenoids, sterols, steroids and surfactants and also enzymes and

exopolysaccharides (Fan & Chen, 2007; Gupta et al., 2012a). The company DSM

commercially produces DHA rich oil Life’s DHATM using a Schizochytrium sp that

accumulate about 40-45% of DHA and 2% of EPA, for the food, beverage and

supplement industries. DSM uses a different strain of Schizochytrium, which produce

DHA and EPA rich oils in 2:1 ratio, to compete with anchovy fish oil that generally

has a higher EPA than DHA level. The trade name of this oil is Life’s OmegaTM, and

it contains 24% DHA and 12% EPA (Winwood, 2013).

Carotenoids are commonly used as a food colouring agent, particularly

astaxanthin that is used in aquaculture for salmon coloration. Carotenoids produced

either synthetically or derived from microbial production. Production of carotenoids

from microbial fermentation is an attractive area of research interest, as carotenoids

have been found to have protective activity against cancer, aging, ulcers, heart attack,

and coronary artery disease (Rao & Rao, 2007).

1.3.2. Strain selection and sustainable lipid production

Screening of high lipid producing microalgae is a key step for toward the

success of large scale cultivation and biodiesel production from microalgae. These 12

Page 34: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 microalgae are specific to particular environments and certain climate conditions

(Brennan & Owende, 2010; Williams & Laurens, 2010). These organisms occur in

various environments such as freshwater, lacustrine, brackish, marine, maturation

ponds and hyper-saline conditions (Mutanda et al., 2011). During the selection of

suitable microalgae for lipid production, organisms which have high biomass

productivity and lipid accumulation are preferred for mass cultivation than high lipid

storage with low productivity (Amaro et al., 2011; Mata et al., 2010). In addition, there

are other factors which have to be considered such as lipid profile and the capacity to

grow under specific conditions. Some microalgae produce fatty acids, making them

suitable for biodiesel production, whereas others, such as polyunsaturated fatty acids,

are not suitable because their unsaturation bonds are more prone to oxidative

degradation (Chisti, 2007). Biodiesel production depends on lipid profiles, which can

be regulated by the nutritional supply, processing and growth conditions (Amaro et al.,

2011). In some microalgal species, lipid accumulation is maximum during the

stationary phase. This can be induced by limiting one of the growth controlling

nutrients such as nitrogen, phosphorous and silicon (Guschina & Harwood, 2006).

Microalgae accumulate more lipid when subjected to stress conditions and by limiting

one of the growth nutrients, which enhances the production of lipid in microalgae

(Brennan & Owende, 2010; Pruvost et al., 2011).

In addition to bioprospecting for high lipid producing microalgae, finding a

cheaper raw materials for lipid production can reduce production cost. Researchers

have used molasses, crude glycerol and agricultural waste as a carbon source for lipid

production from microorganisms (Karatay & Dönmez, 2010).

1.3.3. Estimation of lipid content and fatty acid profile

Conventional lipid quantification methods involve microalgal biomass being

exposed to extraction solvents that extract lipids from the microalgal cellular matrix.

The whole mixture is centrifuged to separate the lipid layer from water and cell debris,

and the solvent evaporated. The microalgal crude oil thereby obtained is measured

gravimetrically and calculated as a percentage of dry biomass. This method only gives

the amount of lipid present in microalgae cells and not the lipid profile. To analyse the

lipid profile, obtained crude oil is subjected to transesterification process to convert

microalgal lipids to fatty acid methyl esters (FAMEs). The lipid composition and

quantities can be determined using gas chromatography (Bendikiene et al., 2011; Vyas

13

Page 35: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 et al., 2010). This method provides accurate quantity of lipid weight, but requires many

chemicals and multi-step process. In particular, this method is not favourable in

screening microalgae strains for high lipid production.

1.3.3.1. Rapid quantification of lipids

Microalgae strain selection depends on biomass and lipid productivity since

these are the key factors involved in economical biodiesel production. These factors

are critical to achieve, because the conditions favour high biomass may not favour

lipid accumulation, and vice versa (Huo et al., 2011). The biomass and lipid

production, particularly triacylglycerols, can be modulated by varying the culture

conditions such as nutrient limitation. In addition metabolic engineering can be applied

as a relatively new strategy for enhancing lipid accumulation in different microbial

species (Liang & Jiang, 2013). Therefore, a rapid and reliable lipid quantification

technique is required in identifying potential lipid producing strain from

environmental samples and screening large numbers of mutants. Traditional methods

for lipid quantification from microalgae are labour intensive, require solvents and

treatment steps (Wawrik & Harriman, 2010).

Staining of intracellular lipid using fluorescent lipophilic dyes, such as Nile red

for lipid quantification, popularly used for the dyes affinity towards neutral lipids (Han

et al., 2011; Rumin et al., 2015). Lipid content has been successfully quantified in

certain microalgae using Nile red dye, but unsuccessful with other microalgae because

of their thick cell wall that prevent the penetration of the dye. At elevated temperature,

Chen et al (2009) assessed the effectiveness of different solvents as a stain carrier and

found DMSO as good stain carrier for the species Chlorella vulgaris. Using this

optimised method, they successfully quantified lipids in microalgae using 96 well

plates (Chen et al., 2009). In a different study by Chen et al (2011), both conventional

and improved Nile red methods were ineffective in quantifying the lipid content in

green algae such as Pseudochlorococcum sp. and Scenedesmus dimorphus. This

problem was solved by treating the samples with microwaves for 50-60 seconds prior

to lipid quantification (Chen et al., 2011). Several studies have successfully quantified

the lipid content in various microalgae using Nile red methods (Bertozzini et al., 2011;

Gerde et al., 2012; Kou et al., 2013; Satpati & Pal, 2014).

Colorimetric quantification of lipids is achieved by sulfo-phospho-vanillin

methods (SPV). The SPV method was initially introduced by Chabrol and Charonnat14

Page 36: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 (1973) to quantify the total serum lipids in humans (B. Chabrol, 1937; Vatassery et al.,

1981). This method was further improved by Frings and Dunn (1970) and is still

widely used for serum lipid quantification.

1.4. Extraction of lipids from dry cell biomass

Marine microalgae are single-celled microorganisms with potential industrial

applications: as a source of aquaculture feedstock and production of valuable bioactive

compounds such as lipids, carotenoids and enzymes (Amin, 2009; Cao et al., 2013;

Mercer & Armenta, 2011). Lipid extraction is the most costly and critical step in

biodiesel production. Generally, organic solvent extraction method is most commonly

used as it is well established (Bligh & Dyer, 1959; Folch et al., 1957). However,

effective extraction of total lipids from microalgae is not possible due to the rigid cell

wall. In case of slow extraction due to the presence of strong cell wall, it is advised to

perform an extraction dynamic study in order to improve extraction efficiency (Cho et

al., 2012a). The most commonly used organic solvents are benzene, hexane, acetone

and chloroform and these have shown to be effective for lipid extraction from

microalgae by disrupting their cell walls (Ramluckan et al., 2014). Selection of the

solvent for lipid extraction should take into account its ability to enter into the algal

cells efficiently and match the polarity of the desired compounds (for instance, hexane

for non-polar lipids). Physical contact with solvent and lipid material can be achieved

by mechanically disrupting the microalgal cells and adding the solvent later (Li et al.,

2014b). The effect of cell disruption in lipid extraction from microalgae has been

reported and shown to significant improve lipid extraction yields (dos Santos et al.,

2015; Grimi et al., 2014; Halim et al., 2013; Lee et al., 2010b; Yap et al., 2014).

1.4.1. Selection of suitable extraction solvents

In order to optimise lipid extraction yields solvent selection is important.

Chloroform / methanol (1: 2 v/v) is the most commonly used and reliable organic

solvent mixture for total lipid extraction from microalgae (Bligh & Dyer, 1959). Using

this organic solvent system, residual endogenous water in the microalgal cells acts as

a ternary component that enables the complete extraction of both neutral and polar

lipids. This method does not require the complete drying of microalgal biomass. Once

the cell debris is removed, more chloroform and water are added to induce biphasic

partitioning. The lower organic phase (chloroform) contains most of the lipids (both

15

Page 37: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 neutral and polar) while the upper aqueous phase (water with methanol) constitutes

most of the non-lipids (proteins and carbohydrates) (Medina et al., 1998). The

chloroform and methanol lipid extraction process is fast and quantitative. However,

the use of chloroform associated with health, security, and regulatory problems, and

as such, its usage is undesirable. Folch et al, originally developed chloroform and

methanol lipid extraction procedure to extract lipids from brain tissue (Folch et al.,

1951). The lipid extraction yield depends on the solvent mixture used in the extraction

process. As a result, its efficiency in lipid extraction from microalgal biomass need to

be further investigated.

Jeon et al, examined 15 different solvents and analysed their efficiency in lipid

extraction from Chlorella vulgaris biomass. Lipid extraction yields increased to 25%

with methanol:dichloromethane as a solvent mixture compared with other solvents

used (Jeon et al., 2013). Mandal et al, also studied the effect of different solvents and

their combinations in lipid recovery from Scenedesmus obliquus. Around 13% of lipid

extracted using chloroform and methanol (2:1) solvent (Mandal et al., 2013).

Ramluckan et al, investigated the efficiency of 13 solvents and their combinations with

a range of polarities and solubilities. The greatest lipid extraction efficiency (11.76)

was obtained upon using 1:1 mixture of chloroform:ethanol (Ramluckan et al., 2014).

Table 1.3 summarises the recent studies investigating the effect of various solvents on

lipid extraction.

Table 1.3. Recent studies investigating the use of organic solvent in the extraction of

microalgal lipids.

16

Page 38: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1

Marine

Microorganis

ms

Dried

biomas

s (g)

Organic solvents v/v Reaction conditions

Duration

(min)/agitation

(rpm)/Temp (oC)

Wt% of

total

extracted

lipids

Reference

Schizochytriu

m S31

0.05 Chloroform/ethanol (2:1) NA 22 (Byreddy et

al., 2015)

Chlorella sp. 1 Chloroform/ethanol (1:1) 3 h/NA/boiling point 11.76 (Ramlucka

n et al.,

2014)

Scenedesmus

obliquus

1 Chloroform/ethanol (2:1) NA/NA/room

temperature

12.9 (Mandal et

al., 2013)

Nannochloro

psis gaditana

Tetraselmis

suecica

Desmodesmu

s communis

0.05 N,N-

dimethylcyclohexylamine

50 mg/ml

(Biomass/solvent)

24 h/stirring/room

temperature

29.2

57.9

31.9

(Samori et

al., 2013)

Chlorella

vulgaris

50 dichloromethane:methano

l (1:1)

1 h/stirring/37 oC 25 (Jeon et al.,

2013)

Arthrospira

platensis

1 chloroform/methanol

(2:1)

NA 20 (Baunillo et

al., 2012)

Pavlova sp 10 Water/methanol/ethyl

acetate (10:24:48)

NS 44.7; (Cheng et

al., 2011a)

Chlorococcu

m sp

4 Hexane,

hexane/isopropanol (3:2)

450/800/25 6.8;

hexane/isop

ropanol

(Halim et

al., 2011)

Synechocystis

PCC 6803

15 Chloroform/methanol/

water ( 1:2:0.8)

24 h/stirring/room

temperature

4.2 (Sheng et

al., 2011b)

Schizochytriu

m limacinum

1 Water/chloroform/methan

ol (4:6:12)

NA 57 (Johnson &

Wen, 2009)

17

Page 39: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1

Aurantiochyt

rium sp.

strain T66

28-32 Methanol/chloroform/wat

er (2:2:1)

NS/NS/40 55;

Methanol/c

hloroform/

water

(Jakobsen

et al., 2008)

Phaeodactylu

m

tricornutum

10 Ethanol 1440/500/25 6.3 (Fajardo et

al., 2007)

S. mangrovei

IAo-1

100-

200

Chloroform/methanol

(2:1)

NS/Ns/38 33.2 (Leano et

al., 2003)

Chlorella sp 0.1 Soxhelt: methylene

chloride:methanol (2:1)

Batch 1:

chloroform/methanol/50

mM phosphate buffer

(35:70:28)

Batch 2: n-hexane/IPA/

distilled water (70:47.7:3)

Soxhlet: 180/NS/NS

batch 1: 2160/NS/NS

batch

2: overnight/NS/NS

11.9;

Soxhlet

with

methylene

chloride/me

thanol (2:1)

(Guckert et

al., 1988)

Botryococcus

braunii

0.12 Chloroform/methanol

(2:1), hexane/isopropanol

(3:2),

dichloroethane/methanol

(1:1),

dichloroethane/ethanol

(1:1),

acetone/dichloromethane

(1:1)

50/high/NS 28.6;

chloroform/

methanol

(2:1)

(Lee et al.,

1998)

18

Page 40: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1

1.4.2. Lipid extraction from wet biomass

Drying microalgae biomass prior to lipid extraction accounts for 90% of the

total production energy cost in dry algal lipid extraction for biodiesel production.

Microalgae biomass drying is an energy intensive process, more than 25% of energy

required for this process can be saved by using wet algal biomass (Du et al., 2015).

There are few studies on microalgae lipid extraction using wet biomass. Yao et al,

investigated the use isopropanol as a potential solvent in extracting lipids from

Nannochloropsis sp. wet biomass and achieved 70% of lipid extraction (Yao et al.,

2012). Dejoye, et al, developed a new environmentally safer method, called

simultaneous distillation and extraction process for lipid extraction from wet

microalgae Nannochloropsis oculata and Dunaliella salina. This method used

alternative solvents such as d- -pinene and p-cymene. The lipid recovery

yields were compared with standard soxhlet extraction and Bligh & Dyer method and

observed similar lipid yields (Dejoye Tanzi et al., 2013). Yang, et al, proposed the use

of ethanol as an organic solvent for lipid extraction from wet microalgae Picochlorum

sp. This study investigated the effect of extraction temperature and time and the

Isochrysis

galbana

5 Chloroform/methanol/wat

er (1:2:0.8)

hexane/ethanol (1:2.5),

hexane/ethanol (1:0.9),

butanol, ethanol,

EtOH/water (1:1)

60/constant/25 8.9;

chloroform/

methanol/w

ater

(1:2:0.8),

(Grima et

al., 1994)

Chaetoceros

muelleri

and

Monoraphidi

um

minutum

60 1-butanol, ethanol,

hexane/2-propanol

(2/3), water/

methanol/chloroform

as a control system

90/high (NS)/close to

boiling point of each

chemical

Control

100%

extraction.

2nd

Highest

efficiency

with 1-

butanol

at 94%

(Nagle &

Lemke,

1990)

19

Page 41: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 influence of solvent biomass ratios on extraction efficiency and lipid class, with results

indicating that the extracted lipid was comparable with Bligh & Dyer method (Yang

et al., 2014). Another recent study confirms the suitability and efficiency of ethanol as

an extraction solvent for lipid extraction from some microalgae (Yang et al., 2015).

Sathish, et al, developed and modified a lipid extraction procedure from wet

microalgae biomass using acid and base hydrolysis. This method was capable of

extracting 79% of total transesterified lipids from Chlorella and Scenedesmus sp.

(Sathish et al., 2015; Sathish & Sims, 2012).

Recently, oil extraction from wet biomass and direct biodiesel conversion

called direct (in-situ) transesterification, has been investigated (Park et al., 2015). Chen

et al, developed a novel method for microalgal biodiesel production using

Chlamydomonas sp. JSC4 with 68.7 wt% water content as the feedstock. This method

involves microwave disruption, partial dewatering (via combination of methanol

treatment and low-speed centrifugation), oil extraction, and transesterification without

removal of solvent and they achieved 97% biodiesel conversion (Chen et al., 2015a).

A recent study by Kim et al, investigated the effect of factors influencing the direct

biodiesel conversion from wet Nannochloropsis salina biomass. The reaction

conditions such as temperature, reaction time, and solvent and acid quantity were

optimised and 90% of total conversion was obtained (Kim et al., 2015).

1.4.3. Microalgae cell disruption for lipid extraction

A large number of lipid extraction procedures are used for microalgal biomass

and their extraction efficiencies have been documented. All microalgae are single cell

organisms having an individual cell wall. Furthermore, some species of microalgae

have high lipid levels but they are located inside a tough cell wall. The presence of

tough cell walls of microalgae increase the lipid extraction cost, since a cell disruption

step is required for effective lipid extraction. Cell disruption is an energy consuming

process but an important step in liberating intracellular lipid molecules from

microalgae (Lee et al., 2012). A number of cell disruption methods are available for

lipid extraction from microalgae. Different microalgae have different disruption

propensities and as a result, no single method of disruption can be universally applied

to the various microalgae species (Halim et al., 2012b).

Cell disruption methods (Figure. 1.3.) generally used at the laboratory scale are

classified according to the working procedure by which microalgal cells are disrupted, 20

Page 42: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 for example mechanical and non-mechanical (Chisti & Moo-Young, 1986; Günerken

et al., 2015; Middelberg, 1995). Mechanical methods include bead mill, press, high-

pressure homogenisation, ultrasonication, autoclave, lyophilisation, and microwave,

while non-mechanical methods often involve lysing the microalgal cells with acids,

alkalis, enzymes, or osmotic shocks (Chisti & Moo-Young, 1986; Günerken et al.,

2015).

Figure. 1.3. Classification of cell disruption methods. Modified from (Chisti & Moo-

Young, 1986).

1.4.3. Effect of cell disruption methods on lipid extraction

Microalgal cell wall disruption in a non-specific manner is generally achieved

by mechanical methods using solid-shear forces. The basic principle of bead milling

involves physically grinding the microalgal cells against the solid surfaces of glass

beads in a violent agitation (Chisti & Moo-Young, 1986). Some key factors include

high throughput process, high biomass loading, good temperature control, easy scale

up procedures, low labour intensity and high disruption efficiency, which in

combination make bead milling a suitable method of implementing large scale

applications (Günerken et al., 2015; Jahanshahi et al., 2002). The method has been

used for years to disrupt microorganisms. The size of beads used in the disruption

21

Page 43: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 process is important. The optimal diameter of the beads for bacterial cell disruption is

0.1 mm and 0.5 mm for yeast and other unicellular organisms. The cell disruption

efficiency can be further improved by using zirconia-silica, zirconium oxide or

titanium carbide made beads. This is because of their hardness and density. In addition,

the separation of beads from the agitated solution is easily achieved due to the high

density of beads (Lee et al., 2012c).

High pressure homogenisation is an effective and simple cell disruption

method. In this process microalgal suspension pumped through narrow orifice of a

valve under high pressure. Cell disruption is achieved by high-pressure impact (shear

forces) of the accelerated fluid jet on the stationary valve surface as well as

hydrodynamic cavitation from the pressure drop induced shear stress (Chisti & Moo-

Young, 1986). Some recent studies observed that high working pressure and cycle

number have a positive effects on cell disruption efficiency (Halim et al., 2012b). In

addition, microalgae biomass concentration and species have major effect on specific

energy consumption (Halim et al., 2013). High speed homogenization is a simple and

effective method by which a stirring device which can rotate at high rpm with in a

stator-rotor assembly. The stator-rotor are usually made of stainless steel with a variety

of designs. High speed homogenization achieves cell disruption via hydrodynamic

cavitation, generated by stirring at high rpm, and shear forces at the solid-liquid

interphase. Hydrodynamic cavitation occurs when the impeller tip reaches

approximately 8500 rpm due to the local pressure decreases nearly down to the vapour

pressure of the liquid (Gogate & Pandit, 2008; Kumar & Pandit, 1999).

Microalgal cells can be disrupted by an ultrasonication process that is

transmission of sonic waves. An ultrasonicator converts the electrical energy into

mechanical vibrations. The probe intensifies the mechanical vibrations, resulting in the

formation of pressure waves in the liquids. This mechanism forms significant amounts

of cavities (bubbles) in the liquid, which expand during the low pressure phase and

collapse violently during the high pressure phase. This phenomenon is referred as

“cavitation”, which releases pressure and heat. Although the cavitational collapse lasts

only for a few microseconds, the cumulative effect of bubbles is to release high levels

of energy into the liquid (Chisti & Moo-Young, 1986).

Generally the non-mechanical cell disruption methods involve use of

chemicals, enzymes and osmotic shock (Monks et al., 2013). These methods primarily

22

Page 44: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 depend on the interaction between chemicals or enzymes and cell walls of the

microalgae, which modify the outer membrane layer and release the intracellular

components. Enzymatic cell disruption is a desirable process due to biological

specificity, mild operating conditions, low energy requirements, low capital

investment. The use of enzymes normally circumvents the need to use aggressive

physical conditions such as high shear stress and elevated temperatures (Choi et al.,

2010; Harun & Danquah, 2011). The discovery of the enzyme lysozyme led interest

in the application of lytic enzymes in cell lysis. During enzymatic cell lysis, enzymes

bind to cell membranes at specific sites and hydrolyse specific bonds leading to

degradation of cell membrane (Choonia & Lele, 2013; Fu et al., 2010; McKenzie &

White Jr, 1991). A variety of chemical agents such as such as antibiotics, chelating

agents, chaotropes, detergents, solvents, hypochlorites, acids and alkali can cause cell

disruption. The efficiency of the cell disruption depends on the selectivity of the

chemical agent and composition of the microorganisms (Jalalirad, 2013; Middelberg,

1995).

Lee at al., (2010) assessed the effect of prior mechanical cell disruption on lipid

extraction for Botryococcus sp., Chlorella vulgaris, and Scenedesmus sp. using

chloroform and methanol (2:1 v/v) as an extraction solvent. Among the methods used

(autoclave, bead beating, microwave, sonication, and osmotic shocks), the microwave

method was resulted in the higher lipid extraction yields in all three microalgal species.

Another study by Zheng et al., (2011) investigated the effect of different mechanical

and non-mechanical cell disruption methods for release of lipids from a marine species

of Chlorella vulgaris. Overall, grinding in liquid nitrogen was identified as the most

effective method in terms of disruption efficiency.

A similar study was conducted by Halim et al, (2012), in which four cell

disruption methods were compared to determine their lipid extraction efficiency.

Among the four methods (Bead beating, ultrasonication, acid treatment and high-

pressure homogenisation), high-pressure homogenisation was the most effective in

disrupting Chlorococcum cells. Summary of recent studies on microalgal cell

disruption and their lipid extraction efficiencies are given in table. 1.4.

23

Page 45: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 Table 1.4. Extraction of lipids improved by cell disruption methods.

Microalgae Cell disruption

method

Maximum

lipid yield (%)

Reference

Schizochytrium

S31

Thraustochytrium

AMCQS5-5

Osmotic shock 48.7

29.1

(Byreddy et al.,

2015)

Chlorella vulgaris Ultrasonication 55 (dos Santos et al.,

2015)

Scenedesmus sp Freeze drying +

Microwave

29.6 (Guldhe et al.,

2014)

Chlorella vulgaris pressure-assisted

ozonation

27 (Huang et al.,

2014)

Mixed culture Microwaves 33.7 (Florentino de

Souza Silva et

al., 2014)

Chlorella vulgaris H2O2 + FeSO4 17.34 (Steriti et al.,

2014)

Nannochloropsis

sp

Microwave 38.3 (Wahidin et al.,

2014)

C. vulgaris Ultrasonication 52.5 (Araujo et al.,

2013)

Chlamydomonas

reinhardtii

Osmotic shock NA (Yoo et al.,

2012b)

Microcystis sp Sonication 53.89 (Supriya &

Ramachandra,

2012)

C. vulgaris (SAG

211-12)

Grinding+microwaves+

sonication

9.82 (Sostaric et al.,

2012)

Synechocystis

PCC 6803

Microwave+pulsed

electric fields

9-13 (Sheng et al.,

2012)

Chlorella vulgaris Microwaves NA (Prommuak et

al., 2012)

24

Page 46: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 Scenedesmus sp high-pressure

homogenization

24.9 (Cho et al.,

2012b)

Synechocystis

PCC 6803

pulsed electric field 25-75 (Sheng et al.,

2011a)

Chlorella sp

Nostoc sp

Tolypothrix sp

Sonication

Sonication

Microwave

21

18

16

(Prabakaran &

Ravindran,

2011)

Botryococcus sp

Chlorella vulgaris

Scenedesmus sp

Microwave

28.6

10

11

(Lee et al.,

2010b)

Scenedesmus

dimorphus

Chlorella

protothecoides

Wet milling

Bead-beater

25.3

18.8

(Shen et al.,

2009)

1.5. Lipases

Lipase have emerged as an important biocatalyst with potential applications

in many industries including organic synthesis, paper manufacturing, oleochemistry,

dairy, cosmetics, perfume, biosensors, and detergents (Hasan et al., 2006). Lipases

catalyse the breakdown of fats into free fatty acids and glycerol at the water oil

interface. Lipase also synthesise ester from free fatty acids and glycerol at the

hydrophobic interface. Lipases often exhibit high chemo-, region- and enantio-

selectivity at different conditions contributing to their utility as industrial biocatalysts

(Gotor-Fernández et al., 2006).

1.5.1. Marine lipases

The marine ecosystem and marine microorganisms in particular represents an

enormous pool of potential enzyme candidates for industrial and biotechnological

applications. Marine microbes are an important source of bioactive compounds (Soliev

et al., 2011). The demand for the novel enzymes is also increasing, due to their utility

in the production of chemically and structurally important molecules. Marine

microorganisms are a potential source for the isolation of enzymes with novel

properties, because of their stability and their potential for wider applications than

25

Page 47: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 those of enzymes isolated from plant or animal sources (Bull et al., 2000; Lam, 2006).

The potential demand for enzymes in industrial applications is increasing every year.

Marine environments provides unique genetic structures and life habitats such as

nutrient rich regions, nutrient limited regions, high salinity, high pressure, low

temperature and special light conditions. Microorganisms must adapt to survive in the

marine environment and so new biodiversity is created (Kennedy et al., 2008; Zhang

& Kim, 2010).

Lipases are found in all forms of life, includes bacteria, fungi, plants and

animals (Wang et al., 2008). Microorganisms are potent sources of enzymes by

comparison with plants and animals, because of the availability of their wide variety

of catalytic activities, the high yields possible, relative ease of genetic manipulation,

regular supply due to the absence of seasonal fluctuations and rapid growth of

microorganisms (Iftikhar et al., 2010). Microbial lipases are normally more stable than

the other lipase sources which makes them more suitable for industrial applications

(Peng et al., 2014). Fungi are the most studied source for the production of commercial

lipases. Some of the common lipase producing fungi belong to the following genus,

Rhizopus sp., Aspergillus sp., Penicillium sp., Geotrichum sp., Mucor sp., and

Rhizomucor sp. Candida antarctica and candida rugosa lipases are most extensively

studied fungal species (Gopinath et al., 2013; Singh & Mukhopadhyay, 2012). Lipase

production depends on the carbon source, nitrogen source, pH and temperature of the

growth medium (Cihangir & Sarikaya, 2004). Unusual lipase producing

microorganisms can be isolated from the extreme environments, such as sites

contaminated with oil, dairy waste, vegetable oil waste and industrial waste (Sztajer et

al., 1988).

1.5.2. Isolation of lipase producing microbes and screening methods

The increasing demand for lipases in industrial applications has led the

identification of novel lipases with unique properties. Screening for specific lipase

activity and new lipases is facilitated by the development of rapid, reliable, specific,

selective, and sensitive analytical methods for evaluating their catalytic activity

(Stoytcheva et al., 2012). Quantifying lipase activity can be challenging since lipases

are water soluble enzymes but act on water insoluble substrates (Kanchana et al.,

2011). Methods based on using chromogenic and fluorogenic substrate are amongst

the simplest and most widely used methods for screening for lipase activity (Hasan et 26

Page 48: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 al., 2009). A preliminary plate method is normally used in the initial screening for

lipase producing strains. The medium in the agar plates are supplemented with lipid

substrate (olive oil or tributyrin) and microorganisms are streaked on agar plate and

periodically observed (Singh et al., 2006). Microbes that are producing lipase can

make a halo in the plate and then these strains are selected for lipase production in the

liquid medium (Cardenas et al., 2001); Hasan et al., 2009).

Lipases hydrolyse ester bonds and liberate free fatty acids that can be

titrimetrically measured by neutralizing the released free fatty acids by manually

adding NaOH or by a pH stat method (Borkar et al., 2009). In spectrophotometric

methods, substrate are designed to give a colour compound upon enzyme hydrolysis.

Lipase activity can be measured in microbial culture supernatant by using p-

nitrophenyl esters of various chain lengths. The lipolytic activity on carboxylic ester

leads to the release of yellow coloured p-nitrophenol that can be monitored

quantitatively by spectrophotometry (Winkler & Stuckmann, 1979).

Spectrofluorometric methods are rapid, simple and sensitive. Substrates used in this

method are acylglycerol analogs coated with fluorophores such as dansyl, resorufin,

or 4-methylumbelliferone (MUF). These analogues are released by lipolytic activity,

which can be measured fluorometrically (Li et al., 2014a).

Chromatographic techniques can be used to measure lipase activity by

quantifying the released free fatty acids from triacylglycerols. Most general techniques

use thin-layer chromatography, gas-liquid chromatography, HPLC or gas

chromatography- mass spectrometry (Hao et al., 2007; Stoytcheva et al., 2012). Lipase

activity estimation by HPLC was developed by Maurich et al 1999, using beta-

naphthyl or para nitrophenol esters. Substrate were incubated with the enzymes and

released naphthol was quantified by reverse phase HPLC. Gas chromatography is used

to determine the products released by the enzymatic reaction (Stoytcheva et al., 2012).

Since lipases are inducible enzymes, the medium composition, particularly the carbon

source, influences lipase production. Most microbial lipases are extracellular lipases.

Lipase production can be induced by adding lipase substrates into the media, such as

triacylglycerols, fatty acids, hydrolysable esters, tweens, bile salts, and glycerol (Gupta

et al., 2004). In addition, the nitrogen source and some important micronutrients also

play a crucial role in the production of lipases. Lipase production levels can be

increased by the optimisation of culture condition, generally optimised by changing

27

Page 49: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 one variable at a time while keeping other variables constant. However, this systematic

approach is time consuming and doesn’t effectively explore the interaction between

variables. A more versatile approach is the Plackett-Burman (PB) method, which is an

efficient and widely accepted approach for the optimisation of multiple culture

variables (Mazzucotelli et al., 2015).

1.5.3. Extraction of intracellular lipases

Enzymes that are synthesized and remained inside the cells are called

intracellular enzymes. To extract intracellular enzymes, the cells have to be broken by

mechanical or non-mechanical methods to release the products. Some cell disruption

methods are sonication, homogenisation, bead beating, French press, or chemical and

enzymatic methods (Safi et al., 2014b). Jermsuntiea et al, extracted the intracellular

lipase from Mortierella alliacea by homogenising with glass beads and investigated

its substrate selectivity. The novel lipase preferentially hydrolysing monoacylglycerols

containing long-chain unsaturated fatty acids and phosphatidylcholine (Jermsuntiea et

al., 2011). Perez et al, isolated and purified a novel halophilic lipase from

Marinobacter lipolyticus SM19 by ultrasonic treatment. This lipase was employed for

hydrolysing fish oil and preferentially hydrolysed EPA from triglycerides over DHA

(Pérez et al., 2011).

1.5.2. Lipase based methods for concentration of omega-3 fatty acids

Lipases are able to catalyse unique reactions such as hydrolysis, ethanolysis or

transesterification of triglycerides. Fatty acids are naturally distributed in the glyceride

‘backbone’ of several marine oils, and these unique properties of lipases are very

useful both in omega-3 concentration from fish oil and for the production of structured

lipids (Ando et al., 1992; Tengku-Rozaina & Birch, 2014). Lipase based methods can

improve the quality of fish oil concentrates by causing less damage to EPA and DHA

during processing, which impacts the sensory quality and stability of the final products

(Venkateshwarlu et al., 2004). EPA and DHA have multiple double bonds in the cis-

configuration and these are very prone to oxidation, polymerization or isomerisation.

Concentrated omega-3 oils can contain high levels of products such as polymers and

trans fatty acids due to oxidative degradation (Kralovec et al., 2010). An effective way

of concentrating omega-3 fatty acids (EPA and DHA) is selective hydrolysis or

esterification of shorter chain fatty acids and their separation from the omea-3 fatty

28

Page 50: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 acid enriched glycerides (Carvalho et al., 2009; Kralovec et al., 2010). Many

commercial lipases available from fungi and bacteria discriminate against omega-3

fatty acids in that they preferentially hydrolyse shorter chain fats (Haraldsson, 1999).

Saturated and monounsaturated fatty acids are partially selectively cleaved from

triglycerides and long-chain omega-3 fatty acids primarily remain in the glyceride

portion as residual mono-, di- and triacylglycerols. The resistance of omega-3 fatty

acids to lipase hydrolysis by lipases is related to the presence of cis double bonds that

causes steric hindrance and results in bending of the fatty acid chains, which brings

the terminal methyl groups close to the ester bond (Okada & Morrissey, 2007). A more

direct approach is selective hydrolysis of omega-3 fatty acids. However, no omega-3

selective lipases have been characterised, although fish digestive juice may contain

such a lipase (Halldorsson et al., 2004). A more recent study also observed the

presence of a lipase with preferential EPA and DHA selectivity from fish waste

(Kurtovic & Marshall, 2013). A novel lipase isolated from the halophilic bacterium

Marinobacter lipolyticus SM19 showed high efficiency in the enrichment of EPA from

fish oil. These results indicate that selective concentration of omega-3 fatty acids from

marine oils using lipases is possible through selective EPA and/or DHA hydrolysis.

Table 1.5. Main patents relating production of omega-3 fatty acids processing using

lipases from the last 10 years.

S.

NO

Year Title Patent Number Inventors

1 2015 Separation of omega-3 fatty

acids

WO2015024055

A1

Colin James

Barrow, Taiwo

Olusesan Akanbi

2 2014 Method of producing oil/fat

comprising highly-

unsaturated fatty acids by

means of lipase

EP 2682453 A1 Nobushige Doisaki,

Hideo Ikemoto,

Kazuhiko Hata,

Shinji Tokiwa,

Kazunori

Matsushima

3 2013 Process for production of

oil or fat containing highly

EP 2578691 A1 Hideo Ikemoto,

Nobushige Doisaki,

Yasuo Umehara

29

Page 51: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1

unsaturated fatty acid using

lipase

4 2013 Process for separating

polyunsaturated fatty acids

from long chain unsaturated

or less saturated fatty acids

EP 2585570 A1 Gudmundur G.

Haraldsson, Bjorn

Kristinsson

5 2013 Method for enrichment of

eicosapentaenoic acid and

docosahexaenoic acid in

source oils

WO

2013043641 A1

Subramaniam

Sathivel, Huaixia

Yin

6 2012 Enrichment of marine oils

with omega-3

polyunsaturated fatty acids

by lipase-catalysed

hydrolysis

WO2012087153

A1

Kahveci Derya, Xu

Xuebing

7 2011 High-Purity Purification

Method for Omega-3

Highly Unsaturated Fatty

Acids

US

20110091947

A1

Gap Jin Kim, Hong

Joo Son, Wu Song

Whang, Yoon Mo

Koo, Jin Il Kim, Jin

Hyo Yang,

8 2011 Processes to generate

compositions of enriched

fatty acids

WO

2011067666 A1

Inge Bruheim,

Svein Rye

9 2011 Method for producing oil

containing polyunsaturated

fatty acid using lipase

CA 2800674 A1 Hideo Ikemoto,

Nobushige Doisaki,

Yasuo Umehara

10 2010 Triglycerides with high

content of unsaturated fatty

acids

WO

2010115764 A2

Per Munk Nielsen,

Jesper Brask,

Steffen Ernst

30

Page 52: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1

11 2010 A polyunsaturated fatty

acid (PUFA) enriched

marine oil comprising

eicosapentaenoic acid

(EPA) and docosahexaenoic

acid (DHA), and a process

of production thereof

EP 2147088 A1 Patrick

Adlercreutz, Ann-

Marie Lyberg

12 2010 Production method of oil or

fat containing

polyunsaturated fatty acid-

containing triglyceride

US 7767427 B2 Kengo Akimoto,

Motoo Sumida,

Kenichi

Higashiyama,

Shigeaki Fujikawa

13 2010 A polyunsaturated fatty

acid (PUFA) enriched

marine oil comprising

eicosapentaenoic acid

(EPA) and docosahexaenoic

acid (DHA), and a process

of production thereof

EP 2147088 A1 Patrick

Adlercreutz, Ann-

Marie Lyberg

14 2010 Enzymatic modification of

oil

EP 2183376 A2 Colin James

Barrow, Jaroslav

A. Kralovec,

Weijie Wang

15 2010 Method for production of

epa-enriched oil and dha-

enriched oil

EP 2172558 A1 Kiyomi Furihata,

Hiroyuki

Kawahara, Hideaki

Yamaguchi, Hideo

Ikemoto,

Nobushige Doisaki

16 2009 Method for production of

condensed polyunsaturated

fatty acid oil

EP 2006389 A9 Nobushige Doisaki,

Kiyomi Furihata,

Hiroyuki Kawahara

31

Page 53: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1

17 2008 Process of selective

enzymatic enrichment of a

mixture containing omega-3

WO

2008093378 A1

Giovanni Cotticelli,

Raul Salvetti

There are numerous research articles and patents published in relation to the

production of omega-3 fatty acids from marine oils using lipases. Table. 1.6

summarises the patents filed in the last ten years on the production of omega-3 fatty

acids using lipase catalysed reaction. Linder et al, hydrolysed salmon oil using a sn-1,

3 specific lipase to concentrate PUFA 39.20 mol% to 43.05 mol% (Linder et al., 2002).

Another study by Gámez-Meza et al. 2003, investigated the enzymatic hydrolysis of

sardine oil and concentration using urea complexation. They also compared the effect

Pseudomonas lipase hydrolysis with chemical hydrolysis using KOH, ethanol and

hexane. This study concluded that the immobilized lipase with 10% protein was the

most effective system by which to achieve the best purity of omega-3 fatty acids

(86.58%) with an acceptable yield (78.0%). Although chemical hydrolysis yielded

higher omega-3 recovery (90.5%), the purity was lower (83.13%). Moreover, the

lipase based method resulted in easier separation of the products using a packed-bed

reactor (Gámez-Meza et al., 2003). A recent study optimised the hydrolysis of anchovy

fish oil and the concentration of omega-3 fatty acids using TL 100 lipase, from which

24% DHA was concentrated in the glycerol portion under optimised conditions

(Akanbi et al., 2013). Kahveci and Xu, reported the repeated hydrolysis of salmon fish

oil after distillation using Candida rugosa lipases in a 1 L stirred tank reactor resulted

in PUFA content of 38.7mol%, double of the initial level (Kahveci & Xu, 2011). These

approaches show the potential of lipase based methods to improve the nutritional value

of marine oils. Compared to existing chemical processes, lipase based approaches are

more environmentally friendly. Further research is required to obtain more selective

lipases and to optimise suitable production and immobilisation methods.

32

Page 54: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 1 1.6. Research aim and objectives

This research was focused on downstream processing of lipid and lipase from

thraustochytrids for the concentration of omega-3 fatty acids. The main objectives

were:

1. To establish a rapid lipid quantification method for the screening and

identification of thraustochytrid strains from marine samples collected from

Queenscliff, Victoria.

2. To determine which enzymatic activities are present in these thraustochytrid in

house isolates.

3. To develop methods for the extraction of total lipids and lipases from selected

thraustochytrids.

4. To investigate lipase production from thraustochytrid in a stirred tank reactor.

5. To develop a novel enzymatic method for concentrating omega-3 fatty acids

from marine oils, including thraustochytrid oils.

33

Page 55: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 2

Chapter 2: A quick colorimetric method for total lipid

quantification in thraustochytrids

34

Page 56: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 2 3.1. Introduction

Biodiesel is a renewable fuel of commercial interest, driven in part by the rising

cost of fossils and negative environmental impacts associated with conventional fuel

sources (Sadeghinezhad et al., 2014). First generation biodiesel was produced from

biomass, vegetable oils, animal fats and waste oils by transesterification (van Eijck et

al., 2014). The use of edible oils for first-generation biodiesel has resulted in the

undesired effect of increasing global competition for food crops (Kumar & Sharma,

2015; Pinzi et al., 2014). Second-generation biofuels utilise non-competing crops such

as jatropha, karanja, jojoba, mahua as well as the use of waste cooking oil, grease, and

animal fats. While the use of “inedible” crops reduces competition with food crops,

second-generation biofuels still compete for arable land, which ultimately negatively

impacts the world food supply and can contribute to the destruction of soil resources.

In addition, these feedstock cannot meet current energy demands (Nautiyal et al., 2014;

Zhu et al., 2014). Sustainable alternatives must be found that can provide the large fuel

volumes required, while not compromising the use of agricultural land for food

production. Microalgae have attracted attention as a means of addressing rapidly

growing demand for feedstock in biodiesel production (Baganz et al., 2014; Bellou et

al., 2014; Chen et al., 2015b). Microalgae are easy to cultivate and can be grown on

non-arable land, as well as accumulate more lipids as a percentage of their biomass

than plants (Challagulla et al., 2015).

Production of biodiesel from heterotrophic microalgae requires expensive

nutrients, particularly glucose as a carbon source (Slade & Bauen, 2013). Glucose can

constitutes 80 % of the total medium cost and up to a third of total fermentation

production costs (Li et al., 2007). To reduce process costs and assist in the

commercialisation of biodiesel production from heterotrophic microalgae, sustainable

low-cost carbon sources are required. One approach is to couple microalgae cultivation

with the use of inexpensive carbon sources such as carbon rich waste water, or use

glycerol derived from the biodiesel industry (Di Caprio et al., 2015; Xin et al., 2010).

Production costs can also be decreased by discovering new microalgae with improved

lipid productivity (Ahmad et al., 2015). An important reason for the slow development

in potential microalgae strain isolation and identification is due to lack of quick and

high-throughput assays (Cheng et al., 2011b). Existing techniques used for the

screening and quantification of microbial lipids include gravimetric, chromatographic

35

Page 57: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 2 and fluorescence dye methods (Bligh & Dyer, 1959; Han et al., 2011; Ren et al., 2015).

Analysis of microbial lipids by gravimetric method is usually effective, but has

limitations. The gravimetric method is time-intensive, requires the use of toxic

chemicals and utilises relatively high levels of biomass. Spectrophotometric

approaches can use a fluorescent dye, Nile Red, for lipid quantification. However, the

sensitivity of this method is poor and microbial cell components such as proteins,

pigments and some other environmental factors may interfere with lipid quantification

using these methods (Cheng et al., 2011b). A disadvantage of using fluorescent dye is

it can photo-bleach during exposure to light (Govender et al., 2012). Microbial lipid

quantification using triethanol-amine copper salts is an alternative method, however,

analyses of fatty acids in the ultraviolet region results without any colour development

(Chen & Vaidyanathan, 2012).

Colorimetric sulfo-phospho-vanillin (SPV) method for lipid quantification

firstly introduced by Charbo and Charonnat, 1937 for lipid quatification in human

cerebrospinal fluid samples (Vatassery et al., 1981). Later, this method was modified

for quantification of lipids in various samples include serum, foods and ecological

samples . Van handel modified SPV methods

into micro scale for total lipids quantification in single mosquitos (Van Handel, 1985).

Thraustochytrids, marine heterokonts and classified as oleaginous

microorganism, have been identified as a potential candidate for feedstock in the

heterotrophic production of lipids (Chen et al., 2015d). Extensive screening from

marine habitats has led to the discovery of strains which produce more than 50% of

their biomass as lipids (Raghukumar, 2008). In this study, we developed a colorimetric

sulfo-phospho-vanillin (SPV) method (initially developed for the serum lipids

quantification (Charbol & Charonnat, 1937)) for quantifying of lipids in

thraustochytrids and other oils containing materials. The accuracy of the SPV method

was compared with gravimetric and gas chromatography analysis. This method is

useful for rapid screening of microalgal strains for biodiesel production.

3.2. Materials and Methods

2.2.1. Chemicals used

All chemicals and medium components used in this study were procured from

Sigma-Aldrich, Australia. Instant ocean sea salts were procured from Aquarium 36

Page 58: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 2 Systems Inc., USA. Schizochytrium oil (batch number 11071602J) was obtained from

Ocean Nutrition Canada, Canada.

2.2.2. Cultivation of thraustochytrids

A total of four thraustochytrids, Schizochytrium sp. S31 (ATCC 20888),

Thraustochytrium sp. PRA 296, Thraustochytrium sp. AMCQS5-5 and

Schizochytrium sp. DT3 (in-house isolates) were selected for the current study.

Thraustochytrium S31 and PRA 296 were procured from the ATCC and used as

standard cultures, whereas AMCQS5-5 (JX993841) and DT3 (KF682125) were

isolated from Australian marine samples (Gupta et al., 2013). Seed culture was

prepared in GYP medium containing: Glucose 0.5 %, yeast extract 0.2 %, peptone 0.2

% in 50 % artificial seawater (Instant Ocean sea salts) and incubated at 150 rpm, 25

°C for 48 h. The inoculum (5 %) was later transferred into production media containing

glycerol (1 %), yeast extract (0.2 %) and peptone (0.2 %) and incubated at 150 rpm,

25 °C for 120 h. The cultures were harvested by centrifugation at 8000 rpm for 10 min,

freeze dried and stored at -20 °C for further use.

2.2.3. Standard curve preparation

For preparing a standard curve, commercial Schizochytrium oil was dissolved

in chloroform (10 mg in 10 ml) to a final concentration 1 mg mL-1. For the preparation

of the standard curve, different concentrations (15-90 μg) of lipid samples were taken

in clean glass tubes. The tubes were incubated at 60 °C for 10 min to evaporate the

chloroform. 100 μl of distilled water was added to each tube. Lipid quantification was

done by SPV reaction.

2.2.4. Quantification of thraustochytrid lipids by SPV

Vanillin phosphoric acid reagent was prepared by dissolving 0.120 g of vanillin

in 20 ml of distilled water and the final volume was adjusted to 100 ml with 85 %

phosphoric acid. This reagent was stored in the dark until further use. SPV reagent was

prepared fresh, which resulted in high activity with lipid samples. A known amount of

thraustochytrid dried biomass was suspended in distilled water and added to each test

tube (0.2-1 mg). Tubes were incubated at 100 °C for 10 min and cooled for 5 min in

an ice bath. Vanillin-phosphoric acid reagent (5 ml) was added to each tube and

37

Page 59: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 2 incubated in a shaker incubator at 200 rpm at 37 °C for 15 min. Absorbance was

measured at 530 nm.

The major steps involved in this method are: (a) phospho-vanillin reagent

preparation using phosphoric acid and vanillin, (b) addition of concentrated sulfuric

acid to a sample that contains lipids of interest, and heating of the mixture, (c) addition

of phospho-vanillin reagent, and (d) measurement of absorbance at 530 nm (Knight et

al., 1972). Fig. 2.1 illustrates in details the steps involved in the SPV method.

Figure. 2.1. Schematic diagram of SPV method. Concentrated sulphuric reacts with

unsaturated lipids microalgae and forms a carbonium ion. Phosphoric acid and vanillin

reacts and form a phosphate ester, which increases the reactivity of carbonyl group.

Carbonium ion and carbonyl group (from phospho vanillin reaction) reaction results

in formation of a stable colour compound.

2.2.5. Lipid extraction

Lipid was extracted and quantified from freeze dried biomass every 48 h

according to a previously published protocol (Lewis et al., 2000) with some 38

Page 60: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 2 modifications. Freeze-dried bio solvent (chloroform

and methanol in 2:1 ratio) and the sample vortexed for 2 min, followed by

centrifugation at 13,000 rpm for 15 min. This extraction process was repeated three

times. The three supernatants were collected and dried at 50 °C. The lipid percentage

was measured gravimetrically.

All the experiments were performed in triplicates. Results are represented as

mean values and standard deviation is < 10%.

2.3. Results and discussion

2.3.1. Biomass and lipid production

Production of biodiesel from inexpensive carbon sources will reduce oil

production costs. Glycerol is an economical carbon sources for microbial fermentation

(Leite et al., 2015) and has been tested as a carbon source for growing thraustochytrids

(Gupta et al., 2013; Scott et al., 2011). After 5 days, the biomass obtained were 938

mg L-1 d -1 in DT3, 950 mg L-1 d -1 in S31, 740 mg L-1 d -1 in AMCQS5-5 and 256 mg L-

1 d -1 in PRA-296 (Table 1). Schizochytrium sp. S31 and DT3 produced similar biomass

yields, whereas Thraustochytrium sp. AMCQS5-5 and PRA-296 exhibited lower

biomass productivity, depending on how efficiently the strain utilised the carbon

source to produce biomass. However, lipid productivity was documented to be higher

in DT3 at 287.96 mg L-1 d -1 when compared to other strains. The next highest biomass

was produced by S31 (227.05 mg L-1 d -1) followed by AMCQS5-5 (158.36 mg L-1 d -

1) and PRA-296 (51.71 mg L-1 d -1) (Table 1).

Table 2.1. Biomass and lipid production in glycerol medium by thraustochytrids.

Item DT3 S31 AMCQS5-5 PRA-296

Biomass (g L-1) 4.69 ± 0.09 4.75 ± 0.1 3.7 ± 0.01 1.28 ± 0.005

Biomass productivity (mg L-1 d -1) 938 ± 0.19 950 ± 0.02 740 ± 0.002 256 ± 0.001

Average lipid content (mg L-1) 1439 ± 92 1135 ± 68 791 ± 49 258 ± 15

Lipid productivity (mg L-1 d -1) 287 ± 28 227 ±17 158 ± 11 51 ± 4

39

Page 61: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 2 2.3.2 Microalgae oil quantification by SPV method

The colorimetric method was first validated to generate the calibration curve

with known quantities of Schizochytrium oil. A known quantity of lipid standard in the

range of 15-90 μg was taken in clean test tubes and placed in a hot air oven at 60 °C

to evaporate chloroform, followed by the SPV reaction protocol described in the

experimental section. Lipid content was measured by plotting absorbance at 530 nm

and lipid concentration. A high correlation coefficient was obtained with R2 value

0.9939 (Fig. 2.2). This demonstrates the high accuracy of the SPV method in

quantifying lipids in aqueous samples.

Figure. 2.2. Standard curve representing relationship between Schizochytrium oil

concentration and increase in the absorbance. The lipid concentrations were from 15

to 90 μg. Standard curve representing relationship between Schizochytrium oil

concentration and increase in the absorbance.

Interference by other cellular and medium components such as protein (BSA),

carbohydrate (Glucose) and glycerol did not occur with the SPV reagent (Fig. 2.3),

which is consistent with a previous study (Izard & Limberger, 2003). These medium

compounds were used as false positives. Thus, other compounds in the medium

showed no effect on the absorbance value.

y = 0.0033x - 0.0258R² = 0.9939

0

0.05

0.1

0.15

0.2

0.25

0.3

0 20 40 60 80 100

Abs

orba

nce

at 5

30 n

m

Schizochytrium oil (μg)

40

Page 62: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 2

Figure. 2.3. Comparison of various lipids and other compounds with sulfo-phospho-

vanillin reaction. Triolein yielded intense colour compared with other lipid standards.

Compounds with no bonds in their structure do not react with the SPV reagent.

The basic principle underlying the SPV method is a three step reaction. When

the lipids are heated in the presence of sulphuric acid at boiling temperature (low

temperature leads to lower reaction rates), a carbonium ion is formed. An aromatic

phosphate ester is formed from the reaction of vanillin and phosphoric acid. The

activated carbonyl group of phospho-vanillin reagent reacts with the carbonium ions

to form a stable pink colour, which has a maximum absorbance at 530 nm (Knight et

al., 1972). Colour formation depends mainly on the molecular structure of fatty acids

present in the lipids. The molecules which do not possess double bonds do not

detectably react with sulfo-phospho-vanillin reagent. Colour intensity varies from one

lipid to another due to differences in fatty acid structure. It was observed that the colour

intensity of the triolein was greater than that of the polyunsaturated fatty acids EPA

and DHA (Fig. 2). The presence of multiple double bonds in EPA and DHA did not

yield more colour due to steric hindrance near these double bonds. Saturated fatty acids

such as palmitate and stearate did not react with sulfo-phospho-vanillin reagent (Frings

& Dunn, 1970; Knight et al., 1972). Among the plant oils tested in this study, olive oil

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0.5

0 20 40 60 80

Abs

orba

nce

(530

nm

)

Different oil samples (μg)

SunflowerOliveCanolaTrioleinEPADHAPalmitateSterateGlycerolGlucoseBSA

False positives

Lipid samples

41

Page 63: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 2 resulted in higher levels of colour formation than did canola and sunflower oil,

probably due to the presence of more than 70% oleic acid in olive oil. In general, higher

levels of unsaturated fatty acids lead to higher absorbance, although the presence of

multiple double bonds in a single fatty acid (polyunsaturated) may lead to the

formation of intense colour (Knight et al., 1972). The presence of various fatty acids

in microbial lipids limits the accuracy of the method. This can be rectified by selecting

a suitable oil with the same fatty acid content as a reference. Previous studies have

reported the quantification of bacterial and microalgal lipids with reference to triolein

and canola oil, respectively (Izard & Limberger, 2003; Mishra et al., 2014). We used

commercial Schizochytrium oil as reference for quantification of lipids in

thraustochytrids.

2.3.3. Lipid quantification in thraustochytrid whole cells via SPV method

The established colorimetric method was applied to the quantification of lipids

in whole thraustochytrid cells. A known quantity of dried biomass of all four

thraustochytrid cultures was suspended in distilled water at a concentration of 20 mg

ml-1. Aliquots of 10-50 μl were taken in clean test tubes and the final volume was made

up to 100 μl with distilled water. Fig. 3 shows the correlation coefficient between

biomass concentration and absorbance with an R2 > 0.99. All four cultures showed an

increase in the intensity of pink colour formation with respect to the increasing

biomass concentration (Fig. 2.4). Lipid content in all four cultures was quantified by

converting the absorbance into mass using a reference curve with a low standard

deviation. Colour change from low concentration to high concentration of biomass

(lipid) in four different thraustochytrids shown in Fig. 2.5.

42

Page 64: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 2

Figure. 2.4. Profile of colorimetric detection of thraustochytrid whole cell lipids. The

relationship between the concentration of biomass and absorbance at 530 nm for

DT3 (R2= 0.992), S31 (R2=0.999), AMCQS5-5 (R2=0.993) and PRA-296 (R2 =

0.998)

y = 0.993x + 0.0039R² = 0.9926

y = 0.888x - 0.0154R² = 0.999

y = 0.7308x + 0.0138R² = 0.9932

y = 0.516x + 0.0121R² = 0.9985

0

0.2

0.4

0.6

0.8

1

1.2

0 0.2 0.4 0.6 0.8 1 1.2

Abs

orba

nce

at 5

30

Biomass (mg)

DT3 S31AMCQS5-5 PRA 296

43

Page 65: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 2

Figure. 2.5. Change of lipid colour after reaction with sulfo-phospho-vanillin reagent.

Increase of colour represents the correlation between the concentrations of lipid in

thraustochytrids. (a) DT3, (b) S31, (c) AMCQS5-5 and (d) PRA-296.

Strain development and lipid quantification are major milestones on the path

to sustainable production of biodiesel from the microbial sources. A rapid lipid

quantification technique is required to screen potential lipid producing strains from

environmental samples. In this study, we used a colorimetric method (SPV) to quantify

the total lipids in an aqueous medium containing thraustochytrids, a potential lipid

producer. The lowest amount of biomass used for quantifying the lipid content was 50

μg. This enables us to quantify the lipids from the thraustochytrids cells without an

extraction step. Moreover, the colorimetric method not only quantified lipid content

accurately, but also required less than one hour to complete. In contrast, gravimetric

quantification of lipids, extraction and GC analysis are more time and labour intensive.

2.3.4. Validating the accuracy of the SPV method

Gravimetric quantification of lipids from the four selected thraustochytrids was

used in order to validate the accuracy of the SPV method. Thraustochytrid lipids were

44

Page 66: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 2 extracted using a conventional solvent extraction method with modifications (Lewis

et al., 2000). The total lipid percentages determined gravimetrically from the four

thraustochytrids DT3, S31, AMCQS5-5 and PRA-296, were 33.3, 25.8, 22.8 and 18.0,

respectively. Lipid percentages determined also by the SPV method showed similar

quantities in all thraustochytrids. Lipid contents determined by SPV, gravimetric

methods and gas chromatography were compared (Table 2.2). The lipid profiles

determined by gas chromatography were 27.3, 21.8, 20.3 and 18.4, respectively. In

two thraustochytrid strains, AMCQS5-5 and PRA 296, the lipid percentages were the

same as those obtained using the SPV method. However, in DT3 and S31 there was

small (1.9-2.6 %) difference in the lipid percentages. These results obtained from

standard gravimetric method for lipid quantification compared with the SPV method

in quantifying the total lipids from thraustochytrid whole cells, and the variation was

less than 3%, which confirms the suitability of SPV method in screening microbial

samples.

Table 2.2. Comparison of thraustochytrid lipid content by SPV, gravimetric and GC

methods.

Thraustochytrid

strains

Lipid

percentage

by SPV

method

Lipid

percentage by

gravimetric

method

Variation Lipid

percentage by

gas

chromatography

DT3 33.3 ± 1.2 30.7 ± 0.59 2.6 27.3 ± 1.3

S31 25.8 ± 0.9 23.9 ± 0.86 1.9 21.8 ± 0.84

AMCQS5-5 22.8 ± 1.1 21.4 ± 0.78 1.4 20.36 ± 0.9

PRA-296 18.0 ± 1.4 20.2 ± 0.62 2.2 18.47 ± 0.81

Microalgal lipids are useful for biodiesel production. Isolation and

identification of the high lipid-producing microalgae is a key step in achieving

economically sustainable biodiesel production. Quantification of lipids from

microalgae must be performed frequently in microbial screening and other research

areas, such as in the screening of genetically engineered microbes for high lipid

production and so higher throughput accurate methods are important.

45

Page 67: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 2 2.4. Conclusion

In conclusion, this study demonstrated the use of the sulfo-phsopho-vanillin

reaction to quantify lipid accumulation in thraustochytrids. The SPV method resulted

in similar lipid content determination in thraustochytrid whole cells as did gravimetric/

gas chromatography analysis. Triolein contains only mono-unsaturated fatty acids and

the colour intensity was greater than for polyunsaturated fatty acids with multiple

double bonds. Carbohydrates, proteins and glycerol did not interfere with the SPV

reaction. SPV is a rapid method that requires low levels of chemicals and so is useful

for the screening of microalgal cultures for biodiesel production.

46

Page 68: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3

Chapter 3: Comparison of cell disruption methods for

improving lipid extraction from thraustochytrid strains

47

Page 69: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3 2.1. Introduction

The limited availability of fossil fuels and increasing greenhouse gases

emission has led to interest in alternative biomass derived energy. As an alternative

fuel, biodiesel is promising (Azeem et al., 2016). Biomass derived biofuels are

classified as solid (bio-char), liquid (ethanol, vegetable oil, and biodiesel), and gaseous

(biogas, biosyngas and biohydrogen) fuels (Mubarak et al., 2015). Biodiesel is

renewable and a relatively clean energy in terms of carbon dioxide generation and

greenhouse gas emissions. Biodiesel accounts for 10% of total biofuel production

globally and its estimated production is about 6 billion litres/year (Nogueira, 2011).

The present approach to biodiesel (defined as the monoalkyl esters of long-chain fatty

acids) production involves transesterification of plant oils such as soybean oil,

sunflower oil and rapeseed oil with methanol using alkali catalysts (Daroch et al.,

2013; Meher et al., 2006). The use of plant oils for biodiesel production has been

associated with some drawbacks such as high viscosity, low volatility and deposition

in combustion chambers (Lin et al., 2011). Also, the use of edible oils for biodiesel

production has led to an increase in the price of oils, initiating a food versus fuel debate

regarding biofuel sustainability. Feedstock useful for biofuel include soapstocks, acid

oils, tallow oils, used cooking oils, various animal fats, non-edible plant oils and

microbes including algae (Chisti, 2007; Silitonga et al., 2013).

Microalgae are promising vehicles for the production of biodiesel and possess

advantages such as higher growth rate and productivity, grow in various environments

(fresh, brackish or salt water), do not compete for land, and have high oil productivity

(20-50% by dry weight basis) compared to conventional crops (Singh et al., 2011).

Microalgae lipid content differs between microalgae species. Microalgae lipid

composition and fatty acids profiles depends on culture media composition (carbon

source and nitrogen source), temperature, pH and aeration (Halim et al., 2012a). The

process of lipid extraction from microalgae is an energy intensive and expensive

because the use of solvents and recovery methods in downstream processing.

Selection of efficient microalgae species and suitable lipid extraction methods are

important for commercial biodiesel production (Griffiths & Harrison, 2009; McMillan

et al., 2013). Biodiesel production from microalgae involves four major stages that are

cultivation, cell harvest, lipid extraction and finally conversion of lipids into biodiesel

(Lee et al., 2010b; Schnurr & Allen, 2015). Therefore, a suitable lipid extraction

48

Page 70: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3 technique is a prerequisite for microalgal lipid extraction. Lipid extraction efficiency

is dependent on the polarity of the solvent and combination of solvent mixture (Lee et

al., 2010b; Lee et al., 1998; Lewis et al., 2000). The combination of a polar and non-

polar solvent mixture can in some cases extract more lipids from microalgae

(Ramluckan et al., 2014; Ryckebosch et al., 2012b). For example, the Bligh and Dyer

method uses chloroform and methanol for lipid extraction from a range of biological

samples (Ramluckan et al., 2014). The use of chloroform and ethanol in a 1:1 ratio

provided maximum lipid extraction from Chlorella sp. (Ramluckan et al., 2014),

whereas, a combination of dichloromethane and ethanol increased lipid extraction

efficiency by 25% in the same organism (R et al., 2015).

Microalgae cell walls block the release of lipid present inside the cells. To get

higher product recovery and quality lipids with lower operating costs from microbial

cells, a suitable cell disruption method is required. Cell disruption enhances the release

of intracellular lipids from microalgae by improving the access of the extracting

solvent to fatty acids (Halim et al., 2012b). Cell disruption methods used in microalgae

lipid extraction are classified as chemical and mechanical. Chemical cell disruption

methods involve use of acids, alkalis, enzymes or osmotic shocks and mechanical

methods use microwave, ultrasonication, bead mill, drying, or supercritical fluid

extraction. The choice of method influences lipid extraction yields from a range of

microalgae (Lee & Han, 2015; Lee et al., 2010b; McMillan et al., 2013; Zheng et al.,

2011).

The aim of this study was to compare various organic solvents and cell

disruption methods for effective lipid extraction from a new strain of thraustochytrids

(a newly isolated strain from the Queenscliff region, Victoria, Australia). This strain

was used as a representative microalgae in this work due to its ability to accumulate

high levels of lipids (Gupta et al., 2013). In addition, microalgae biomass was used for

different extraction methods that were successfully used for efficient algal lipid

extraction in previous studies. Finally, the conversion of lipids to fatty acid methyl

ester (FAME) was performed to determine yield and fatty acid distribution after

extraction.

3.2. Experimental Section

3.2.1. Chemicals

49

Page 71: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3

All the chemicals used in this study were of analytical grade. Medium

components such as glucose, yeast extract and mycological peptone (Sigma-Aldrich,

USA) and sea salt (Instant Ocean, USA) were used for biomass production while

solvents such as acetone, ether, hexane (Merck, Australia), methanol and ethyl acetate

from Fischer and Honeywell, Australia.

3.2.2. Strain selection and biomass production

Schizochytrium sp. S31 (ATCC 20888) was procured from American Type

Culture Collection (ATCC) and used as standard culture. Thraustochytrids used in this

study were maintained on GYP medium consisting (g L-1): glucose 5, yeast extract 2,

mycological peptone 2, agar 10 and artificial seawater 50% at 25 °C and sub-cultured

after 15 days.

Thraustochytrium sp. AMCQS5-5 (an in house isolate; Genbank accession

number JX993841), was grown in a medium containing (g L-1): glucose 5, peptone 2,

yeast extract 2 and artificial sea water 50% for inoculum preparation with shaking at

150 rpm for 2 days at 25 oC. The medium was autoclaved at 121 oC for 20 min.

Inoculum (5% v/v) was used to inoculate production medium (100 ml contained in 500

ml flask) and incubated for 5 days in a shake flask at 25 oC and 150 rpm. The resultant

biomass was harvested by centrifugation (4000 rpm for 15 min) and was freeze-dried

and kept at -20 oC until further use.

The thraustochytrids grown in culture medium were harvested at the interval

of 24 h up to 120 h. Optical density at 600 nm and dry cell weight (DCW) was

measured at 24 h intervals. A calibration curve was plotted between OD and dry cell

weight. Results are presented as mean ± SD of duplicates repeated twice.

The biomass and lipid productivity was calculated from the formula mentioned below:

Productivity = Biomass or Lipid content/T1-T0

Where, T1 = Final day of biomass harvesting and T0 = Initial day of incubation

3.2.3. Lipid extraction from thraustochytrids by organic solvents

The nine solvents, chloroform, dicholoromethane, diethylether, ethanol,

heptane, hexane, isopropanol, methanol and toluene at ratios of 1:1, 1:2 and 2:1 were

tested for maximising extraction of total lipids. For solvent extraction, 50 mg of freeze

50

Page 72: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3 dried biomass of thraustochytrid was blended with 3 ml of various solvents. The

mixture was vortexed for 2 min and the sample was then centrifuged at 4000 rpm for

15 min. Supernatant (organic phase) was carefully collected in the pre-weighed glass

vials and the solvent was evaporated under nitrogen gas at room temperature. Lipid

content (% dry weight basis) was determined gravimetrically. To determine the

optimal organic solvent mixture for lipid extraction from thraustochytrids, different

ratios of the best three single solvents were investigated. For chloroform-methanol, the

Bligh and Dyer method was followed (Bligh & Dyer, 1959).

3.2.4. Cell disruption for lipid extraction

Freeze dried biomass was blended with 3 ml of chloroform and methanol (2:1)

and disrupted by means of different cell disruption methods as detailed below. After

each treatment, lipid extraction from thraustochytrids was done according to Gupta

and co-workers (Gupta et al., 2012c). After centrifugation (4000 rpm for 15 min), the

upper layer was collected and dried under nitrogen gas. Lipid content (% dry weight

basis) was determined gravimetrically.

3.2.4.1. Grinding with liquid nitrogen

A sample of freeze dried thraustochytrid biomass (50 mg) was taken in the

ceramic mortar. About 10-15 ml liquid nitrogen was added and the sample was allowed

to thaw and grinded with pestle for 2 min. After grinding the lipid was extracted using

organic solvents.

33.2.4.2. Bead vortexing

Thraustochytrid cell suspension (50 mg) was taken in glass tube (35 ml) and 3

ml of solvent and 1 ml of beads (zirconia beads, size 0.4-0.6 mm, Klausen Pty Ltd,

Australia) were added and contents were vortexed for 20 min, using a vortex in 30

second bursts. Samples were kept on ice in between the bursts and lipid was extracted

using organic solvents.

3.2.4.3. Sonication

Thraustochytrid cell biomass (50 mg) was suspended in 3 ml of solvent in a 15

ml centrifuge tube. Sample was sonicated at 20 kHz, 40% amplitude and the pulse was

51

Page 73: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3 40 seconds on and 20 seconds off with total working time of 20 min (Sonics, USA).

Sample tubes were kept in ice during the sonication process to prevent overheating.

3.2.4.4. Osmotic shock

Thraustochytrid suspension was disrupted using an osmotic shock method. 50

mg of thraustochytrid biomass was suspended in 3 ml of 10% NaCl solution (mass

concentration 16.6 Kg/m-3, 0.6 Kg of NaCl) and vortexed for 2 min and incubated for

48 h at room temperature, followed by solvent extraction. Cell disrupting pressure was

calculated by the Morse equation:

kPa), i is van 't Hoff factor, M is molar concentration of NaCl, R is 0.0821 L atm/K

mol, and T is absolute temperature in °K at which the osmotic shock was performed.

Osmotic shock was calculated using an online osmotic pressure calculation hosted by

Georgia State University.

3.2.4.5. Water bath

3 ml samples containing 50 mg of biomass taken in 15 ml centrifuge tube was

placed in preheated water bath (Ratek Instruments Pty Ltd, Australia) to induce

thermolysis. Samples were kept in a water bath for 20 min at 90 °C. Three tubes were

treated simultaneously as replicates without shaking.

3.2.4.6. Shake mill

Thraustochytrid cell suspension was disrupted in plastic sample bottle using

shake mill (SPEX Mill 8000M, USA). The biomass and bead ratio was 3:1 ratio

(Zirconia beads 0.4-0.6 mm) at maximum speed (1060 cycle/min) and exposed for 5

min.

All the cell slurries were observed under the microscope using differential

interference contrast (Axio-imager, Zeiss, Germany) to check the disruption of

thraustochytrid cells. Microbial smear was prepared on the glass slide, air dried and

observed under a microscope. All experiments were performed three times.

52

Page 74: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3

3.5. Fatty acids methyl esters (FAMEs) analysis

Fatty acids were converted to methyl esters by acid-catalysed trans-

esterification according to a published method (Han, 2010). 1 ml toluene was added to

methanol (2 ml) was also added to the tube and kept for overnight incubation at 50 °C.

Fatty acid methyl esters (FAMEs) were extracted into hexane. The hexane layer was

removed and dried over sodium sulphate. FAMEs were concentrated using nitrogen

gas. The samples were analysed by a GC-FID system (Agilent Technologies, 6890N,

Santa Clara, CA, USA). The GC was equipped with a capillary column (SGE, BPX70,

ier gas at a flow rate

of 1.5 ml min 1. The injector was maintained at 250 °C and a sample volume

was injected. Fatty acid peaks were identified on comparison of retention time data

with external standards (Sigma-Aldrich, St. Louis, MO, USA) and corrected using

theoretical relative FID response factors (Ackman, 2002). Peaks were quantified with

Chemstation chromatography software (Agilent Technologies, Santa Clara, CA,

USA). Results are presented as mean ± SD of triplicates.

3.3. Results and Discussion

3.3.1. Biomass and lipid production

Schizochytrium sp. S31 showed highest biomass productivity at 0.81 g L-1 d-1

and Thraustochytrium sp. AMCQS5-5 at 0.64 g L-1 d-1 at the end of 5 days (Table 1).

The lipid productivity of Schizochytrium sp. S31 and Thraustochytrium sp. AMCQS5-

5 were 100.74 mg L-1 d-1 and 64.2 mg L-1 d-1, respectively. Selection of suitable

microalgae strain with adequate biomass and oil productivity is important for cost

effective biodiesel production. Our results on biomass and lipid productivity are in

agreement with previous findings by Vello et al. 2014 (0.30 g L-1 day-1 and 34.53 to

230.38 mg L-1 day-1) that demonstrated the suitability of Chlorella strain as a

promising candidate for biodiesel production (Vello et al., 2014).

53

Page 75: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3 Table 3.1. Biomass, lipid contents and productivity of thraustochytrids.

Properties Thraustochytrid strains

Schizochytrium sp. S31 Thraustochytrium sp. AMCQS5-5

Dry weight (g L-1) 4.06 3.23

Biomass productivity (g L-1 d-1) 0.88 0.64

Average lipid content (mg L-1) 503.7 321.3

Lipid productivity (mg L-1 d-1) 100.74 64.2

Lipid content and fatty acid profile of a microorganism is dependent on the

growth conditions (Rodolfi et al., 2009; Salama et al., 2013). Medium composition

influences the percentage of total lipid and type of fatty acids in the microbe. For

example, the addition of tween 80 in the production medium led to the accumulation

of oleic acid in the thraustochytrids (Taoka et al., 2011). A recent study reported the

effect of seasonal variation and nitrogen limitation in the total lipid production and

fatty acid composition of Nannochloropsis oculata. They observed an increased

accumulation (up to 90%) of saturated and mono-unsaturated fatty acids (Olofsson et

al., 2014). Since the aim of this work was to find a suitable disruption method for lipid

extraction, higher biomass productivity was not pursued further.

3.3.2. Lipid extraction from thraustochytrid by organic solvents

The lipid extraction by organic solvents was examined to confirm the lipid-

extraction characteristics of thraustochytrids. Fatty acids present in the lipid govern

the polarity, based on the principle “like dissolves like”, thus a suitable solvent should

be identified for total lipid extraction, however a universal solvent cannot be applied

to all microbes with different fatty acid composition. Total lipid extraction yields vary

primarily with solvent polarity (Phukan et al., 2011).

To understand the efficacy of organic solvents in lipid extraction from

Schizochytrium sp. S31, nine solvents and their combinations (based on high lipid

yield) were selected, since the lipid extraction is highly dependent on the polarity of

the solvents and their ratios (Ramluckan et al., 2014). The effect of single solvent as

54

Page 76: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3 well as the combinations of the best 3 solvents with respect to the lipid extraction from

Schizochytrium sp. S31 is shown in Figure 1. Among the single organic solvents,

maximum lipid (12.5%) was extracted with hexane, followed by 11 % in heptane and

9.7% in chloroform. However, among the combination of solvents, the mixture of

chloroform: methanol (2:1) showed maximum lipid extraction (22%), followed by

chloroform and hexane (2:1) with 13.4% (Fig. 3.1). Chloroform and hexane showed

comparatively low efficacy in lipid extraction indicating that a range of polar to non-

polar lipids were present in Schizochytrium sp. S31. A similar trend was observed

during lipid extraction from Chlorella sp. (Ramluckan et al., 2014). A mixture of

hexane:heptane (1:1) extracted the least amount of lipid (2.2%). Higher lipid yields

with chloroform: methanol indicates the presence of more polar and neutral lipids in

the algae. It was observed that the percentage of polar and neutral lipids were

approximately 78% and non-polar lipids were 22% in some algae (Kale, 2013).

Combination of polar and non-polar solvents could extract more lipids than individual

solvents (Lewis et al., 2000; Ryckebosch et al., 2012b). However, a study (Shen et al.,

2009) contradicts this conclusion, indicating that the combination of hexane and

ethanol could not extract more lipid than hexane alone from Scenedesmus dimorphus

and Chlorella protothecoides. This suggests the efficiency of lipid extraction depends

on the algal species and their lipid compositions. Also, results obtained from

chloroform and methanol (2:1) indicate the presence of more polar and neutral lipids

in algae (Ramluckan et al., 2014). Mixtures of chloroform:methanol extract

hydrocarbons, carotenoids, chlorophyll, sterols, triglycerides, fatty acids,

phospholipids and glycolipids (dos Santos et al., 2015; Singh et al., 2013).

55

Page 77: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3

Figure. 3.1. Lipid extraction percentage from dry biomass using various solvents from

Schizochytrium sp. S31. The symbols represent Chl: Choloroform, Hex: Hexane, Hep:

Heptane, Met: Methanol.

3.3.3. Comparison of lipid extraction methods

Six methods were evaluated in order to determine the efficiency of cell

disruption methods for total lipid extraction from thraustochytrids. The effectiveness

of the cell disruption methods were quantified using lipid yield percentages. Cell

disruption breaks the cells and improves the accessibility to the intracellular

components for extraction (Wang et al., 2015). Figure 2 shows the percentage of total

lipids extracted as a function of cell disruption methods. All the cell disruption

methods used in this study were able to disrupt thraustochytrid cells, although lipid

yield varied.

0

5

10

15

20

25

30

Tota

l lip

id c

onte

nt (d

cw, %

)

Solvents and their combination used

56

Page 78: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3

Figure. 3.2. Effect of different cell disruption methods for lipid extraction from

thraustochytrids. Average lipid extraction of each method was reported as % of total

lipids extracted.

Thraustochytrid cells subjected to disruption resulted in rupture of the cell

walls and release of intracellular components (Figure 3.3a and b). It was observed that

lipid content of Schizochytrium sp. S31 was higher than that of Thraustochytrium sp.

AMCQS5-5. Maximum lipid was extracted from both Schizochytrium sp. S31 (48.7%)

and Thraustochytrium sp. AMCQS5-5 (29.1%) cells using osmotic shock (Figure 2).

Grinding, sonication, shake mill and water bath treatments extracted 44.6%, 31%,

30.5% and 20.8% of lipids, respectively. Bead vortexing resulted in 25% lipid

extraction from Thraustochytrium sp. AMCQS5-5 cells, whereas other methods (water

bath, grinding, shake mill and sonication) resulted in lower yields. Osmotic shock

resulted in a 2.2- folds increment in lipid extraction from Schizochytrium sp. S31 and

a 2.8-folds increase from Thraustochytrium sp. AMCQS5-5, compared to control. A

similar study for a Chlorella sp. indicated that osmotic shock was an effective method

for extracting lipids (Prabakaran & Ravindran, 2011). This method also consumes less

energy than traditional methods. In a recent study, when Chlamydomonas reinhardtii

cells were incubated in a high osmotic environment, it led to a 2-folds improvement in

lipid extraction (Yoo et al., 2012). Table 3.2 summarises some cell disruption methods

reported for lipid extraction from microalgae. Due to differences in cell wall structure,

0

10

20

30

40

50

60

Control Grinding Beadvortexing

Osmoticshock

Water bath Sonication Shake mill

Tota

l lip

ids c

onte

nt (%

)

Cell disruption methods

S31 AMCQS5-5

57

Page 79: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3 not all microalgae respond the same to pretreatment. For example, Lee et al. (2010)

observed that microwave treatment was optimum for the disruption of Botryococcus

sp., Chlorella vulgaris and Scenedesmus sp. cells. Another study showed that grinding

with liquid nitrogen facilitated higher levels of lipid extraction from Chlorella vulgaris

(Zheng et al., 2011). Available literature suggests that cell disruption methods improve

lipid extraction from microalgae, and efficiency depends on microalgae species, age

of the culture and composition of cell wall. Therefore, results obtained from one

species cannot be generalised to all other species (Halim et al., 2012b).

Figure. 3.3. Effect of cell disruption on thraustochytrids cells. Thraustochytrid cells

before (A) and after (B) cell disruption (scale bar 20 μm).

Table 3.2. Comparison of cell disruption methods employed to extract total lipids from

different microalgae.

S.

No.

Cell

disruption

methods

used

Efficient

method

Organism used Lipid

content

(%)

Reference

1 Autoclaving

Bead beating

Microwaves

Sonication

Osmotic

shock

Microwaves Botryococcus sp.

Chlorella vulgaris

Scenedesmus sp.

28.6

11

11.5

(Lee et al.,

2010b)

58

Page 80: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3 2 Sonication

Osmotic

shock

Microwave

Autoclave

Bead beating

Sonication Chlorella sp.

Nostoc sp.

Tolypothrix sp.

20.1

18.2

14

(Prabakaran

&

Ravindran,

2011)

3 Grinding

Sonication

Bead Milling

Enzymatic

Lysis

Microwaves

Grinding Chlorella vulgaris 29 (Zheng et

al., 2011)

4 Grinding

Bead

vortexing

Osmotic

shock

Water bath

Sonication

Shake mill

Osmotic

shock

Schizochytrium

sp. S31,

Thraustochytrium

sp. AMCQS5-5

48.7

29.1

This study

There are many lab scale cell disruption methods discussed in the literature, however,

only a few mechanical methods either alone or with the intervention of enzymes/

chemicals can be scaled up for industrial applications. For instance, bead mill, high

pressure homogenizer and Hughes press are used extensively at large scale, which

reduce unit operation steps compared to chemical and enzymatic methods (Chisti &

Moo-Young, 1986; Schutte & Kula, 1990). An osmotic shock method was

implemented in thraustochytrid cell disruption and lipid extraction, to reduce the

energy consumption and production cost. During the osmotic shock treatment, the

resulting wastewater can be recycled through reverse osmosis technology (Arnal et al.,

2005). The same method has been applied at pilot-scale for enhancing the release of

ectoine (Sauer & Galinski, 1998). A recent study by Jayaranja and Rekha (2015)

indicated that osmotic shock was a suitable method for maximising the extraction of

intracellular products, and this method can also be industrially scaled up (Kar &

59

Page 81: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3 Singhal, 2015). The advantages and disadvantages of selected investigated methods

are summarised in Table 3.3.

Table 3.3. Advantages and disadvantages of the investigated cell disruption methods.

Cell disruption

methods

Advantages Disadvantages

Manual grinding Quickest and efficient

- 2 min process

- Localised heating caused

denaturation of molecules

Bead vortexing - Can be established easily

and relatively effective

- High heat generation,

- Incomplete cell lysis

Osmotic shock - Lower energy

consumption

- Easier scale-up

- Generation of waste salt

water

- Time consuming

Water bath - Maximum disruption

- Easy in handling at lab

scale

- Increases the viscosity

- Energy intensive

Sonication - Faster extraction

- suitable for all cell type

- Damage chemical

structure of molecules

Shake mill - Rapid method - High energy intensive

- High heat generation

3.3.4. Fatty acid composition of extracted lipid

The fatty acid profiles of the lipids extracted following different cell disruption

methods from thraustochytrids are presented in Figure 4a and 5a. Major fatty acids

such as myristic acid (28.1%), palmitic acid (27.3%), palmitoleic acid (20.7%) and

oleic acid (12.8%) were detected in Schizochytrium sp. S31 (Figure 3.4a) after osmotic

shock disruption. Total saturated, monounsaturated and polyunsaturated fatty acid

contents after osmotic shock facilitated extraction were 58.7%, 34.6% and 4.8%,

respectively, making it a potential feedstock for biodiesel production (Figure 3.4b).

The other prominent fatty acids based on grinding, sonication and shake mill identified

in the lipid extracts were saturated (49-57%), mono-unsaturated (31-35%) and

polyunsaturated fatty acids (2-18%). Saturated and monounsaturated fatty acids are

useful major components for microalgal biodiesel production because of their

60

Page 82: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3 relatively high oxidative stability (Cao et al., 2014). The general properties of biodiesel

such as viscosity, specific gravity, cetane number, iodine value, and low temperature

performance metrics are determined by the structure (length and unsaturation) of fatty

acid esters (Hoekman et al., 2012).

Figure. 3.4a. Effect of different cell disruption methods on fatty acid profiles of

Schizochytrium sp. S31.

0

5

10

15

20

25

30

C14:0 C15:0 C16:0 C16:1n7 C18:0 C18:1n9C18:1n7C20:5n3C22:5n6C22:6n3

Tota

l lip

id c

onte

nt (%

)

Major fatty acids

Control Grinding with liquid N2Bead vortexing Osmotic shockWaterbath SonicationShake Mill

61

Page 83: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3

Figure. 3.4b. Effect different cell disruption methods on SFA, MUFAs and PUFAs of

Schizochytrium sp. S31.

Figure. 3.5a. Effect of different cell disruption methods on fatty acid profiles of

Thraustochytrium sp. AMCQS5-5.

0

10

20

30

40

50

60

70

SFAs MUFAs PUFAs

Tota

l lip

id c

onte

nt (%

)

Fatty acids composition

Control Grinding with liquid N2Bead vortexing Osmotic shockWaterbath SonicationShake Mill

0

10

20

30

40

50

60

C16:0 C18:0 C18:1n9 C18:1n7 C20:4n6 C20:5n3 C22:5n6 C22:5n3 C22:6n3

Tota

l lip

id c

onte

nt (%

)

Major Fatty acids

Control Grinding with liquid N2Bead vortexing Osmotic shockWaterbath SonicationShake Mill

62

Page 84: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3

Figure. 3.5b. Effect of different cell disruption methods on SFA, MUFAs and PUFAs

of Thraustochytrium sp. AMCQS5-5.

In Thraustochytrium sp. AMCQS5-5, palmitic acid (31.6%) and docosahexaenoic acid

(31.5%) were the major fatty acids (Figure 3.5a), with other polyunsaturated fatty acids

(C20:4n6, C20:5n3, C22:5n6 and C22:5n3) ranging from 2-9%, when cells were

disrupted using osmotic shock. Saturated, monounsaturated and polyunsaturated fatty

acid contents were 35.5%, 7.4% and 51%, respectively (Figure 3.5b). The highest

polyunsaturated fatty acid percentages were extracted by grinding and sonication

methods, 67% and 57%, respectively. Only shake mill resulted in a significant

percentage of monounsaturated fatty acids extraction (37.5%). Most methods resulted

in different lipid yields, but no difference in the fatty acid profiles for extraction from

Thraustochytrium sp. AMCQS5-5 (Figure 3.5b). If the omega-3 fatty acids could be

separated from other fatty acids in the Thraustochytrium sp. AMCQS5-5 extract, then

these fatty acids, particularly DHA, could be used as nutritional products and offset

the cost of biofuel production in this strain (Ryckebosch et al., 2012a; Singh et al.,

2015b).

3.3.5. Prediction of biodiesel properties

We analysed fatty acid profiles of two different thraustochytrid strains to

determine the suitability of these microalgae for biodiesel production. The fatty acid

0

10

20

30

40

50

60

70

80

SFAs MUFAs PUFAs

Tota

l lip

id c

onte

nt (%

)

Fatty acid composition

Control Grinding with liquid N2Bead vortexing Osmotic shockWaterbath SonicationShake Mill

63

Page 85: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3 values were taken as an input in predicting the biodiesel properties by using an open

access software "Biodiesel Analyzer© Ver. 1.1" (Talebi et al., 2013). Some of the

important parameters of biodiesel are cetane number (CN), Iodine value (IV) and

oxidative stability (OS), which determine the combustion behavior, quality of

biodiesel, stability and performance, respectively (Islam et al., 2013). The CN and IV

values in Schizochytrium sp. S31 (47.4 and 98.61) when compared to

Thraustochytrium sp. AMCQS5-5 (44.01 and 157.44) were observed. According to

ASTM D6751-12, the standard values were 47-51 (CN) and 120 g I2/100g maximum,

indicating suitability of selected strain S31 for biodiesel, however, its further

characterization will be a follow up study. The OS value of 6.5 for Schizochytrium sp.

S31 was higher than 1.65 for Thraustochytrium sp. AMCQS5-5 suggesting oxidation

stability decreased with the increase of polyunsaturated fatty acid content. Other

properties such as saturated fatty acids (SFA), monounsaturated fatty acids (MUFA),

polyunsaturated fatty acids (PUFA), degree of unsaturation (DU), saponification value

(SV), long chain saturated fatty (LCDF), cold filter plugging point (CFPP), cloud point

(CP), allylic position equivalent (APE), bis-allylic position equivalent (BAPE),

Table 3.4. Predicted biodiesel properties of given thraustochytrids strains.

Properties Units Strain S31 Strain AMCQS5-

5

Saturated fatty

acid

% (m/m) 59 44

Monounsaturated

fatty acids

% (m/m) 35 4.6

Polyunsaturated

fatty acids

% (m/m) 14 33

Degree of

unsaturation

63 70

Saponification

value

mg KOH/g oil 233.98 164.73

Iodine value g I2/100 g 98.61 157.44

64

Page 86: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3 Cetane number min 47.44 44.01

Long chain

saturated factor

% (m/m) 3.8 5.85

Cold filter plugging

point

oC 4.54 1.90

Cloud point oC 9.74 10.79

Allylic position

equivalents

83 167.60

Bis-Allylic position

equivalents

70 165.20

Oxidation stability h 6.5 1.65

Higher heating

value

oC 42.14 32.33

Kinematic

Viscosity

mm2/s 1.28 0.94

Density Kg/m3 0.94 0.73

3.3.6. Energy analysis

Energy consumption of the investigated cell disruption methods was carried

out to establish their potential as large scale processes. The comparative estimated

energy consumptions and processing times of the investigated cell disruption methods

are presented in Table 5. In comparison, water bath and sonication methods resulted

in highest energy consumption, 2400 MJ Kg-1 dry mass and 1200 MJ Kg-1 dry mass,

respectively. Shake mill energy consumption was estimated to be 690 MJ Kg-1 dry

mass. The osmotic shock method consumed modest energy (4.8 MJ Kg-1) and highest

lipid recovery, and so was the preferred method. NaCl present in the lysate solution

exerted osmotic pressure which was estimated to be 4.21 KPa, using an online osmotic

pressure calculator. Another study showed that the osmotic pressure of 1.9 KPa was

enough to break the microbial cells (Kar & Singhal, 2015). It has been demonstrated

that energy consumption for the microwave (28 MJ Kg-1) and ultrasound (44 MJ Kg-

1) methods enhanced lipid extraction from Chlorella sp. (Olofsson et al., 2014), which

shows that osmotic shock consumed less energy than either of these methods.

65

Page 87: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 3 Table 3.5. Energy consumption comparison.

Cell disruption methods Lipid yield

(%)

Energy consumption

(MJ Kg-1 dry mass)

Processing time

(min)

Control 22.04 Nila 0

Manual grinding 44.6 ND b 2

Bead vortexing 22.8 48 20

Osmotic shock 48.7 4.8 c 2

Water bath 20.8 2400 20

Sonication 31.05 1200 20

Shake mill 30.5 690 5

Thraustochytrids mass concentration of 16.6 kg/ m-3 was used for energy analysis.

Where Nila represents no energy was consumed; NDb represents physical effort cannot

be quantified; cOsmotic pressure by virtue of salt addition

3.4. Conclusion

In this study, we have shown that some cell disruption methods, particularly

osmotic shock, result in both different oil yields and variation in the percentage of

saturated fatty acids, monounsaturated fatty acids and polyunsaturated fatty acids in

the extracted oil from thraustochytrid species.

66

Page 88: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4

Chapter 4: Optimised cell disruption methods for lipase

extraction from Australian thraustochytrids

67

Page 89: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4 4.1. Introduction

The marine environment has a unique biodiversity and is a good source of

enzymes with unusual properties with potenial utility in biotechnology (Lee et al.,

2010a). Considerable effort has gone into screening marine extracts for novel bioactive

compounds, with unique structural classes being discovered (Kurtovic & Marshall,

2013; Vester et al., 2014). Marine microorganisms are particularly useful for bioactive

compunds discovery since they can be fermented and so offer ease of scaleup versus

other sources (Sana, 2015). Among marine microbes, thraustochytrids are one of the

most scaleable strains since they are commerically useful for the production of omega-

3 oils and carotenoids. These heterotrophic microorganisms also are a potentially

useful source of new enzymes (Gupta et al., 2012a; Gupta et al., 2013).

Thraustochytrids can accumulate approximately 40% docosahexaeonic acid (DHA) as

a percentage of total lipids (Liu et al., 2014b). Thraustochytrids have been used as

animal feed, aquaculture feed and poultry feed in addition to their use in omega-3 and

carotenoid production (Milledge, 2012; Sprague et al., 2015). Thraustochytrids have

been reported to produce degradative enzymes such as lipases, peptidases,

phosphatases and cellulases, although the utility of their enzymes has not been fully

investigated (Bongiorni et al., 2005; Kanchana et al., 2011).

Cell disruption is an important downstream processing step as it impacts

extraction yields and therefore the cost of bioactive compound production (Lee et al.,

2012c). Various methods–physical, chemical and mechanical–are available for

disrupting microbial cells, with mechanical methods generally used for large-scale

processing (Günerken et al., 2015). The selection of a suitable method of cell

disruption for extracting intracellular products depends on cell wall strength,

intracellular location of products, stability and the final use of the recovered products

(Klimek-Ochab et al., 2011). For example, the effect of chemical and mechanical cell

disruption methods was investigated for extraction of intracellular -D-galactosidase

(Puri et al., 2010).

Lipases are hydrolytic enzymes that catalyse the hydrolysis or synthesis of esters

and possess positional as well as fatty acid selectivity (Ferreira-Dias et al., 2013; Singh

& Mukhopadhyay, 2012). Lipases isolated from marine organisms generally showed

maximum activity at low temperatures and have unusual fatty acid specificity (Vester

et al., 2014). Lie and Lambertsen, observed that the crude enzyme mixture collected

68

Page 90: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4 from cod intestine showed an unusual fatty acid selectivity, preferentially hydrolysed

omega-3 fatty acids (Lie & Lambertsen, 1985). Similar fatty acid selectivity has also

been observed from the crude digestive juice from salmon and rainbow trout collected

from the pyloric area of the stomach (Halldorsson et al., 2004). However, in each of

these cases no omega-3 lipase has been characterised. A recent study reported that a

novel lipase isolated from the marine microbe Marinobacter lipolyticus SM19 showed

a high preference for hydrolysing eicosapentaenoic acid (EPA) from fish oil (Pérez et

al., 2011).

We report here the development of a cell disruption method for release of

intracellular lipases from thraustochytrids. The efficiency of different cell disruption

methods for the release of intracellular proteins and their effect on lipolytic activity

was investigated. To the best of our knowledge this is the first report on optimisation

of cell disruption methods for lipase extraction from thraustochytrids.

4.2. Materials and methods

4.2.1. Chemicals

All chemicals such as p-nitrophenyl palmitate (p-NPP), Tween 80, and medium

components (yeast extract and peptone) were procured from Sigma-Aldrich, USA.

Acid washed glass beads from Thomas Scientific, Australia and Sea salts from Instant

Ocean, USA. The API ZYM kit was procured from BioMérieux, Australia Pty. Ltd.

4.2.2. Strain selection and cultivation

Recently isolated strains designated as AMCQS5-3 (Genbank accession

numbers JX993839), AMCQS5-4 (accession numbers JX993840), AMCQS5-5

(accession numbers JX993841), AMCQS1-9 (accession numbers JX993843) and

AMCQ-4b27 (accession numbers JX993842) were used for screening lipase activity

(Gupta et al., 2013). Two microbial cultures, Schizochytrium S31 (S31) and PRA 296,

were procured from the American type of culture collection (ATCC).

Thraustochytrids were maintained in a GYP medium containing glucose (5 g L-

1) , peptone (2 g L-1) and yeast extract (2 g L-1) and artificial sea water (50% v/v) at

20 oC. Stock cultures were subcultured onto fresh plates at regular intervals. All

thraustochytrids were grown in modified Vishniac’s Medium (0.1% glucose, 0.01%

yeast extract, 0.01% peptone, 0.1% gelatine hydrolysate, 1.2% agar in seawater, pH

69

Page 91: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4 7.0) for enzyme screening (Raghukumar, 2002). The incubation conditions were 20 oC

and 150 rpm agitation until the cells reach the late exponential growth phase (96 h).

Before harvesting, cells were observed under a microscope for growth.

4.2.3. Enzymatic profiling of isolates

The thraustochytrid cells were harvested at 5000 rpm for 15 min and washed

with sterile artificial sea water. Cells were suspended in artificial sea water and 65 μl

of sub aliquots (corresponds to 4.0 x 105 cells) inoculated into (19 microcupules except

in control) API ZYM strips and incubated at 20 oC overnight. After incubation, 30 μl

of API ZYM A and B were added to the microcupules and incubated for 5 min.

Enzyme activities were measured using a colorimetric scale (Tiquia, 2002).

4.2.4. Lipase production

Thraustochytrid were grown in medium containing peptone (500 mg L-1), yeast extract

(500 mg L-1) and Tween 80 (1% v/v) (Arafiles et al., 2011). The production medium

was pre-adjusted to pH 6.0 using 0.1 M HCl. A volume of 50 ml medium contained

in a 250 ml Erlenmeyer flask was inoculated with a loop full of culture from the plate.

The flask was incubated at 20 oC with shaking at 150 rpm for 48 h. Inoculum (5% v/v)

was used for the production of biomass. The production culture was incubated for 120

h under the same conditions. The biomass was harvested by centrifugation (5000 rpm,

10 min). Harvested biomass was resuspended in distilled water to remove the Tween

80, and this process was repeated three times (Li et al., 2001). The biomass was kept

at -80 oC overnight and freeze-dried (Christ, Germany) for 24 h for further use.

4.2.5. Cell disruption methods for lipase release

4.2.5.1. Bead vortexing

A sample of freeze-dried biomass (100 mg) was suspended in 1 ml of

extraction buffer (100 mM Tris pH 7.2) with EDTA (10 mM) and NaCl (100 mM).

One ml of acid washed glass beads (425-600 μm, Klausen Pty Ltd, Australia) were

added and the contents were vortexed for 10 min, in 30 second burts. Ground biomass

was centrifuged at 12,000 rpm for 60 min at 4 oC. The supernatant was analysed for

lipase activity and protein levels as described below.

70

Page 92: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4 4.2.5.2. Grinding with liquid nitrogen

A sample of freeze-dried biomass (100 mg) was suspended in 10-15 ml of

liquid nitrogen and ground with a pestle for 2 min. 1 ml of extraction buffer was added

and the resultant biomass was collected and centrifuged at 12,000 rpm for 60 min at

4 oC. The supernatant was analysed for lipase activity.

4.2.5.3. Sonication

A sample of freeze-dried biomass (100 mg) was suspended in 1 ml of

extraction buffer. The suspension was sonicated at 20 kHz, 40% amplitude and the

pulse was 40 seconds on 20 seconds off through a time-range of 5-25 min (Sonics,

USA). After every 5 min, cells were observed under a microscope to check the level

of disruption. During sonication, the sampling tubes were kept in ice-bath to avoid

heat-mediated denaturation of crude enzyme. The suspension was centrifuged at

12,000 rpm for 60 min at 4 oC. The supernatant was analysed for lipase activity.

4.2.5.4. Homogenization

Homogenization was performed using a rotor-stator type homogenizer

(Unidrive 1000, CAT Scientific, USA). A sample of freeze-dried biomass (100 mg)

was suspended in 1 ml of extraction buffer and the suspension was subjected to

homogenization at 8500 rpm/min for 5-25 min at 5 min intervals. After every 5 min,

the suspension was observed under the microscope. The suspension was centrifuged

at 12,000 rpm for 60 min at 4 oC. The supernatant was analysed for protein

concentration and lipase activity.

4.2.6. Microscopic observation of the native and disrupted thraustochytrid cells

The cell slurries were observed under a microscope using differential

interference contrast (Axio-imager, Zeiss, Germany) to check for disruption.

Microbial smears of normal and disrupted cells were prepared on the glass slide and

air-dried

4.2.7. Lipase and protein assay

Lipase activity was measured from the supernatant obtained after the cell

disruption using p-nitrophenol palmitate pNPP as a substrate, following a previously

described method with some modifications (Winkler & Stuckmann, 1979). The lipase

substrate solution was prepared by adding solution A consisting of 30 mg of pNPP in

10 ml of isopropanol and the solution B consisting of 0.4% Triton X-100, 0.1% Gum 71

Page 93: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4 Arabic and 50 mM Tris-HCl buffer, pH 7.0. Solution A (10 ml) was added to solution

B (90 ml) drop by drop with stirring. The substrate solution (2.9 ml) and 0.1 ml of

enzyme solution was incubated at 37 oC for 10 min and absorbance was measured at

410 nm. One unit (IU) of lipase activity was defined as one nmol of pNPP released per

minute under the standard assay conditions.

Protein concentration was estimated from the supernatant using the Bradford

method with bovine serum albumin as the standard (Bradford, 1976). All the enzyme

and protein assays were done in four replicates and the standard error (SE) calculated

using Microsoft excel software.

4.3. Results and discussion

In this study, seven thraustochytrids isolates were screened for 19 different

enzyme activities. An API ZYM kit was used that contained enzyme strips that

consisted of 20 microcupules with 19 different chromogenic substrates and one

control. Table 4.1 shows the wide spectrum of enzyme activities that were exhibited

by the thraustochytrids isolated from Austrlian marine water. The new isolates were

isolated form two different sample collection trips. The investigated thraustochytrid

strains exhibited high esterase, lipase, acid and alkaline phosphatase activities. The

peptide degrading enzyme leucine arylamidase was produced by all strains, except

AMCQS5-4. Maximum lipase activity was observed in AMCQS1-9. Other peptidases

such as cysteine arylamidase and trypsin were produced in all isolates, except

AMCQS5-4 and AMCQS1-9. PRA 296 and AMCQS1- -chymotrypsin

-galactosidase -glucosidase activities were observed in the strains

PRA 296 and AMCQS5-5.

72

Page 94: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4 *Table 4.1a. Enzymatic activities of thraustochytrids isolated form Australian marine

waters (First collection trip, October 2011).

Enzyme S31 PRA GTC 7 6a13 S53 S54 4b27 S19 S55Alkaline Phosphatase

LE HE

Esterase HE

Esterase + Lipase

Lipase

Leucine arylamidaseValine arylamidase

Cysteine arylamidaseTrypsin

a-chymotrypsin

Acid phosphatase

Napthol-AS-BI-phosphohydrolase

a-galactosidase

b-galactosidase

b-glucuronidase

a-glucosidase

b-glucosidase

N-acetyl-b-glucosaminidase

a-mannosidase

a-fucosidase

(Indicate what is shown in red, green and straw yellow colour)

*Samples were collected by Dr A Gupta and Dr T Thyagrajan from Barwon region,

Victoria, Australia.

73

Page 95: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4 **Table 4.1b. Enzymatic activities of thraustochytrids isolated form Australian marine

waters (Second collection trip, September 2013).

Enzymes PP31 AGT-0

AGT-8

TS-1

TS-7

TS-8

TS-4

DSP5-1

DSP 2-4

DSP1-1B

DSP 2-3

AVI-1

AVI-2

Alkaline PhosphataseEsteraseEsterase + LipaseLipaseLeucine arylamidaseValine arylamidaseCysteine arylamidaseTrypsina-chymotrypsinAcid phosphataseNapthol-AS-BI-phosphohydrolase

-galactosidase-galactosidase-glucuronidase-glucosidase-glucosidase

N-acetyl-b-glucosaminidase

-mannosidase-fucosidase

Footnote: Semi-quantitative screening of thraustochytrid for different enzymatic

activities was done using a commercial enzyme kit. Thraustochytrid cells were

inoculated into the strips and observed for colour formation due to the enzymatic

action. (Green) high activity, (yellow) medium activity; (red) low activity; (no colour)

no activity. GenBank accession numbers for isolates given in Appendix.

Cluster analysis revealed that the enzymatic activities of thraustochytrid strains

were homogenous, with 86% similarity (Fig. 4.1). Bongiorni et al. also observed a

range of enzyme activity in some thraustochytrids (Bongiorni et al., 2005).

**Others bioprocessing lab members Dr A Gupta, Dr T Thyagrajan, Dr D Singh and

Dr S Sonkar were involved in the collection of various isolates from the Queenscliff

region, Victoria.

74

Page 96: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4

Figure. 4.1. Cluster analysis of thraustochytrids based on various enzyme activities.

Cluster analysis was done using XLSTAT version and reveals that degree of similarity

was 86% based on presence/absence in their enzymatic profiles.

Lipase production was inhibited by high concentrations of glucose in the

medium (Kanchana et al., 2011). To improve lipase production levels, thraustochytrids

were grown in a Tween 80 containing medium. Intracellular lipase activity was

measured using a colorimetric enzyme assay. The lipase activity from S31 (39 IU/g)

was higher than that observed for the other strains used in the current study. The

maximum lipase activities observed for the thraustochytrid isolates were 34 IU/g for

AMCQ-4b27 and 36 IU/g for AMCQS1-9 (Table 4.1). 18r DNA sequence analysis

identified these strains as a Thraustochytrium sp (AMCQ-4b27) and a Schizochytrium

sp (AMCQS1-9) (Gupta et al., 2013).

75

Page 97: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4 Table 4.1. Lipase activity of thraustochytrids new isolates and ATCC cultures. Lipase

activity was determined by p-NPP assay using whole cell as catalyst.

Thraustochytrid strains Lipase activity (IU/g)

In house isolates AMCQS5-3 26±2.0

AMCQS5-4 21±1.0

AMCQS5-5 32±2.0

AMCQS1-9 36±1.6

AMCQ-4b27 34±4.2

ATCC S31 39±3.7

PRA 296 28±2.0

4.3.1. Cell disruption for lipase extraction

Cell disruption is a key factor in the extraction of intracellular enzymes from

microbial cells. Different methods are available in the literature to solubilize the

intracellular proteins that depend on the location of enzymes, its application and

stability (Kurtovic & Marshall, 2013). The thraustochytrids strains examined had a

range of enzymes activities, with some lipase activity. We assessed the effect of cell

disruption techniques on lipase extraction from two thraustochytrid strains. The impact

of different cell disruption methods on enzyme yield varies with enzyme stability and

cell type, so that a balance between maximum cell degradation effectiveness and

minimum protein denaturing needs to be achieved.

The effect of cell disruption methods on lipase extraction from thraustochytrid

was studied via quantifying protein content released, lipase extraction and also by

direct microscopic observation (Fig. 4.2 a-d). Undisrupted cells of AMCQS1-9 and

AMCQ-4b27 remained intact (Fig. 4.2a and b), whereas, cells subjected to disruption

ruptured to release intracellular components (Fig. 4.2c and d).

76

Page 98: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4

Figure. 4.2. Microscopic observation of thraustochytrid after cell disruption.

Undisrupted/Intact cells of AMCQS1-9 and AMCQ-4b27 (a and b), and disrupted cells

of AMCQS1-9 and AMCQ-4b27 (c and d) by sonication reaction conditions ( 20 KHz,

40% amplitude, 15 min) were observed after the treatment using differential

interference contrast.

4.3.1.1. Bead vortexing

Thraustochytrid cells AMCQ-4b27 and AMCQS1-9 were disrupted by bead

vortexing. The lipase activity and specific activity after cell disruption by bead

vortexing are shown in Fig. 4.3. More protein was obtained from AMCQ-4b27 cells

(130 μg/ml) than from AMCQS1-9 (63 μg/ml), however, lipase activity was greater in

protein extracted from AMCQS1-9 cells (0.045 IU/μg versus 0.125 IU/μg,

respectively). The effectiveness of bead vortexing versus other cell disruption methods

77

Page 99: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4 appears to be cell type dependant. -galactosidase

from Kluyveromyces lactis by disruption with glass beads. They also observed the

volumetric and specific activity were minimum when compared with sonication and

other methods (Becerra et al., 2001).

AMCQ-4b27 AMCQS1-9

6.0

6.5

7.0

7.5

8.0

Acti

vity (

IU/m

l)

Thraustochytrids

Activity Protein yield Specific activity

60

70

80

90

100

110

120

130

140

Prot

ein yi

eld (μ

g/ml)

0.04

0.06

0.08

0.10

0.12

0.14

Spe

cific

activ

ity (I

U/μg

)

Figure. 4.3. Effect of bead vortexing on thraustochytrid cell disruption and lipase

activity. Bead vortexing of thraustochytrid cells resulted in more protein release from

AMCQ-4b27 than AMCQS1-9. S1-9 cells were relatively more susceptible for high

enzyme activity than AMCQ-4b27 cells (in all case SE <10%).

4.3.1.2. Grinding with liquid nitrogen

Grinding with liquid nitrogen is a simple and quick method for disrupting

microbial cells (Zheng et al., 2011). Grinding with liquid nitrogen was faster than other

methods for cell disruption. The total protein extracted by this method was higher than

that obtained by other methods, for AMCQ-4b27 (156 μg/ml) and AMCQS1-9 (107

μg/ml). Fig. 4.4 showed that grinding with liquid nitrogen led to a higher degree of

cell disruption, consistent with higher protein yields. Lipase activity of AMCQ-4b27

(25 IU/ml) was higher than that for AMCQS1-9 (18 IU/ml). However, specific activity

from AMCQ-4b27 (0.160 IU/μg) was lower than that for AMCQS1-9 (0.168 IU/μg).

Lipase from AMCQ-4b27 more readily degraded with temperature than lipase from

AMCQS1-9. Our results are consistent with the study investigated the effect of

different chemical and physical methods for extraction of -D-galactosidase, where

78

Page 100: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4 grinding with river sand is effective in enzyme extraction compared with other

methods (Puri et al., 2010).

AMCQ-4b27 AMCQS1-917

18

19

20

21

22

23

24

25

26

Acti

vity (

IU/m

l)

Activity Protein yield Specific activity

Thraustochytrids

100

110

120

130

140

150

160

Prote

in yie

ld (μ

g/ml)

0.160

0.162

0.164

0.166

0.168

Spec

ific a

ctivit

y (IU

/μg)

Figure. 4.4. Effect of grinding on thraustochytrid cell disruption and lipase activity.

AMCQS-4b27 cell were more susceptible to cell disruption compared with AMCQS1-

9. Grinding with liquid nitrogen resulted in more protein released than did other

methods.

4.3.1.3. Sonication

Sonication was previously shown to be an effective method for releasing

intracellular bioactive compounds (Liu et al., 2013). In order to understand the optimal

sonication time for maximum protein release and lipase activity, we varied the

sonication treatment time. The impact of sonication exposure time on thraustochytrid

cells was investigated. Enzyme released by sonication can be calculated as:

R = Rm (1- e –kt)

Where R and Rm are the released enzyme activity and maximum enzyme

activity that can be released, respectively (U/mg/min), k is the disruption constant for

sonication and t is the time of sonication in minutes (Doulah, 1977). Lipase release

from thraustochytrid cells was investigated at an acoustic power of 20 kHz and 40 %

amplitude for 5 to 25 mins. Figure 4.5 [(a) and (b)] shows higher levels of protein

extraction as sonication time increased for thraustochytrids AMCQ-4b27 and

AMCQS1-9 cells. However, lipase activity increased to 15 minutes but plateaued after

that, so that the lipase concentration in extracted protein decreased after 15 minutes 79

Page 101: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4 extraction (Figure 4.5). The maximum lipase activity obtained from AMCQ-4b27 was

47 U/ml and from AMCQS1-9 was 42 U/ml. This indicates that either heat produced

during sonication may have denatured lipase, or that lipase extracted more readily than

other protein (Farkade et al., 2006). Decreased lipase activity when acoustic power

was increased beyond 40% amplitude is consistent with heat-induced degradation. A

recent study indicated deactivation of Yarrowia lipolytica lipase upon increasing

sonication time (Kapturowska et al., 2012).

Fig. 4.5 (b)

5 10 15 20 2520

25

30

35

40

45

50

Acti

vity (

IU/m

l)

Activity Protein yield Specific activity

Sonication time (min)

48

50

52

54

56

58

60

62

64

66

Pro

tein y

ield (μg

/ml)

0.5

0.6

0.7

0.8

0.9

Spe

cific

activ

ity (I

U/μg

)Fig. 4.5 (b)

5 10 15 20 250

5

10

15

20

25

30

35

40

45

Acti

vity (

IU/m

l)

Activity Protein yield Specific activity

Sonication time (min)

20

30

40

50

60

Prot

ein yi

eld (μ

g/ml)

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

Spe

cific

activ

ity (I

U/μg

)

80

Page 102: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4 Figure. 4.5. Effect of sonication on thraustochytrid cell disruption and lipase activity.

Time course effect of sonication on the protein release and lipase activity from

AMCQ-4b27 and AMCQS1-9.

4.3.1.4. Homogenization

During homogenization, microbial cells enter into a rotor-stator creating shear

force that disrupts the cells (Lee et al., 2012c). After 25 min homogenization 133 μg/ml

protein was extracted from AMCQ-4b27 cells and 75 μg/ml protein from AMCQS1-9

cells. As expected, protein content increased with increasing number of cycles (Halim

et al., 2012b). Lipase activity increased substantially in AMCQ-4b27 (25 IU/ml) and

AMCQS1-9 (37 IU/ml) until 20 min of homogenization and started declining

thereafter [Fig. 4.6 (a) and (b)], likely due to heat deactivation of the lipases. The

decrease in lipase activity after 20 min of homogenization can be attributed to enzyme

inactivation from the heat produced during the process. Overall lipase activity was

relatively low when compared to sonication, indicating that the combination of shear

forces and surface tension during disruption process increases lipase degradation (Lee

Ying Yeng et al., 2013; Taubert et al., 2000).

Fig. 4.6 (a)

5 10 15 20 25

10

12

14

16

18

20

22

24

26

Acti

vity (

IU/m

l)

Activity Protein yield Specific activity

Homogenization time (min)

40

60

80

100

120

140

Prote

in yie

ld (μ

g/ml)

0.16

0.17

0.18

0.19

0.20

0.21

0.22

Spec

ific a

ctivit

y (IU

/μg)

Fig. 4.6 (b)

81

Page 103: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4

5 10 15 20 2510

15

20

25

30

35

40

Acti

vity (

IU/m

l)

Activity Protein yield Specific activity

Homogenization time (min)

30

40

50

60

70

80

Prote

in yie

ld (μ

g/ml)

0.46

0.48

0.50

0.52

0.54

0.56

Spec

ific a

ctivit

y (IU

/μg)

Figure. 4.6. Effect of homogenisation on thraustochytrid cell disruption and lipase

activity. Time course of extraction and its effect on lipase activity by homogenization.

4.3.2. Comparison of different cell disruption methods for lipase extraction

To release intracellular microbial products, such as lipases, efficient cell disruption

methods are required. However, these should not denaturation the compounds of

interest. In this study the four different cell disruption methods, bead vortexing,

grinding with liquid nitrogen, sonication and homogenization, were compared for

impact on protein and lipase recovery from thraustochytrid extraction (Table 4.2).

Sonication resulted in maximum lipase yield from AMCQ-4b27 (0.903 IU/μg) and

AMCQS1-9 (1.44 IU/μg). Lower specific activities were obtained using bead

vortexing (0.045 IU/μg and 0.125 IU/μg), grinding with liquid nitrogen (AMCQ-4b27

0.160 IU/μg and AMCQS1-9 0.168 IU/μg) and homogenization (AMCQ-4b27 0.211

IU/μg and AMCQS1-9 0.556 IU/μg). Extracts from strain AMCQS1-9 showed

approximately 1.5 times higher lipase activity than those from AMCQ-4b27. Overall,

sonication resulted in the highest lipase activity, followed by homogenization,

grinding with liquid nitrogen and bead vortexing.

82

Page 104: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 4 Table 4.2. Disruption of thraustochytrid cells for lipase extraction.

Cell disruption

method

AMCQ-4b27 AMCQS1-9

Lipase

activity

(IU/ml)

Protein

released

(μg/ml)

Lipase

activity

(IU/ml)

Protein

released

(μg/ml)

Bead vortexing 6±0.7 130±2.1 8±1.0 63±2.0

Grinding with liquid

nitrogen

25±3.0 156±2.4 18±0.5 107±1.6

Sonication 47±0.2 52±3.0 42±3.7 29±1.3

Homogenization 25±5.0 126±17 37±4.0 73±9.0

4.4. Conclusion

Thraustochytrids contain a range of potentially useful enzyme activities,

including lipase activity. Cell disruption methods were used to increase the extraction

yield of intracellular lipase from thraustochytrid cells. We conclude that AMCQ-4b27

cell were sensitive to protein release than the AMCQS1-9. Moreover, AMCQS1-9 has

yielded high total specific lipase activity than the AMCQ-4b27. Sonication resulted in

the higher lipase activity yield, with strain AMCQS1-9 yielding more lipase activity

than did strain AMCQ-4b27. Lipase activity yield did not completely correlate with

protein extraction yield, probably due to partial lipase degradation under conditions of

high temperature or sheer.

83

Page 105: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 5

Chapter 5: Tween 80 influences the production of

intracellular lipase by Schizochytrium S31 in a stirred tank

reactor

84

Page 106: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 5 5.1. Introduction

Lipases (triacylglycerol lipases, EC 3.1.1.3) belongs to the hydrolase family of

enzymes that catalyse the hydrolysis of triglycerides into free fatty acids and glycerol

(Schreck & Grunden, 2014). However, under certain conditions, they are also able to

catalyse synthetic reactions (Martins et al., 2014). In addition, lipases have some

important functional properties such as substrate selectivity, regio-selectivity and high

enantio-selectivity (Gupta et al., 2007; Yuan et al., 2014). These properties make them

potential catalyst in numerous industrial processes, such as food, detergent, chemical,

and pharmaceutical industries (Sangeetha et al., 2011; Yuan et al., 2014). Lipases are

ubiquitous enzymes widely distributed in plants, animals and microorganisms

(Mazzucotelli et al., 2015). Among them, microbial lipases have displayed a broad

range of substrate specificities. This property might have evolved through the access

of microbial lipases to different carbon sources (Hasan et al., 2006; Salihu & Alam,

2015). There is a growing demand for the novel lipases from a microbial sources,

because of their ease in production and access to genetic manipulations. Also because

they work under mild reaction conditions without requirement of any co-factors (Pérez

et al., 2011).

The health benefits of omega-3 fatty acids have been evaluated by a number of

clinical studies (Barrow, 2010; Nichols et al., 2014). There is strong evidence for the

efficacy of omega-3 fatty acids in the prevention of sudden death from cardiovascular

disease and rheumatologic conditions (Abuajah et al., 2015; Bhangle & Kolasinski,

2011; De Caterina, 2011). The anti-cancer activity of omega-3 fatty acids, particularly

against colorectal cancer, has recently been shown (Cockbain et al., 2012). Infant brain

and eye development is influenced by DHA, which has led to the addition of DHA to

the infant formulae. Learning capacity of school going children has been shown to

increase by taking DHA enriched food supplements (Qawasmi et al., 2013; Wu et al.,

2015). Oily marine fish are the primary source for the industrial production of omega-

3 fatty acids. These fish derived omega-3 fatty acids from their diet, primarily from

phytoplankton. Researchers have implemented different methods for concentrating

omega-3 fatty acids (Rubio-Rodríguez et al., 2010). A promising green approach for

concentration of omega-3 fatty acids is the use of lipases. Generally the enzymatic

methods are faster than other methods and can be carried out under milder conditions

then methods involving chromatographic separation, molecular distillation or urea

85

Page 107: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 5 complexation (Kahveci et al., 2010; Kralovec et al., 2010). Commercially available

lipases have been exploited for omega-3 fatty acid concentration (Akanbi et al., 2013;

Casas-Godoy et al., 2014; Fernández-Lorente et al., 2011; Sharma et al., 2013;

Valverde et al., 2013). However, selectivity of fatty acid hydrolysis is poor and no

know lipases selectivity hydrolyse DHA and EPA, a property that would be useful in

selectivity concentrating these fatty acids from fish or microbial oils.

Thraustochytrids produce lipases and because of their high productivity of

omega-3 fatty acids, particularly DHA, they are potential sources of omega-3 selective

lipases, although none have yet been characterised (Bongiorni et al., 2005; Kanchana

et al., 2011). We recently reported lipase activity from a strain of Schizochytrium

named S31 (Table 4.2). In this work we optimised culture conditions to enhance lipase

production from Schizochytrium sp. S31 in a bioreactor and optimised the extraction

of this lipase using sonication. Using a recently developed p-nitrophenyl esters (p-NP)

assay (Nalder et al., 2014) we showed that lipase extracted from S31 preferentially

hydrolysed DHA over EPA, an unusual selectivity for lipases.

5.2. Materials and methods

5.2.1. Microorganism and maintenance of culture

Schizochytrium S31 (ATCC 20888) was procured from the American type

culture collection (ATCC). The stock culture was maintained on GYP medium

consisting of glucose 5, yeast extract 2, mycological peptone 2, agar 10 g L-1 and

artificial seawater 50% at 25 °C and sub-cultured after 15 days. All medium

components and chemicals including p-NP were obtained from Sigma, Australia. The

p-NP fatty acid esters of -linolenic acid (C18:3),

eicosapentaenoic acid (C20:5) and docosahexaenoic acid (C22:6) were synthesised

based on a recently published protocol (Nalder et al., 2014).

5.2.2. Schizochytrium S31 growth and lipase production

Schizochytrium S31 was grown in a medium containing peptone 0.5, yeast

extract 0.5 g L-1 and Tween 80 (1% w/v; as sole carbon source). 50 ml of medium in a

250 ml Erlenmeyer flask was inoculated with a loop full of culture from the plates.

This culture flask was incubated for 48 h at 20 oC and shaken at 150 rpm. A 48 h old

5% seed culture (v/v) was used to inoculate sterilised production medium (100 ml in

86

Page 108: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 5 500 ml Erlenmeyer flask) containing peptone 0.5, yeast extract 0.5 g L-1 and Tween 80

for the production of lipase. The flasks were incubated in a rotary shaker for up to 9

days. Samples were withdrawn aseptically at regular time intervals and then dried and

weighted and lipase activity determined. For dry weight determination, the entire

content of the liquid culture was transferred to a pre-weighed centrifuge tube and

harvested by centrifugation at 3500 rpm for 10 min and the supernatant was discarded.

Harvested cells were washed thoroughly with phosphate buffer (pH 7.0) and then

freeze-dried for 24 h. The biomass weight was determined gravimetrically. Tween 80

concentrations were estimated based on a colorimetric method (Cheng et al., 2011b).

A standard curve was prepared by varying the Tween 80 concentration and the

unknown sample concentration was determined using the standard curve. Temperature

(15 to 30 oC) and initial pH (5-8) of the medium were optimised to maximise lipase

activity. All experiments were carried out in triplicates.

Fermentation of thraustochytrids was carried out in a 2 L stirred tank reactor

(New Brunswick, USA) with a working volume of 1.4 L at 25 oC with aeration of 0.5

vvm and agitation of 150 rpm. Varying concentration of Tween 80 (0.5 to 1.5%, v/v)

was used to optimise biomass, lipase and lipid production in the bioreactor. Samples

were withdrawn at regular time intervals and analysed for residual substrate. Changes

in pH were also monitored during the course of fermentation.

5.2.3. Estimation of Schizochytrium S31 growth

The growth of the organism was estimated by taking 1-ml aliquot periodically

and measuring the optical density (OD) at 600 nm in a UV-visible spectrophotometer

(Shimadzu1601, Japan) with medium as a blank. The OD value was converted into

cell mass concentration (mg ) using a standard curve.

5.2.4. Optimisation of cell disruption through sonication

Cell disruption of Schizochytrium S31 biomass was performed as previously

described (Byreddy et al., 2015). Thraustochytrid cell biomass (50 mg) was suspended

in 3 ml of solvent in a 15 ml centrifuge tube. Sample was sonicated at 20 kHz, 40%

amplitude and the pulse was 40 sec on and 20 sec off with total working time of 20

min (Sonics, Newtown, CT, USA). Sample tubes were kept on ice during the

sonication process to prevent overheating. Sonication conditions such as time and

87

Page 109: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 5 acoustic power were optimized. The sonicator used in this study was a horn type with

500 W power output and 20 KHz.

5.2.5. Lipase assay and protein estimation

Lipase activity was measured in the supernatant obtained after cell disruption

using p-nitrophenol palmitate (pNPP) as a substrate, following a previously described

method with some modifications (Winkler & Stuckmann, 1979). The lipase substrate

solution was prepared by adding solution A consisting of 30 mg of pNPP in 10 ml of

isopropanol and the solution B consisting of 0.4% Triton X-100, 0.1% Gum Arabic

and 50 mM Tris-HCl buffer, pH 7.0. Solution A (10 ml) was added to solution B (90

ml) drop by drop with stirring. The substrate solution (2.9 ml) and 0.1 ml of enzyme

solution was incubated at 37 oC for 10 min and absorbance was measured at 410 nm.

One unit (IU) of lipase activity was defined as one nmol of p-nitrophenol released per

minute under the standard assay conditions. The relative activity of crude enzyme is

calculated as a percentage activity divided by the most hydrolysed substrate

(designated 100% for each substrate).

Protein concentration was estimated from the supernatant using the Bradford

method with bovine serum albumin as the standard (Bradford, 1976). All enzyme and

protein assays were done in four replicates and the standard error (SE) calculated using

Microsoft excel software.

5.2.6. Lipid extraction

Lipid was extracted and quantified from freeze-dried biomass every 48 h

according to a previously published protocol (Lewis et al., 2000) with some

modifications. Freeze-

and methanol in 2:1 ratio) and the sample vortexed for 2 min, followed by

centrifugation at 13,000 rpm for 15 min. This extraction process was repeated three

times. The supernatants were collected in a pre-weighed vials and dried in an oven at

50 °C. The lipid percentage was measured gravimetrically. All the experiments were

performed in triplicate.

5.3. Results and discussion

5.3.1. Optimisation of culture conditions for lipase production

88

Page 110: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 5

Lipase production is influenced by the type of carbon source used, temperature

and initial pH of the culture medium (Orlando Beys Silva et al., 2005). Tween 80 has

been successfully used as a carbon source for the production of lipases from various

microbes (BoekeMa et al., 2007; Li et al., 2001; Sharma et al., 2014). In this study,

Tween 80 was used as the sole carbon source and its concentration was optimised to

maximise biomass and lipase productivity (Fig. 5.1). Tween 80 (1%) resulted in the

highest level of biomass (0.9 g L-1) and lipase (39 IU/g) productivity by

Schizochytrium S31. Similar results were found previously for Acinetobacter

radioresistens (Li et al., 2001). Lipase production did not improve when higher tween

80 concentration (1.5%) was used in the medium, although the reason for this is

unclear. Remaining experiments were performed with 1% tween 80.

Figure. 5.1. Effect of Tween 80 concentration on lipase production by Schizochytrium

S31 at shake flask, pH 6.0 and incubation temperature 25 oC.

Initial pH and temperature are known to influence lipase production (Priji et

al., 2015). In Schizochytrium S31 maximal lipase activity was observed at pH 6 (38

U/g) for the pH range of 5 to 8. Similar results were reported for lipase production

from Thraustochytrium sp. AH-2 and TZ when the medium pH was pre-adjusted to 6

(Kanchana et al., 2011). The effect of incubation temperature on lipase production was

investigated for the four temperatures 15 oC, 20 oC, 25 oC and 30 oC at pH 6 and

agitation 150 rpm, with lipase levels being highest at 25 oC (data not shown). Our

00.10.20.30.40.50.60.70.80.91

05

1015202530354045

0 0.5 1 1.5 2

Bio

mas

s (g/

L)

Lip

ase

activ

ity (I

U/g

)

Tween 80 (%)

Biomass (g/L) lipase (IU/g)

89

Page 111: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 5 results are comparable with lipase production from Colletotrichum gloeosporioides

where maximum lipase activity at 25 oC (Balaji & Ebenezer, 2008).

5.3.2. Production of lipase in stirred tank reactor

Growth kinetics of Schizochytrium S31 was studied in terms of biomass

production, lipid accumulation in cells, and lipase production. After 48 h

acclimatisation in the medium, dry cell weight went from 0.16 g L-1 to 0.2 g L-1, with

the log phase beginning after this time, up to 120 h when growth slows and the

stationary phase begins (Fig. 5.2). Schizochytrium S31 cells grew well in a

fermentation medium containing Tween 80 (1%) as sole carbon source. Fig. 5.2 shows

the time course of the investigated parameters at different growth phases and also

showed that lipase and lipid production were growth associated. Biomass increased

with incubation time and was maximal (0.9 g) after 5 days of incubation. Lipase

production reached a maximum of 39 U/g in the late exponential phase, then the lipase

activity declined (10.7 U/g) up until 9 days of incubation, perhaps due to product

inhibition. Our results are in agreement with other researchers who observed the

maximum lipase activity of Yarrowia lipolytica KKP 379 in late exponential phase

(after 72 h) (Fabiszewska et al., 2014). Another study conducted by Kanchana et al,

(2011), found that the maximum lipase activity was observed in thraustochytrids

between 4-7 days of incubation (Kanchana et al., 2011). The pH and substrate

concentration of the medium decreased as the incubation time increasing, which may

be due to the hydrolytic activity of lipase on Tween 80 releasing oleic acid for

utilisation. In the bioreactor, lipase production increased with increasing biomass

concentration as a result of increasing Tween 80 concentration from 0.5 to 1 %. In

addition, improved control of environmental factors in the vessel resulted in increased

enzyme production, compared with that achieved in flasks.

90

Page 112: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 5

0 2 4 6 8 100.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

0 2 4 6 8 103.5

4.0

4.5

5.0

5.5

6.0

6.5

7.0

0 2 4 6 8 10

0

50

100

150

200

250

300

350

Biom

ass (

g/L)

Biomass

pH

Incubation time (day)

pH

0

10

20

30

40

50

60

70

80

90

100

Sub

strat

e de

pletio

n (%

)

Substrate depletion

0

10

20

30

40

Lipas

e Ac

tivity

(IU/

g)

Lipase

Lipid

(mg/

g)

Lipid

Figure. 5.2. Fermentation profile for the production of lipase by Schizochytrium S31

in a stirred tank reactor (Tween 80 containing medium was used as sole carbon source,

pH 6.0 and temperature 25 oC).

Figure. 5.3. Biomass, lipid and DHA production reached a maximum after five days

of incubation. After that the DHA percentage decreased more rapidly than total lipids.

To investigate lipid content and fatty acid composition, biomass was collected

and analysed at different time intervals up to 9 days. DHA content increased with total

0

20

40

60

80

100

120

0

5

10

15

20

25

30

35

1 2 3 4 5 6 7 8 9

Bio

mas

s (%

)

Lipi

d/D

HA

(%)

Incubation time(daya)

Lipid (%) DHA (%) Biomass (%)

91

Page 113: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 5 lipid content and reached a maximum of 5.4% on day 5. Biomass, lipid and DHA all

decrease after day 5, although DHA decreases more rapidly than does total lipid (Fig.

5.3). Lipid is used for energy and membrane synthesis during starvation (Thompson

Jr, 1996) and DHA is associated with many cellular functions. DHA appears to be used

preferentially in the late stages, which also indicates that lipase is acting on DHA

during this stage (Fig. 5.2 and 5.3). The fatty acid composition of oil from

Schizochytrium S31 grown in Tween 80 medium is given in Fig. 4, with growth in

glucose medium as a control. The oleate in Tween 80 was directly accumulated into

cells (Fig. 5.4). A similar pattern was observed in Thraustochytrium aureum ATCC

34304, where Tween 80 in the medium increased the oleic acid content in lipid (Taoka

et al., 2011). A corresponding decrease in the percentages of other fatty acids was

observed in Tween 80 medium (Fig. 5.4). There was also changes in the ratio of fatty

acids when Tween 80 medium was used, with relative decreases in C16 (26%), C18

(70%) relative to DHA.

Figure. 5.4. Fatty acid composition of Schizochytrium S31 was compared between

lipid extracted from biomass grown GYP medium and that from grown in Tween 80

medium.

5.3.3. Disruption of Schizochytrium S31 using sonication

0

10

20

30

40

50

60

70

80

90

C14:0Myristic

acid

C16:0Palmitic

acid

C16:1n7Palmitoleic

acid

C18:0stearic acid

C18:1n9cOleic acid

C20:5n3EPA

C22:5n6DPA

C22:6n3DHA

Tot

al fa

tty

acid

s (%

)

Fatty acids

Control GYP TW 80

92

Page 114: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 5

In order to investigate the fatty acid selectivity of lipase from Schizochytrium

S31, sonication was used to release the intracellular lipase. The effect of sonication on

Schizochytrium S31 was evaluated by microscopic observation and cells were shown

to be fully broken at high acoustic power. Table 5.1 shows lipase activity extracted

from thraustochytrid cells at an acoustic power of 20 kHz and 40 % amplitude for

intervals of 5 to 25 mins. Lipase activity increased to 15 minutes but plateaued after

that, indicating partial lipase inactivation with increased time over 15 minutes. Since

sonication releases intracellular enzymes but also degrades, optimising time and power

are important for maximising lipase extraction (Kadkhodaee & Povey, 2008; Mawson

et al., 2011; Nguyen & Le, 2013).

Table 5.1. Effect of sonication on protein release and lipase activity from

Schizochytrium S31.

Cell disruption

methods

Cell disruption

parameters

Lipase

activity

(IU/ml)

Protein

concentratio

n (μg/ml)

Specific

activity

(IU/μg)

Sonication 5 min

10 min

15 min

20 min

22.8±1.4

32.5±1.6

37.8±1.5

26.3±1.3

43±1.5

60±1.8

82±2.1

105±9

0.512

0.533

0.451

0.247

Fig. 5.5 (a) and (b) show the effect of sonication on lipase activity from

Schizochytrium S31 at different levels of sonication time and power. As the sonicator

power increased from 30% to 70%, the total proteins released increased from 53 to

122 μg/ml and the most lipase activity was observed at 50% power (64 IU/ml). Total

proteins released after sonication treatment was 3 folds higher at 50 % (0.75 IU/μg)

sonication power than 70% (0.25 IU/μg) (Fig. 5a). Our results are in agreement with

previous studies showing that the amount of protein release from microbial cells

increased linearly with increasing sonication time and power (Apar & Özbek, 2008;

Iida et al., 2008). The highest protein extraction levels from Schizochytrium was

observed after 15 min of sonication time at 70% of sonication power. However, both

total lipase activity and specific lipase activity decreased at 70% power, with extended

time causing additional lipase activity degradation. A recent study by Kapturowska et

93

Page 115: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 5 al, 2013, demonstrated the negative effects of extended sonication time on lipase

activity from Yarrowia lipolytica KKP 379 (Kapturowska et al., 2012).

00.10.20.30.40.50.60.70.8

0

20

40

60

80

100

120

140

30 50 70

Spec

ific

activ

ity

Lip

ae a

ctiv

ity/p

rote

in c

onc

Sonication power (%)

Lipase activity (IU/ml) Protein conc (μg/ml) Specific activity (IU/μg)

00.20.40.60.811.21.41.61.82

0

20

40

60

80

100

120

140

3 6 9

Spec

ific

activ

ity

Lip

ase

activ

ity/P

rote

in C

onc

Sonication Time (min)

Lipase activity (IU/ml) Protein conc (mg/ml) Specific activity (IU/mg)

a

b

94

Page 116: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 5 Figure. 5.5. Cell disruption optimisation of Schizochytrium S31 a) as a function of

sonication power, b) time duration. (Sonication power 70% for 10 mins was subjected

to 100 mg biomass for the lipase extraction).

5.3.4. Hydrolysis of p-Nitrophenyl esters

Understanding and controlling lipase fatty acid selectivity is important in the

modification of fats and oils for high value products and in the production of structured

lipids (Qin et al., 2014). The substrate selectivity of S31 crude lipase activity was

determined using the standard assay in the presence of 8 mM of the different p-NP

esters of various chain lengths: p-NP acetate (C2); p-NP butyrate (C4); p-NP caprylate

(C8); p-NP decanoate (C10), p-NP laurate (C12), p-NP myristate (C14), p-NP

palmitate (C16), p-NP stearate (C18), p-NP-EPA and p-NP-DHA. Lipase extracted

from Schizochytrium S31 showed a broad range of substrate activity (Fig. 6).

Maximum activity was observed towards short chain length fatty acids, followed by

medium chain length fatty acids. An unusual selectivity was observed towards p-NP-

DHA, when compared with p-NP-EPA. This type of unusual fatty acid selectivity was

previous observed for a fish digestive lipase, although this lipase was never

characterised (Halldorsson et al., 2004). However, a study investigated the fatty acid

selectivity of five commercial lipases using p-NP esters found that all lipase showed

preference to EPA over DHA (Nalder et al., 2014). This unusual selectivity might be

due to the requirement of the Schizochytrium to process DHA more than EPA, since

EPA is present in only trace amounts in the oil from this strain, whereas DHA levels

are relatively high, in contrast to most microalgae.

95

Page 117: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 5

Figure. 5.6. Substrate specificity of lipase from Schizochytrium S31 (p-nitrophenyl

esters of different chain lengths were incubated with crude enzyme).

5.4. Conclusion

Schizochytrium S31 was grown in a bioreactor using Tween 80 as the carbon

source, which resulted in high biomass and oleic acid levels. Lipase extraction by

sonication was optimised and the fatty acid selectivity of the extracted lipase

investigated using p-NP fatty acid esters. The lipase extract preferentially hydrolysed

shorter chain fatty acids, however, DHA was preferentially hydrolysed over EPA. This

preferential hydrolysis of DHA is unusual and may be industrially useful. This

selectivity may be a result of this Schizochytrium species having much higher levels

of DHA then EPA, and so DHA processing is more important for the organism growth

and function.

0

20

40

60

80

100

120

Rel

ativ

e ac

tivity

(%)

p-NP acyl esters

96

Page 118: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6

Chapter 6: Bead milling for lipid recovery from

thraustochytrids cells

97

Page 119: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6 6.1. Introduction

Marine fish and microalgal oils are rich sources of omega-3 fatty acids. The

health benefits of omega-3 oil supplements have been documented in numerous studies

(Ellulu et al., 2015). Methods available for the concentration of omega-3 fatty acids

from fish and microbial oils include urea complexation, chromatography, distillation,

low temperature crystallisation, supercritical fluid extraction and enzymatic methods

(Rubio-Rodríguez et al., 2010). The primary advantages of enzymatic methods are

operation under mild reaction condition and less or no use of chemicals which can

degrade the oxidatively sensitive structures.

Omega-3 fatty acid production from microalgae is potentially useful due to

their ability to accumulate up to 30-70% of their dry weight as lipid (Adarme-Vega et

al., 2014; Martins et al., 2013). Microalgae can be used to produce bulk ingredients for

applications in pharmaceuticals, cosmetics, nutraceuticals, animal feeds and functional

foods (Yen et al., 2013). Thraustochytrids are marine heterotrophic protist commonly

found in marine environments, and are classified under Labyrinthulomycetes (Gupta

et al., 2012). They have rapid growth rates and are capable of producing significant

amounts of PUFA, including DHA (Gupta et al., 2013). Thraustochytrids have recently

been identified as a potential microalgae for omega-3 fatty acid and carotenoids

production (Singh et al., 2015a). A range of studies have also demonstrated the

suitability of thraustochytrids for biodiesel production (Chen et al., 2015c; Gupta et

al., 2015; Gupta et al., 2013).

Specific compounds produced from microbial sources, such as lipids, proteins,

carotenoids and carbohydrates, must be extracted and purified using downstream

processes prior to their respective applications (Wijffels et al., 2010). Production of

intracellular metabolites from microalgae involves extensive processing steps

including microalgae cultivation, harvesting, extraction and product recovery. The

selection of suitable technologies depends on the characteristics of the algae and the

stability of the product (Guldhe et al., 2014; Pahl et al., 2013). Research is ongoing

into the development of improved downstream processing techniques for microalgae

lipid recovery, a step which is a substantial contributor to total costs for algal

production systems (Delrue et al., 2012). Following harvesting, cell disruption is an

important step for the extraction of microalgae produced compounds. Different cell

disruption methods have been applied to a range of microalgae, including

98

Page 120: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6 ultrasonication, enzymatic treatment, high-pressure homogenisation, alkali treatment,

microwaves and pulsed electric field (Grimi et al., 2014; Safi et al., 2014a), however,

some of these methods have resulted in low extraction yields of the products of

interest.

Bead mill, a homogeniser used for the grinding of ceramics and paints, has

been used for the mild disruption of microorganisms for the release of proteins

(Schwenzfeier et al., 2011). Recent advancements in temperature control in microalgae

cell suspensions make bead milling a promising technique for microalgae component

recovery (Doucha & Lívanský, 2008; Günerken et al., 2015). The operating principle

of bead milling is based on rapid grinding of thick microbial cell suspensions in the

presence of beads. The operating mode can be either batch (recirculation) or

continuous (single passage through milling chamber) depending on the feedstock used.

Parameters which influence cell disruption efficiency include feedstock residence time

(suspension feed rate in continuous operation, or process time in batch operation)

agitator speed, agitator design, milling chamber design, biomass concentration, bead

diameter, bead density and bead filling (Doucha & Lívanský, 2008). It is hypothesized

that the suspended cells are disrupted in the bead collision zones by compaction or

shear forces with energy transfer from the beads to cells (Gunerken et al., 2015). Bead

milling has been successfully applied in Saccharomyces sp and Chlorella sp cell

disruption for bioactive recovery (Balasundaram & Pandit, 2001; Balasundaram et al.,

2012).

In this study, we optimised thraustochytrid cell disruption using bead mill for

maximising lipid extraction yields. The cells of Schizochytrium were spherical and

filled with numerous lipid bodies. The process was optimised for lipid yield, energy

consumption and processing time. The extracted oil was characterised and further

employed for enzymatic hydrolysis for the concentration of omega-3 fatty acids,

particularly DHA.

6.2. Materials and Methods

6.2.1. Chemicals

All chemicals used in this study were of analytical grade. Medium components

such as glucose, yeast extract, mycological peptone (Sigma-Aldrich, USA) and sea salt

(Instant Ocean, USA) were used for biomass production. Schizochytrium S31 (ATCC

20888) was procured from the American Type Culture Collection (ATCC) and used 99

Page 121: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6 as a standard culture. Solvents such as chloroform, methanol, hexane (from Merck),

acetic acid and diethyl ether (from Fischer and Honeywell, Australia) were used in the

lipid class analysis and gas chromatography. TLC standards (from Nu-Chek Prep,

USA) were used to identify each lipid class.

6.2.2. Thraustochytrids biomass cultivation

Thraustochytrids used in this study were maintained on GYP medium

consisting of glucose 5, yeast extract 2, mycological peptone 2, agar 10 g L 1 and

artificial seawater 50%, at 25 °C and sub-cultured after 15 days. Two strains,

Schizochytrium sp. S31 (ATCC) and Schizochytrium sp. DT3 (an in-house isolate),

grown in glycerol media, showed high biomass and lipid productivity (Gupta et al.,

2013). Seed medium (1% glucose, 0.2% yeast extract, 0.2% mycological peptone,

ASW 50%, pH 6.5) and production medium (4% Glycerol, 1% yeast extract, 0.1%

mycological peptone, ASW 50%, pH 6.5) were autoclaved at 121 °C for 20 min. Seed

medium (50 ml) was inoculated from agar plates and grown for 2 days in a shake flask

at 25 °C, at 150 rpm. Inoculum (5%) was used to inoculate production medium and

incubated under the same conditions for 5 days. Biomass was harvested by

centrifugation (10,000 × g, 15 min) and washed twice with distilled water, then freeze-

dried for 24-48 h. Freeze-dried biomass was stored at -20 °C for further use.

6.2.3. Cell disruption by bead milling

The Dyno-Mill (bead mill; Model No. 130857) from Klausen Pvt. Ltd.,

Germany was used for cell disruption. This mill is well suited to applications in

research and development. The Dyno-mill grinding chamber is positioned horizontally

with a chamber volume 79.6 ml. Algal suspension is fed through a funnel mounted on

the grinding chamber. The beads are accelerated by a motor connected via a central

shaft, causing the grinding effect. The bead mill contained an integrated cooling

system to prevent overheating. A cooling jacket and cooling coil are integrated into

the grinding chamber and supplied by an external water bath to extract heat, and

maintain a constant temperature below 20 oC throughout processing. The cell

suspension residence time in the chamber is controlled via the pump speed. The beads

and product are separated at the end of the chamber by a sieve. Beads used in this study

were (0.4-0.6 and 0.8-1.0 mm) zirconium oxide, which are reported to exhibit

maximum efficiency in cell disruption (Lee et al., 2012a). As per manufacturer advice

100

Page 122: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6 the beads were filled to 54% v/v in the chamber. Keeping these parameter stable, the

effect of biomass concentration, bead size, RPM and processing time were

investigated. In total, four different cell concentrations in distilled water were used;

1%, 3%, 5%, and 7%. Samples were collected at 2 min intervals up to 8 mins and

analysed for total lipid yield and fatty acid profiles.

6.2.4. Microscopic observation

In addition to using lipid yield as a parameter for validating the effectiveness

of bead mill treatment on thraustochytrid cells, the direct effect on the cells was

observed under a differential interference contrast microscope (Axio-imager, Zeiss,

Germany). Microbial smears were prepared on glass slides, and air dried prior to

observation.

6.2.5. Lipid extraction

Fatty acid extraction was performed in accordance with Gupta and co-workers

(Gupta et al., 2012c) with some modifications. 10 mg of freeze dried biomass was

blended with 0.6 ml of solvent. Biomass and solvent were vortexed and then

centrifuged at 4000 rpm for 15 min. Supernatant was place in pre-weighed glass vials

and solvent was evaporated under nitrogen. Lipid content (% dry weight basis) was

determined gravimetrically.

6.2.6. Lipase catalysed hydrolysis of thraustochytrid oil

Lipase-catalysed hydrolysis of oil was carried out using a modification

literature method (Kahveci & Xu, 2011). 500 mg of thraustochytrid oil, a known

amount of lipase (3000 U, EC 232-619-9, Candida rugosa; Sigma-Aldrich, St. Louis,

US) and 1.8 ml of Tris-HCl buffer (50 mM, pH 7.0) were taken in 25 ml round bottom

flasks and heated to 45 oC at 300 rpm stirring. Substrates were stirred for 30 min and

lipase was added to initiate hydrolysis. Samples form the reaction mixture were

withdrawn at specific time intervals over the course of the 1-4 h reaction. The enzyme

hydrolysis reaction was scaled up with 2 g of thraustochytrid oil with optimised

enzyme concentration. All experiments were conducted in triplicates with the mean

values reported.

101

Page 123: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6

6.2.7. Thin Layer Chromatography for lipid class analysis

Both the hydrolysed and unhydrolysed portions of thraustochytrid oil were

analysed by capillary chromatography with a flame ionization detector (Iatroscan

MK5, Iatron Laboratories Inc., Tokyo, Japan). The Iatroscan settings were: air flow

rate; 200 ml/min, hydrogen flow rate; 160 ml/min and scan speed of 30 s/scan. Under

these conditions, the chromarods were cleaned by scanning twice before applying

samples. 1 μL of each lipid fraction in hexane was spotted onto the rods with the aid

of an auto pipette along the line of origin on the rod holder and developed for 22 min

in a solvent tank containing hexane/diethyl ether/acetic acid (60:17:0.2, vol/vol/vol).

TLC standards were used to identify each lipid class.

6.2.8. Gas Chromatography analysis for fatty acid composition

Hydrolysed and unhydrolysed thraustochytrid oil compositions were analysed

by gas chromatography (Agilent 6890) with a flame ionisation detector (FID),

equipped with a Supelcowax 10 capillary column (30 m, 0.25 mm i.d., 0.25

thickness; Supelco). Fatty acids were converted to methyl esters by acid-catalysed

trans-esterification with minor modifications (Christie & Han, 2010). 1 ml toluene was

nonadecanoate (C19

methanol (2 ml) was also added to the tube and kept for overnight incubation at 50 °C.

Fatty acid methyl esters (FAMEs) were extracted into hexane. The hexane layer was

removed and dried over sodium sulphate. FAMEs were concentrated using nitrogen

gas. Helium was used as the carrier gas at a flow rate of 1.5 ml min-1. The injector was

identified on comparison of retention time data with internal standard (Sigma-Aldrich,

St. Louis, MO, USA) and corrected using theoretical relative FID response factors

(Ackman, 2002). The peaks were quantified with chemstation chromatography

software (Agilent Technologies, Santa Clara, CA, USA).

6.3. Results and discussion

6.3.1. Thraustochytrid cultivation

102

Page 124: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6

Thraustochytrids are marine microalge known for their ability to produce large

quantities of lipid, including a high ratio of omega-3 fatty acids, particularly DHA.

The carbon source used in the culture medium has a significant effect on biomass and

lipid productivity (Abedini Najafabadi et al., 2015). In this study, we used production

medium with glycerol as a carbon source for thraustochytrids cultivation, since

glycerol is a low cost carbon source that thraustochytrids grow well on (Gupta et al.,

2013; Li et al., 2015). In this study the highest biomass and lipid production achieved

were 12.23 ± 0.4 gL-1 and 29% ±1.9 for Schizochytrium sp. S31 and 14.65 ± 0.54 g-L-

1 and 35% ± 2.1 for Schizochytrium sp. DT3.

6.3.2. Cell disruption by bead mill

Bead milling has been utilised for cell disruption and extraction of intracellular

metabolites from microorganisms such as yeasts (Middelberg, 1995), bacteria (Schütte

et al., 1983), filamentous fungi (Baldwin & Moo-Young, 1991) and even microalgae

such as Chlorella vulgaris (Postma et al., 2015). Nevertheless, the use of bead milling

in lipid extraction from marine microalgae thraustochytrids has not been previously

reported. Important process parameters such as biomass concentration, bead size, RPM

and process time were optimised in the present study to improve the lipid extraction

from thraustochytrids specifically. The experimental process from thraustochytrid cell

disruption to enzyme concentration used in the present study is depicted in Fig. 6.1.

103

Page 125: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6

Figure. 6.1. Schematic diagram of experimental process followed from thraustochytrid

cell disruption to enzymatic omega-3 fatty acids concentration.

The extent of cell disuprtion depends on residence time, shear forces, type of

microroganism, and cell conentration etc. During experimental stage, biomass

concentration, processing time, bead size (0.4-0.6 and 0.8-1.0 mm zirconium oxide)

and processing speed were optimised and the bead load (54%) was kept constant based

on manufactures advice. Various biomass concentrations (1%, 3%, 5% and 7%) and

processing times (in minutes 2, 4, 6 and 8) were investigated. Thraustochytrids cells

disrupted by bead mill were observed under a microscope to validate the cell

disruption. Fig. 6.2 shows typical microscopic images of disrupted cells at optimised

bead mill conditions.

Figure. 6.2. Thraustochytrid cell disruption (40 X magnification) employing bead

mill at 5% biomass concentration, bead size 0.4-0.6 mm, 4500 rpm and 4 min

processing time.

The rate of cell disurption will increase with increasing stirring speed,

however, simultaneous increase in biomass load may decrease this effect. The degree

of cell disruption increased with increasing bead filling volume in the grinding

chamber, up to about 70% where interference among the grinding elements prevents

the establishment of effective cell disruption (Doucha and Livansky, 2008). As shown

in Fig. 6.3, lipid recovery increased up to biomass concentrations of 5% and then

104

Page 126: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6 decreased, probably due to increased viscosity above 5% (Oh et al., 2009). Our results

are in aggrement with a previous study in which the high cell density of

Chlorochoocum was found not to increase cell disruption (Doucha & Lívanský, 2008),

but are inconsistent with a study showing that yeast mass concentration during bead

milling had no significant effect on cell disruption (Baldwin & Moo-Young, 1991;

Mogren et al., 1974).

Figure. 6.3. Optimisation of lipid extraction by bead mill as function of biomass

concentration. Four different biomass concentrations include 1 g, 3 g, 5 g and 7 g

were used.

Lipid yield increased between 1500 rpm and 4500 rpm while further increase

in rpm had no effect on lipid yield (Fig. 6.4). Doucha & Lívanský, 2008) found that

the agitator speed (rpm) between 4500 rpm and 5500 rpm resulted in maximal cell

disintegration in Chlorococum cells. Increasing milling time increased lipid recovery

at all biomass concentrations. Maximum lipid recovery was achieved with 5%

biomass, 4500 rpm and 4 min processing time, with no increase in lipid recovery after

4 minutes since thraustochytrid cells were completely disrupted at that time (Fig. 6.2).

A recent study using bead milling to extract Chlorella cells resulted in 90-95% cell

disruption in 4 min of bead milling (Postma et al., 2015).

20

25

30

35

40

45

0 2 4 6 8

Lipi

d yi

eld

(%)

Time (min)

1 g3 g5 g7 g

105

Page 127: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6

Figure. 6.4. Effect of bead mill agitator speed (rpm) on lipid yield from thraustochytrid

biomass (5% concnetration) at 4 min processing time

In order to understand the effect of bead size on thraustochytrid cell disruption

and lipid yield, we investigated two diffent bead sizes, 0.4-0.6 mm and 0.8-1.0 mm.

The maximum lipid yield (40.5% ± 0.9) was achieved with 0.4-0.6 mm beads

compared to 38.6 ± 1.2 with 0.8-1.0 mm beads. The optimum bead sizes was

previously indicated to be 0.5-0.7 mm for yeast, 0.25-0.5 mm for bacteria and 0.3-0.7

mm for C. vulgaris (Schütte et al., 1983) (Doucha & Lívanský, 2008; Lee et al.,

2012a).

At low biomass concentration 10 Kg m-3 and 8 min processing time, the highest

energy consumption was 528 MJ Kg-1 biomass and lipid yield was 31.45%. Maximum

lipid recovery was observed with 50 Kg m-3 and 4 min processing time, yielding

40.12%, with an estimated energy consumption of 264 MJ Kg-1. The disruption

efficiency was quantified as a measure of lipid recovery and had the highest efficiency

at 50 Kg m-3 and 4 min processing time with low energy consumption compared to

other parameters (Table 6.1). Lee et al, 2010, reported the energy consumption of bead

milling on three different microalgae Botryococcus, Chlorella and Scenedesmus is

equivalent to 504 MJ kg-1 of the dry mass.

Table 6.1. Energy consumption analysis for different mass concentrations and lipid

yields. Schizochytrium sp. 31 biomass was used in different mass concentration and

energy was calculated where the maximum lipid yield was achieved.

05

1015202530354045

1500 2500 3500 4500 5500

Lipi

d yi

eld

(%)

RPM

106

Page 128: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6 Mass

concentration

(Kg m-3)

Energy

consumption (MJ

Kg-1 mass)

Lipid yields (%) Process time

(min)

10 528 31.45 8

30 396 35.87 6

50 264 40.12 4

70 396 38.4 6

6.3.3. Lipid profile analysis

The optimised bead mill processing parameters were applied to lipid extraction

from two thraustochytrid species, resulting in 49.6% lipid as a percentage of dry

weight, with 39.8% DHA in the extracted lipid, for Schizochytrium sp. DT3. Extraction

of Schizochytrium sp. S31 resulted in lipid (39.06%) as a percentage of dry weight,

with 31.4% DHA in the extracted lipid (Fig. 6.5). Our results are in agreement with a

recent study on lipid and DHA production from Aurantiochytrium limacinum SR21,

which achieved maximum lipid and DHA accumulation using glycerol as the carbon

source (Li et al., 2015). A similar study conducted by Gupta et al, (2015) investigated

the suitability of Schizochytrium sp. DT3 for lipid production from enzyme sachrified

hemp, and reported DHA accumulation of 38% (Gupta et al., 2015). Schizochytrium

sp. DT3 oil contained more DHA than did strain S31 and so was investigated further

for the lipase concentration of DHA.

107

Page 129: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6

Figure. 6.5. Fatty acid composition of lipid extracted from thraustochytrids S31 and

DT3.

6.3.4. Selective hydrolysis of Schizochytrium DT3 oil using Candida rugosa lipase

Before hydrolysis, oil extracted from the DT3 strain was confirmed to be

essentially in the triacylglycerol form (Fig. 6.6a). Candida rugosa lipase (CRL) was

used for DHA concentration since it is known to preferentially hydrolyse saturated

(SFAs), then mono-unsaturated fatty acids (MUFAs), with low activity towards

polyunsaturated fatty acids (PUFAs) from TAGs (Kahveci et al., 2010; McNeill et al.,

1996). Enzyme concentrations from 1000 U-4000 U were tested with 500 mg of

thraustochytrid oil where oil was incubated with different enzyme concentrations and

samples were collected at specific time interval and analysed for hydrolysis. Fig. 6.7

shows that the percentage of oil hydrolysed increased with in increase in enzyme

concnetration, particularly with 3000 and 4000 U enzyme resulting in more than 50%

of hydrolysis after 3 hrs. Enzyme concnetrations of more than 3000 U did not increase

the level of thraustochytrid oil hydrolysis, possibly because of substrate limitation. A

similar trend was observed in the hydrolysis of crude palm oil by Candida rugusa

lipase (You & Baharin, 2006). Enzymatic hydrolysis with Candida rugosa lipase

05

101520253035404550

% T

otal

fatty

aci

ds

Major fatty acids

S31 DT3 isolate

108

Page 130: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6 resulted in free fatty acid increase over time (Fig. 6.8), with corresponding increases

in diacyl and monoacyl glycerolds (Fig. 6.6b). Fifty percent hydrolysis of TG occurred

in 3 hours. Candida rugosa lipase preferentially hydrolysed myristic acid (C14:0) and

plamitic acid (C16:0), with DHA and DPA being only slightly hydrolysed (Fig. 6.9).

a

b

Figure. 6.6. Capillary chromatography (iatrascan) profile of thraustochytrid oil a) before and

b) after hydrolysis by Candida rugosa lipase at 40 C.

-2

0

2

4

6

8

10

12

14

16

0 0.1 0.2 0.3 0.4 0.5 0.6

Sign

al (m

V)

Time (min)

TAG

-2

0

2

4

6

8

10

0 0.1 0.2 0.3 0.4 0.5 0.6

FFA

DA

MAG

TAG

Time (min)

Sign

al (m

v)

109

Page 131: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6

Figure. 6.7. Optimisation of enzyme concetrtation. 500 mg of thraustochytrid oil was

incubated with varying enzyme concentrations and samples were withdrawn at

different time intervals.

Figure. 6.8. Lipid class analysis of thraustochytrid oil after hydrolysis with Candida

rugosa lipase at different time intervals.

0

10

20

30

40

50

60

70

0 1 2 3 4 5

Hyd

roly

sis

(%)

Time (h)

1000 (U)2000 (U)3000 (U)4000 (U)

0

20

40

60

80

100

120

0 1 2 3 4

Lipi

d cl

ass (

%)

Time (h)

TAG FFA

110

Page 132: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6

Figure. 6.9. AG and FFA are separated hydrolysed oil and analysed by gas

chromatography. Short chain fatty acids were preferentially hydrolysed form TAG

and DPA and DHA significantly concentatred at glyceride portion.

In the triacylglycerol portion myristic acid was reduced from 10.2% to 4.3% and

palmitic acid from 35.3% to 10.24%. DPA in the triacylglyercol portion increased

from 19.2% to 29.14%, with DHA increasing from 39.8% to 55.45%. CRL has been

shown previously to preferably hydrolyse short and medium chain fatty acids while

increasing omega-3 levels in the remaining triacylglyercides (Kahveci & Xu, 2011

(Nalder et al., 2014). This fatty acids selectivyt is due to the to the deep tunnel like

active site which is buried inside the protein structure. Under optimised conditions in

this study the omega-3 content of thraustochytrid oil was enriched to 88.7%, 1.4 folds

above initial levels. The fatty acid profiles obtained were as expected, from fatty acid

chain length selectivity of Candida rugosa lipase.

6.4. Conclusion

Schizochytrium sp. DT3 is a renewable source of omega-3 fatty acids,

particularly DHA. We optimised a bead mill based cell disruption method for

maximising lipid extraction with lower energy consumption. We proposed a simple

enzymatic process for the concentration of DHA from marine microalgae

0

10

20

30

40

50

60

70

C14:0Myristic

acid

C16:0palmitic

acid

C18:0stearicacid

C18:1Oleic acid

C20:5n3EPA

C22:5n6DPA

C22:6n3DHA

Fatt

y ac

id c

ompo

sitio

n (%

)

Major fatty acids

Native oil AG FFA

111

Page 133: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 6 Schizochytrium sp. DT3 oil using Candida rugosa lipase. This lipase at 3000 U

hydrolysed 50% of thraustochytird oil in 3 h, increasing DHA levels by 1.4 fold.

112

Page 134: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7

Chapter 7: Phospholipase A1 as a catalyst to enrich omega-

3 fatty acids in anchovy and algal oil

113

Page 135: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7 7.1. Introduction

Consumption of fish oils has been shown to have health benefits, mainly in

cardiovascular events, based on several prospective and retrospective studies (Lavie et

al., 2009; Molfino et al., 2014). Most of the benefits of fish oil were attributed to the

polyunsaturated omega-3 ( -3) fatty acids eicosapentaenoic acid (EPA) and

docosahexaenoic acid (DHA) (Udani & Ritz, 2013). The American Heart Association

recommends consumption of 1g/day of omega- fatty acids to patients with coronary

heart diseases (Kralovec et al., 2012). Omega-3 fatty acids constitute approximately

30% of the total fatty acids in natural fish oils. Concentrated esters of omega-3 fatty

acids have been formulated to deliver higher amounts of EPA and DHA per dose,

particularly in pharmaceutical products where 80% EPA and DHA concentrates are

used (Molfino et al., 2014). However, conversion to ethyl esters followed by fractional

distillation and urea concentration is damaging to these oxidatively sensitive omega-3

fatty acids. Also, re-esterification to triacylglycerides requires further damaging

processing of these oils, and also results in statistical distribution of fatty acids on the

glycerol backbone (Catal et al., 2013).

To overcome heat induced degradation, lipases were employed to selectively

hydrolyse omega-3 fatty acids. Triglycerides from natural fish oils are complex in

composition (Miller et al., 2008), besides containing several kinds of fatty acids,

mainly saturated, monounsaturated and polyunsaturated, the fatty acids are also non-

uniformly distributed on the glycerol backbone. 13C NMR spectral studies on anchovy

fish oils show that DHA in more abundant at sn-2 than at sn-1 and -3 positions, while

EPA is nearly statistically distributed across all three positions (Akanbi et al., 2013).

Five lipases were tested for specificity in the hydrolysis of fish oil, using fatty acid

esters as a model system. Discrimination for EPA and DHA was observed with fatty

acid esters but not with fish oils (Lyberg & Adlercreutz, 2008). Another study on lipase

mediated fish oil hydrolysis suggested that hydrolysis is biased towards the chemical

nature of the fatty acid rather than to their positional abundance (Akanbi et al., 2013).

Statement of Contribution to Research for this Chapter: The research outlined in the chapter was carried

out jointly with Mr. Tushar Ranjan Moharana (PhD student) and Dr Madhusudhana Rao at CCMB,

Hyderabad, India, with funding from an Australia-India strategic research grant. The experiments data

shown in Figures 7.2B, 7.3B, 7.5B, 7.6 and 7.7 were performed by Mr. Moharana and the data shown

in Figures 7.1,7.2A, 7.3A, 7.4, 7.5A, 7.8, 7.9 and Table 1 were performed by me.

114

Page 136: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7

Lipases have been extensively used for selective concentration of -3 fatty acids from

fish oils either by fish oil hydrolysis or selective esterification (Akanbi et al., 2013;

Kosugi & Azuma, 1994; Kralovec et al., 2010; Shimada et al., 1997). Lipases possess

some important properties such as selectivity towards chain length and position of fatty

acid in a triglyceride and also discriminate between fatty acids with single and multiple

double bonds. These properties of lipases make them as suitable candidates for

enzymatic concentration of omega-3 fatty acids (Shahidi & Wanasundara, 1998). Most

lipases preferentially hydrolyses saturated and mono-unsaturated fatty acids from

triacylglycerides and strongly discriminate against omega-3 fatty acids, due to the

presence of double bonds that enhance the steric hindrance at lipase active site (Casas-

Godoy et al., 2014; Halldorsson et al., 2004). Most lipases preferentially hydrolyse

EPA over DHA, probably due to the presence of an additional double bond located

closer to the ester bond in DHA (Akanbi et al., 2013; Shimada et al., 1998). In another

study, pancreatic lipase was observed to preferably hydrolyse docosapentaenoic acid

(DPA) over EPA and DHA (Akanbi et al., 2014).

Phospholipase A1 specifically hydrolyses the fatty esters at the sn-1 position

and release 2-acyl lysophospholipid and a fatty acid. Functions of PLA1 are not clearly

established and some PLA1 were reported to show lipase-like activity (Mishra et al.,

2009; Richmond & Smith, 2011). Lipases that show phospholipase activity have been

extensively investigated, but not the reverse (Aloulou et al., 2014). The catalytic

mechanism between lipases and phospholipases is identical; however, the specificity

emerges from the active site properties. Studies have indicated that lipases with a short

lid and a short 9 loop are more suitable to accommodate polar phospholipids (Carrière

et al., 1998). PLA1 was tested for its specificity on triglycerides in this study and the

activity profiles were compared with those of a lipase from Bacillus subtilis. To

investigate the ability of a phospholipase in regioselective hydrolysis of

triacylglycerols, we choose PLA1 since it is likely to be selective for the sn-1, 3

position in triacylglycerols.

Phospholipase A1 is a recombinant enzyme manufactured by Novozymes by

the fusion of a lipase gene from Thermomyce lanuginosus and a phospholipase gene

from Fusarium oxysporum. This enzyme was observed to hydrolyse both phospholipid

and triglycerides with a single active site (Fernandez-Lorente et al., 2008; Mishra et

al., 2009). The phospholipase activity of this enzyme is applied in the commercial

115

Page 137: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7 degumming of vegetable oils and for modifying phospholipids (Sheelu et al., 2008;

Yang et al., 2008). The lipase activity of PLA1 has been applied in organic synthesis

and in the synthesis of structured lipids (Sheelu et al., 2008). PLA1 also successfully

immobilised on different nanoparticles to further improve its catalytic properties (Liu

et al., 2012; Mishra et al., 2009; Wang et al., 2010). Therefore, in this study, we

attempted to use phospholipase A1 for the concentration of omega-3 fatty acids from

anchovy fish oil. We investigated the chain length selectivity of phospholipase A1

using p-NP esters. The ability of Phospholipase A1 to concentrate omega-3 fatty acids

into the glyceride portion was investigated by examining the preferential fatty acid

hydrolysis using gas chromatography and 13C NMR spectroscopy.

7.2. Material and Methods

7.2.1. Chemicals

Bleached Anchovy oil was supplied by Ocean Nutrition Canada (Canada). The

major fatty acid composition of native oil was determined by Gas chromatography

analysis (Akanbi et al., 2013). Phospholipase A1 (Lecitase Ultra®) and gas

chromatography standards were procured from Sigma-Aldrich (Castle Hill, Australia)

and hexane (Merck). All the buffers were made using analytical grade chemicals. All

solvents used were either analytical grade or higher. Gum Arabic and phospholipase

A1 (PLA1) and methyl nonadecanoate were purchased from Sigma Aldrich. TLC

plates were purchased from Merck.

7.2.2. Hydrolysis of anchovy oil

Reaction mixture containing 50 mM Tris, 25 mM CaCl2, 5% (w/v) gum Arabic

at pH 8.00 and 5% (v/v) anchovy oil were micellised by sonication. 100 unit of PLA1

was added and the reaction was carried at 37 C at constant stirring under nitrogen.

itrating

against 1 M NaOH. 2 ml of sample were withdrawn at every time point and free fatty

acid was separated from glyceride by solvent extraction and the purity of both the

fractions were analysed by TLC. Solvent was evaporated under nitrogen gas and

fractions stored in an airtight polypropylene tube at -20 C for further analysis. Oxygen

was removed by blowing nitrogen whenever the sample was exposed to air.

116

Page 138: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7 7.2.3. Thraustochytrid oil hydrolysis by phospholipase A1

Thraustochytrid lipid extraction and analysis was described in section 6.2.5,

6.2.7 and 6.2.8. Phospholipase hydrolysis was performed based on the method

described in section 6.2.6 with some modifications. 500 mg of thraustochytrid oil and

1.8 ml of Tris HCL buffer (50 mM, pH 7.0) were taken in 25 ml round bottom flask

and heated up to 45 oC and 300 rpm stirring. Substrates were stirred for 30 min and

Phospholipase A1 (2500 U/g oil) added to initiate the hydrolysis. Samples were

collected periodically and analysed.

7.2.4. IATROSCAN analysis

Both the unhydrolysed and hydrolysed portions of the fish oil were analysed

by capillary chromatography with flame ionization detector (Iatroscan MK5, Iatron

Laboratories Inc., Tokyo, Japan). The Iatroscan settings were: air flow rate; 200

ml/min, hydrogen flow rate; 160 ml/min and scan speed of 30s/scan. Under these

conditions, the chromarods were cleaned by scanning twice before applying samples.

One microliter of each lipid fraction in hexane was spotted onto the rods with the aid

of an auto pipette along the line of origin on the rod holder and developed for 22

minutes in a solvent tank containing hexane/diethyl ether/acetic acid (60:17:0.2,

vol/vol/vol). TLC standards purchased from Nu-Chek Prep were used to identify each

lipid class (Akanbi et al., 2013).

7.2.5. Methylation and GC analysis

5 μL (both free fatty acid and glyceride) of sample was methylated by using

acetyl chloride in methanol as described by Christie et al 2010 (Christie & Han, 2010).

Hydrolysed and unhydrolysed fish oil was analysed using gas chromatography

(Agilent 6890) with flame ionisation detector (FID), equipped with a Supelcowax 10

. Fatty acids

were converted to methyl esters by acid-catalysed esterification with minor

modifications. 1 ml toluene was added to the methylation tubes followed by the

(5mM) of internal standard, methyl nonadecanoate (C19:0) and 200

(1mM) of butylated hydroxytoluene (BHT). 2 ml acidic methanol (prepared by

adding 10% acetyl chloride in methanol drop wise on ice bath) was also added to the

tube and kept for overnight incubation at 50°C. Fatty acid methyl esters (FAMEs) were

117

Page 139: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7 extracted into hexane. The hexane layer was removed and dried over sodium sulphate.

FAMEs were concentrated using nitrogen gas and analysed by gas chromatography.

Helium was used as the carrier gas at a flow rate of 1.5 ml min-1. The injector was

identified on comparison of retention time data with external standards (Sigma-

Aldrich, St. Louis, MO, USA) and corrected using theoretical relative FID response

factors (Ackman, 2002). Peaks were quantified with Chemstation chromatography

software (Agilent Technologies, Santa Clara, CA, USA).

7.2.6. Analysis of positional distribution of fatty acid by NMR

300 μl of glyceride were dissolved in 1 ml of CDCl3 (99.8% pure). NMR

spectra were collected by using Bruker 600MHz NMR machine. Peak corresponding

to different fatty acid at different position were assigned as described by (Suárez et al.,

2010). Relative quantification was done by comparing the area under each peak.

Quantitative 13C NMR spectra of the unhydrolysed oils were recorded under

continuous 1H decoupling at 24°C.In order to quantify the residue of each fatty acid at

different positions, peak area ratios were analysed by integration and presented in

percentages (Akanbi et al., 2013).

7.3. Results and discussion

7.3.1. Lipase activity of Phospholipase A1

PLA1 used in this study was isolated from Thermomyces lanuginosus

(LecitaseTM) and is a recombinant enzyme preparation. This enzyme has been explored

primarily for the degumming of vegetable oils (Yang et al., 2006). Some studies

include the use of PLA1 lipase activity for the resolution of mandelate esters and the

regioselective hydrolysis of carbohydrate esters (Mishra et al., 2009). Lipase-like

activity of PLA1 was reported earlier. However its ability to hydrolyse triglycerides is

only marginal in the vegetable oil degumming processes. PLA1 was previously

characterised for substrate selectivity and it was observed that this enzyme does

hydrolyse C2-esters (Mishra et al., 2009). Initially the lipase activity of PLA1 was

tested using p-nitrophenyl esters of various chain length fatty acids. Some of the p-NP

esters were synthesised based on published methods (Nalder et al., 2014; Neises &

Steglich, 1978). Fig. 7.1 shows that PLA1 was able to hydrolyse esters of fatty acid

118

Page 140: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7 chain lengths from C4–C20 with comparable efficiency. The activity was highest for

C10 and lowest with C4 ester (40% of C10). The presence of unsaturation, single or

multiple double bonds, did not impact the activity appreciably. However, esters of

DHA were not efficiently hydrolysed. A similar chain length study was performed for

lipase from Bacillus subtilis (BSL) and it was found that C8 chains were most

efficiently hydrolysed (Dartois et al., 1992).

Figure.7.1 Phospholipase A1 activity towards p-NP esters with different acyl

chains. Activity of PLA1 on p-NP-decanoate was taken as 100%.

Lipase activity of phospholipases and vice versa was studied in few cases. The

ratio of lipase to phospholipase activity of lipases or phospholipases varies widely and

the ratio is not only related to the structure of the lipase but also to the reaction system

(Yang et al., 2008). For example, PLA1 from Thermomyces lanuginosus used in this

study shows hydrolytic activity against both triacylglcyerides and phospholipids and

is used to degum oils (Sheelu et al., 2008). PLA1 shows predominately phospholipase

activity at reaction temperatures above 40 oC. Pancreatic phospholipases hydrolyse

phospholipids in the aqueous phase, while lipase from Fusarium oxysporum

hydrolyses phospholipids in the oil phase (Yang et al., 2008).

7.3.2. Hydrolysis of Anchovy oil by PLA1

Initially hydrolysis of anchovy oil was tested in an oil-in-water emulsion and

the extent of hydrolysis was monitored using Iatroscan. Fig.7.2A shows the scan

119

Page 141: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7 before and after hydrolysis of the oil by PLA1. To obtain more quantitative kinetics of

the hydrolysis, anchovy oil hydrolysis was monitored by a pH Stat, while maintaining

a constant pH, using 5% Gum Arabic as emulsifier. Initially we measured the rate of

hydrolysis of anchovy oil by PLA1 and BSL (Fig.7.2B). Both enzymes were able to

hydrolyse the oil to a comparable extent. At 37 oC the reaction reached 45% hydrolysis

and then plateaued.

Figure. 7.2. Time course of hydrolysis of anchovy fish oil by PLA1 and BSL

A, Iatroscan of unhydrolysed (light trace) and hydrolysed (darker trace) fraction of

anchovy oil by PLA1. B, pH stat based determination of rate of hydrolysis of anchovy

oil in the presence of PLA1 (closed circle) and BSL (open circle).

The fatty acid fraction was separated from the glyceride fraction, methylated

and analysed by GC. Fig.7.3A shows the time course of release of various fatty acids

120

Page 142: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7 from anchovy oil by PLA1. Both saturated and monounsaturated fatty acids were

extensively hydrolysed (approximately 40% hydrolysis), while the hydrolysis of both

the omega-3 fatty acids EPA and DHA was relatively poor 15% and 4%, respectively.

Identical experiments were also performed with BSL (Fig.3B). BSL also hydrolysed

oil with similar efficiency. However, the extents of hydrolysis of various fatty acids

were similar. While the extents of hydrolysis of unsaturated, monounsaturated and

EPA were equal (50%), hydrolysis of DHA was marginally less at 30%. The fatty acid

products obtained after 3 h hydrolysis are shown in Fig.7.4. The data suggests that

PLA1 discriminates against EPA and DHA, while BSL shows marginal discrimination

against DHA. A significant five-fold higher discrimination of DHA by PLA1

compared to BSL was observed.

121

Page 143: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7 Figure. 7.3. Time course of hydrolysis of anchovy oil and fatty acid distribution in the

hydrolysate. Anchovy oil was subjected to hydrolysis by PLA1 (A) and BSL (B). The

reaction product at various times was subjected to methylation and GC analysis.

Percent hydrolysis of each fatty acid was calculated based on its hydrolysis at tx as a

fraction of to.

PLA1 specifically hydrolyses sn-1 acyl esters from phospholipids producing

FFAs and lysophospholipase. PLA1 enzymes normally exhibit very little

lysophospholipase or lipase activity. However, some PLA1 enzymes exhibit some

neutral lipase activity. PLA1 enzymes may be descendants of neutral lipases, and

several PLA1 sequences show substantial sequence similarity to the well characterised

pancreatic, hepatic and endothelial lipases (Aoki et al., 2007; Richmond & Smith,

2011). Although PLA1 enzymes are found in a wide variety of cells and tissues, only

a small number of PLA1 enzymes have been cloned and their substrate preferences

have not been well studied (Richmond & Smith, 2011). Several structural and protein

engineering studies on lipases have focused on the architecture of the active site and

its substrate preferences. The hydrophobicity and 9

loop and the lid domain of lipases play a selective role in preferring a triglyceride or a

phospholipid (Aoki et al., 2007; Tiesinga et al., 2007). The lid domain of guinea pig

pancreatic lipase related protein 2(GPLRP2) is reduced in size and consequent

exposure of hydrophilic residues enables the active site to accept phospholipids and

larger galacto lipids (Carrière et al., 1998). Human pancreatic lipase that does not show

phospholipase activity was not able to accommodate the phospholipid molecule.

Sequence alignment of porcine pancreatic lipases indicates that Val260 is critical for

interaction with lipids (Van Kampen et al., 1998).

PLA1 from Thermomyces lanuginosus, used in this study, discriminated

against DHA and partly against EPA when hydrolysing anchovy oil. In similar

experiments using a lipase from Bacillus subtilis there was no observed selectivity

against EPA and DHA. Anchovy oil has DHA preferentially located at the sn-2

position (60%), compared to the sn-1, 3 positions (40%). If PLA1 preferentially

hydrolyse at the sn-1, 3 position in TG oils then more DHA than EPA should be present

in the unhydrolysed fraction. This was observed in our experiments (Fig.7.4). In our

study PLA1 poorly hydrolysed DHA at all positions, and preferably hydrolysed EPA

122

Page 144: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7 at the sn-1, 3 positions. BSL preferentially hydrolysed DHA at the sn-2 position, while

preferentially hydrolysing EPA at sn-1, 3.

7.3.3. 13C NMR studies of the anchovy oil hydrolysis by PLA1 and BSL

13CNMR spectra of complex triglycerides, such as fish oils, can provide

positional information on the various fatty acids in the oil. This information enables

the study of the positional specificity of hydrolysis by PLA1 and BSL and also to

obtain quantitative information on the extent of hydrolysis of fatty acids at each

position. Initially we acquired the positional information of various fatty acids in

anchovy oil. Table 7.1 provides the details of positional distribution of fatty acids in

Anchovy oil. The ratio of abundance of various fatty acids at sn-1, 3 and sn-2 positions

was observed to be saturated 1.64, monounsaturated 3.06, EPA 4.2 and DHA 0.6. A

ratio of 2 suggests the fatty acids are equally distributed. A ratio of 0.6 for DHA

indicates it is predominantly present at the sn-2 position. The ratio is high (4.2) for

EPA indicating its preferential distribution at sn-1, 3 position, compared with the

distribution of DHA.

Table 7.1. Positional distribution of various fatty acids in anchovy oil.

Fatty acid

Distribution (%)

sn-1,3 sn-2

Saturated 22.9 14.0

Mono unsaturated 18.7 6.1

Stearidonic acid (STA) 6.2 2.7

EPA 14.9 3.5

DHA 3.9 6.8

Anchovy oil hydrolysis by PLA1 and BSL was allowed to proceed until 30%

of the oil was hydrolysed and then the unhydrolysed portion of the oil was extracted

and 13C NMR spectra were obtained. This was repeated with unhydrolysed anchovy

oil. Figure 7.5 shows the 13C NMR spectra of hydrolysed and unhydrolysed anchovy

oil subjected to PLA1 (Fig. 7.5A) and BSL (Fig. 7.5B) treatment. The overlay of

hydrolysed and unhydrolysed spectra for both PLA1 and BSL clearly demonstrates

that BSL preferentially hydrolyses EPA and DHA more than does PLA1. Fatty acids

123

Page 145: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7 remaining in the unhydrolysed portion of the oils with PLA1 and BSL are shown in

Fig. 7.4. Saturated and monounsaturated fatty acids were preferentially hydrolysed by

both the enzymes while omega-3 fatty acids were discriminated against. BSL also

preferentially hydrolysed EPA over DHA.

Figure. 7.4. Percent hydrolysis of various fatty acids (from Fig.7.3) at 32% hydrolysis

of anchovy oil by PLA1 (light) and BSL (dark). The relative proportions at other time

points were also similar.

124

Page 146: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7

Figure. 7.5. 13C NMR spectra of the unhydrolysed oil fraction before (blue) and after

(red) 30% hydrolysis of anchovy oil by PLA1 (A) and BSL (B). Fatty acids and their

positions were marked. Sat (saturated), mono (monounsaturated), STA (stearidonic

acid), EPA (eicosapentaenoic acid) and DHA (docosahexaenoic acid).

The percent accumulation data shown in Fig. 7.6A indicates that EPA at sn-1

was more efficiently hydrolysed by BSL than by PLA1. BSL hydrolysed EPA at sn-1

but not DHA at sn-1or at sn-2, while PLA1 had no preference for EPA over DHA.

These results indicate that the discrimination against omega-3 fatty acids by PLA1 is

a result of the chemical nature of the fatty acid and not its position on the TG. To verify

this observation, we have performed hydrolysis of a structured triglyceride, 1, 3-

dioleoyl-2-hexadecanoic glyceride, using PLA1 and BSL and estimated the fatty acid

products by GC (Fig.7.7). If PLA1 shows absolute preference for fatty acids at the sn-

Sn-1,3 Sat

Sn-1,3 Mono unsat

Sn-1,3STA

M

Sn-1,3EPA

Sn-2 Sat

Sn-2 Mono Unsat

Sn-1,3 STA

Sn-2 DHA

Sn-1,3DHASn-

Sn-2 EPA

A

B

125

Page 147: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7 1, 3 position, the hydrolysis of hexadecanoic acid at sn-2 should be much lower than

that of oleic acid. The difference between the specificity of hydrolysis by PLA1 and

BSL was also negligible, indicating that PLA1 did not show any specificity towards

the sn-1,3positions. However, there is a distinct preference against DHA in anchovy

oil, indicating the fatty acid specificity of this enzyme.

Figure. 7.6. Percent accumulation of different fatty acid classes in TG fraction after

30% hydrolysis of anchovy oil by PLA1 (dark) and BSL (light). A, Percent

accumulation of each fatty acid is calculated as (UHFAx-TotalFAx)/TotalFAx) X 100).

Each of the fatty acid fractions were calculated based on peak area in the NMR spectra.

B, Percent accumulation data replotted with positional information.

126

Page 148: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7

Hexadecanoic acid 9-Octadecenoic acid0

10

20

Fatty

aci

d re

leas

ed (

mol

es)

Figure. 7.7. Lipase hydrolysis of structured triglyceride1, 3-dioeoyl-2-myristic

triglyceride was subjected to hydrolysis by PLA1 (Light) and BSL (Dark) and the

released fatty acids were quantified by GC analysis.

To assess the positional specificity of these hydrolases we have employed a

structured glyceride with fixed positional distribution of fatty acids. Using 3-dioleoyl-

2-myristic glyceride, a structured TG, neither PLA1 nor BSL exhibited any position

specific hydrolysis under our reaction conditions, with both fatty acids being equally

hydrolysed. However, with anchovy oil the hydrolysis by PLA1 clearly shows bias

against DHA, indicating fatty acid selectivity against polyunsaturated fatty acids. In a

previous study five lipases were tested for their specificity by using fish oil and methyl

esters of EPA, DHA and palmitic acid. All lipases discriminated against EPA and

DHA when presented as methyl esters. However, lipase from Thermomyces has shown

discrimination against DHA, particularly in the early stage of the hydrolysis reaction.

A study observed that Candida rugosa lipase was most efficient in the enrichment of

DHA in the glyceride fraction (Lyberg & Adlercreutz, 2008). In another study lipase

from Candida rugosa discriminated omega-3 fatty acids from other fatty acids during

hydrolysis of sardine oil (Okada & Morrissey, 2007). These studies indicate that the

specificity in hydrolysis of natural oils by lipases or phospholipases is strongly

dependent on the temperature and duration. Also, positional selectivity can found fatty

127

Page 149: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7 acid selectivity analyses in natural triacylglycerides, since positional distribution is not

always analysed in selectivity studies.

Our study and similar studies with PLA1 lipase from Thermomyces

lanuginosus suggests that hydrolysis is due to fatty acid selectivity more than to

regioselectivity (Akanbi et al., 2013). A related previous study showed that substrate

binding was dependent on the relative volume of the substrate to the volume of the

active site (Kamal et al., 2015). Thirty eight lipases were investigated in silico, for

their affinity to structured TGs with different chemical and positional compositions.

This study indicated that binding affinity differences between various substrates with

lipases is complex, but helped to identify active site positions that are critical to

binding. This information could be useful for designing site saturation mutagenesis to

identify amino acid substitutions that would enhance hydrolysis of specific fatty acids.

Few studies have successfully identified amino acid positions that are important in

binding TG and phospholipids (Carrière et al., 1998). Further studies are required to

enable the design of enzymes that can specifically hydrolyse omega-3 fatty acids.

7.3.4. Thraustochytrid oil hydrolysis by phospholipase A1

Phospholipase A1 is an engineered carboxyl ester hydrolase, produced from

Thermomyces lanuginosus/Fusarium oxysporum by genetically modified Aspergillus

oryzae. Phospholipase A1 possess activity towards both phospholipids and

triglycerides. (Mishra et al., 2009). Research has focused on its phospholipase activity

and little information is available on the production of structured glycerides using

phospholipase A1 from oils (Liu et al., 2012; Liu et al., 2011). DT3 oil was hydrolysed

to determine phospholipase A1 activity towards concentration of omega-3 fatty acids.

Hydrolysis was gradually increased with incubation time and reached 50% at 150 min.

Fig. 7.8 illustrates the percentage of hydrolysis of DT3 oil by phospholipase A1.

128

Page 150: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7

Figure. 7.8. Lipid class analysis of thraustochytrid oil after hydrolysis with

phospholipase at different time intervals.

Fatty acids profiles of glycerol and free fatty acids of hydrolysed DT3 oil were

analysed by gas chromatography and results are shown in Fig. 7.9. After 150 min of

phospholipase A1 hydrolysis, a significant amount of short chain fatty acids and

monounsaturated fatty acids were cleaved from the glycerol back bone and

concentrated in the free fatty acids portion. In the triacylglycerol portion, the major

saturated fatty acids were reduced, myristic acid was reduced from 10.2% to 5.3% and

palmitic acid from 35.2% to 16.24%. EPA was equally distributed in the free fatty acid

and glyceride portions and DHA was significantly enriched in the glyceride portion,

indicating a strong selectivity for EPA over DHA (Fig. 7.9). The DHA concentration

was enriched by 1.3 fold compared to native oil. These selectivity is similar to those

observed for hydrolysis of fish oil using lipase TL100 (Akanbi et al., 2013).

0

20

40

60

80

100

120

0 30 60 90 120 150 180 210

Hyd

roly

sis

(%)

Time (min)

TAG FFA

129

Page 151: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 7

Figure. 7.9. Fatty acid profiles of hydrolysed and unhydrolysed DT3 oil.

7.4. Conclusion

The position selective hydrolysis of PLA1 was tested for the hydrolysis of

anchovy oil. Fatty acid chain length selectivity of PLA1 was investigated using p-NP

esters of chain length C2 to C22, including the long chain PUFAs, EPA and DHA. For

anchovy oil hydrolysis SFAs and MUFAs were preferentially hydrolysed by PLA1

and omega-3 fatty acids were discriminated against. Of the omega-3 fatty acids, EPA

was more preferentially hydrolysed than was DHA. Lipase from Bacillus subtilis did

not show discrimination of fatty acids to the same extent as PLA1. Hydrolysis of the

structured TG, 1, 3–dioleoyl-2-hexadecanoic glyceride, supported that the

discrimination property of PLA1 was primarily due to the chemical nature of the fatty

acid rather than its position in the triacylglyceride. PLA1 is a potential catalyst for

selective enrichment of omega-3 fatty acids in triacylglycerides.

Phospholipase A1 was used to catalyse the hydrolysis of thraustochytrid DT3

oil to enrich DHA in glyceride portion, with short chain fatty acids being preferentially

cleaved from the triacylglyceride. This results show that phospholipase A1 can be used

to concentrate DHA from triacylglycerides.

0

10

20

30

40

50

60

C14:0Myristic

acid

C16:0palmitic

acid

C18:0stearicacid

C18:1Oleic acid

C20:5n3EPA

C22:5n6DPA

C22:6n3DHA

Com

posi

tion

(%)

Major fatty acids

Native oil AG FFA

130

Page 152: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 8

Chapter 8. Summary and future aspects

131

Page 153: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 8

The marine ecosystem is a large relatively unexplored pool for the isolation of

potential compounds and materials for industrial and biotechnological applications.

Marine microorganisms remain a major source for scientific discovery. The number

of microorganisms isolated and characterised for their potential metabolites is

increasing each year. Researchers have isolated bioactive compounds (lipids,

enzymes, polysaccharides, proteins and pigments) from a range of marine inhabitants.

The demand for novel enzymes is also increasing, particularly for the production of

chemically and structurally important molecules with industrial applicability.

Naturally marine environments provide unique genetic diversity and life habitats such

as nutrient rich regions, nutrient limited regions, high salinity, high pressure, low

temperature and light conditions. Microalgae lipid production is an area that is being

explored, although the cost of production is high, partly due to downstream processing

costs. Thus, cost-effective downstream processing methods play an important role in

achieving commercial scale production of microalgae derived metabolites.

The main objective of this study is the isolation and selection of suitable

thraustochytrid strains for lipid production, the development of lipid extraction and

omega-3 enzymatic concentration methods applicable to the production of

thraustochytrid derived oil. One aim was to develop a rapid lipid quantification method

to screen newly isolated thraustochytrids. Another aim was to optimise lipid and lipase

extraction methods. A third aim was to concentrate omega-3 fatty acids from

thraustochytrids oil using lipases.

The first aim of the study was to develop a rapid colorimetric method for lipid

quantification from thraustochytrid cells. The conventional lipid quantification

methods are laborious and time consuming. Initially, to evaluate the colorimetric

method, different lipid samples (plant oils, pure triolein, DHA and EPA) were

quantified. A straight line showed the range of linearity of the method with different

samples. In addition, the false positives from non-lipid components (glucose, glycerol

and proteins) did not interfere with lipid quantification. In this study, four different

thraustochytrids strains, two of them new isolates, were quantified for lipid content

and the results compared with those obtained using a conventional gravimetric method.

The results obtained from both methods were similar within experimental error.

132

Page 154: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 8

Chapter 3 describes the effect of different solvents and cell disruption methods

on lipid extraction from thraustochytrids. A total of nine solvents; chloroform,

dichloromethane, diethyl ether, ethanol, heptane, hexane, isopropanol, methanol and

toluene and their combinations, were investigated for lipid extraction. Maximum lipid

extraction for single solvents were hexane (12.5%), heptane (11%) and chloroform

(9.7%). However, a solvent mixture of chloroform and methanol (2:1) showed

maximum lipid extraction of 22%, followed by chloroform and hexane (2:1) with

13.4%. Furthermore, different cell disruption methods were evaluated to improve lipid

extraction from Schizochytrium S31 and Thraustochytrium AMCQS5-5. Cell

disruption resulted in increased lipid extraction from both Schizochytrium S31 and

Thraustochytrium AMCQS5-5, and maximal lipid extraction was obtained using

osmotic shock, with 48.7% lipid extracted from Schizochytrium S31 and 29.1% lipid

from Thraustochytrium AMCQS5-5. The other methods, grinding (44.6%), sonication

(31%) and shake mill (30.5%) extracted higher lipids than the control method (22%).

The fatty acid profiles of the lipids extracted from thraustochytrids were investigated

using GC. GC analysis revealed the suitability of S31 fatty acid profiles for biodiesel

production and AMCQS5-5 profiles for nutraceutical applications. In addition, in

silico analysis of fatty acid profiles confirms the potential of S31 lipid for biodiesel

production. The energy consumption was evaluated for the methods investigated in

this study to aid in determining their potential use in large scale processing. The

osmotic shock method was found to be the most economical method compared to other

methods for industrially scaled-up.

Thraustochytrids have been documented for polyunsaturated fatty acids

production, particularly DHA, but little information is available about their enzymatic

activities. The first part of this study (Chapter 4) describes screening of enzymatic

activities from thraustochytrid strains using the API ZYM kit. The isolates investigated

in this study were collected form two different collection trips. The investigated

thraustochytrid strains exhibited high esterase, lipase, acid and alkaline phosphatase

activities and the maximum lipase activity was observed in AMCQS1-9 followed by

AMCQS-4b27. In addition, the effect of different cell disruption methods on protein

release and lipase activity was evaluated from AMCQS1-9 and AMCQS-4b27 strains.

In this study, four different cell disruption methods, bead vortexing, grinding with

liquid nitrogen, sonication and homogenisation, were compared for impact on protein

and lipase recovery from thraustochytrids. Strain AMCQ-4b27 cells were more

133

Page 155: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 8 sensitive to protein release than those from AMCQS1-9 and sonication was found to

be the suitable method for lipase yield from AMCQ-4b27 (0.903 IU/μg) and

AMCQS1-9 (1.44 IU/μg). Lipase activity extracted from strain AMCQS1-9 was

approximately 1.5 fold more than that from AMCQ-4b27.

Lipase production is influenced by the culture conditions. Chapter 5

demonstrated the optimisation of Schizochytrium S31 culture conditions for enhancing

lipase production in a stirred tank reactor. In addition, sonication parameters were

optimised to improve lipase extraction yields. Initially, Tween 80 concentration was

optimised and a Tween 80 concentration of 1% gave the maximum biomass and lipase

productivity. The optimum temperature was 25 oC and the optimum pH was 6. The

process was scaled-up in a 2 L stirred tank reactor to monitor biomass, lipid and lipase

productions and utilisation of substrate and change in pH. To quantify lipid content,

fatty acid composition and lipase activity, biomass was collected and analysed at

different time intervals up to 9 days. Biomass and lipase production reached a

maximum after 5 days of incubation. Based on the results obtained from Chapter 4,

sonication was resulted in the highest level of lipase extraction from thraustochytrids.

In this study, we optimised sonication parameters to enhance lipase extraction.

Optimised sonication conditions resulted in a 5 fold increase in lipase activity

extraction. Novel Schizochytrium S31 lipase was employed to hydrolyse different

chain length p-NP esters. The extracted lipase preferentially hydrolysed shorter chain

fatty acids, however, DHA was preferentially hydrolysed over EPA.

Chapter 6 describes the use of bead milling in thraustochytrid lipid extraction

and enzymatic concentration of DHA from extracted lipids. In this study,

thraustochytrids were grown in glycerol, a low cost carbon source for biomass and

lipid production. The highest level of biomass and lipid produced were 12.23 ± 0.4 g

L-1 and 29% ± 1.9 for Schizochytrium sp. S31 and 14.65 ± 0.54 g L-1 and 35% ± 2.1

for Schizochytrium sp. DT3, respectively. Schizochytrium S31 biomass was used for

optimising lipid extraction using a bead mill. Important process parameters such as

biomass concentration, bead size, RPM and process time were optimised. The highest

lipid yield (40.5%) was obtained with biomass concentration (5%), bead size (0.4-0.6

mm), RPM (4500) and 4 min processing time. Fatty acid profiles of extracted lipid

from the both strains was compared and DHA accumulation was observed more in

strain DT3 (39.8%) than S31 (31.4%). Candida rugosa lipase was used to concentrate

DHA in the glyceride portion. Initially, the lipase concentration was optimised and it 134

Page 156: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 8 was found that 3000 U resulted in 50% oil hydrolysis in 3 hours. The presence of lipid

classes FFA, DAG and MAG were observed by TLC analysis. Under optimised

conditions the omega-3 fatty acid content of thraustochytrid oil was enriched to 88.7%,

a 1.4 fold increase on initial levels.

We also used phospholipase A1, a carboxyl ester hydrolase, for concentrating

omega-3 fatty acids from Schizochytrium DT3 oil extracted from optimised bead

milling process. At 2.5 h of incubation, oil hydrolysis reached 50%, and the fatty acid

profiles were analysed using GC. EPA was equally distributed in the free fatty acid

and glyceride portions, while DHA was significantly enriched in glyceride portion,

indicating a strong selectivity for EPA over DHA.

Chapter 7 demonstrated the application of phospholipase as a catalyst for

enriching omega-3 fatty acids from anchovy fish oil. Initially we investigated the

substrate selectivity of PLA1 using p-NP esters chain lengths of C2 to C22, including

the long chain PUFAs, EPA and DHA. PLA1 showed preferential hydrolysis of EPA

and discriminated against DHA. In anchovy oil hydrolysis, the saturated and mono-

unsaturated fatty acids were hydrolysed from the glyceride portion. Of the omega-3

fatty acids, EPA was preferentially hydrolysed over DHA. We compared fish oil

hydrolysis with the lipase from Bacillus subtilis, which did not discriminate between

fatty acids to the same extent as PLA1. Furthermore, hydrolysis of the structured lipid

1, 3-dioleoyl-2-hexadecanoic confirmed that the observed selectivity of PLA1 was due

to fatty acid selectivity and not positional selectivity.

Microalgae derived biofuels and nutraceuticals are an alternative approach for

sustainable production. However, there are several challenges to be addressed. The

large scale production of lipids from microalgae is only possible with the isolation and

selection of microalgae with high biomass and lipid productivities. This can be

achieved by using robust techniques in isolation and selection of microalgae from

varied habitats for commercial scale production. Although the lipid extraction methods

explored in this thesis were optimised using dry biomass, it is necessary to explore

extraction of wet biomass, since biomass drying is an expensive process. The

development of enzyme assisted and nano particle based lipid extraction and the use

of natural solvents such as terpenes and organic-nanoclays reduce the environmental

impacts and waste generation during processing. Thraustochytrids appears to have

135

Page 157: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

Chapter 8 some potentially useful lipases, including lipases with DHA versus DPA selectivity,

although the structures of these lipases still remain to be identified.

136

Page 158: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References References

Abedini Najafabadi, H., Malekzadeh, M., Jalilian, F., Vossoughi, M., Pazuki, G. 2015.

Effect of various carbon sources on biomass and lipid production of Chlorella

vulgaris during nutrient sufficient and nitrogen starvation conditions.

Bioresource Technology, 180, 311-317.

Abuajah, C., Ogbonna, A., Osuji, C. 2015. Functional components and medicinal

properties of food: a review. Journal of Food Science and Technology, 52(5),

2522-2529.

Ackman, R.G. 2002. The gas chromatograph in practical analyses of common and

uncommon fatty acids for the 21st century. Analytica Chimica Acta, 465(1–2),

175-192.

Adarme-Vega, T.C., Lim, D.K.Y., Timmins, M., Vernen, F., Li, Y., Schenk, P.M.

2012. Microalgal biofactories: a promising approach towards sustainable

omega-3 fatty acid production. Microbial Cell Factories, 11(1), 96.

Adarme-Vega, T.C., Thomas-Hall, S.R., Schenk, P.M. 2014. Towards sustainable

sources for omega-3 fatty acids production. Current Opinion in Biotechnology,

26, 14-18.

Ahmad, F.B., Zhang, Z., Doherty, W.O.S., O’Hara, I.M. 2015. A multi-criteria

analysis approach for ranking and selection of microorganisms for the

production of oils for biodiesel production. Bioresource Technology, 190, 264-

273.

Akanbi, T.O., Adcock, J.L., Barrow, C.J. 2013. Selective concentration of EPA and

DHA using Thermomyces lanuginosus lipase is due to fatty acid selectivity and

not regioselectivity. Food Chemistry, 138(1), 615-620.

Akanbi, T.O., Sinclair, A.J., Barrow, C.J. 2014. Pancreatic lipase selectively

hydrolyses DPA over EPA and DHA due to location of double bonds in the

fatty acid rather than regioselectivity. Food Chemistry, 160, 61-66.

Aloulou, A., Frikha, F., Noiriel, A., Bou Ali, M., Abousalham, A. 2014. Kinetic and

structural characterization of triacylglycerol lipases possessing phospholipase

A1 activity. Biochimica et Biophysica Acta (BBA) - Molecular and Cell

Biology of Lipids, 1841(4), 581-587.

Amaro, H.M., Guedes, A.C., Malcata, F.X. 2011. Advances and perspectives in using

microalgae to produce biodiesel. Applied Energy, 88(10), 3402-3410.

137

Page 159: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Amin, S. 2009. Review on biofuel oil and gas production processes from microalgae.

Energy Conversion and Management, 50(7), 1834-1840.

Ando, Y., Nishimura, K., Aoyanagi, N., Takagi, T. 1992. Stereospecific analysis of

fish oil triacyl-sn-glycerols. Journal of the American Oil Chemists' Society,

69(5), 417-424.

Aoki, J., Inoue, A., Makide, K., Saiki, N., Arai, H. 2007. Structure and function of

extracellular phospholipase A1 belonging to the pancreatic lipase gene family.

Biochimie, 89(2), 197-204.

Apar, D.K., Özbek, B. 2008. Protein releasing kinetics of bakers' yeast cells by

ultrasound. Chemical and Biochemical Engineering Quarterly, 22(1), 113-

118.

Arafiles, K.H.V., Alcantara, J.C.O., Batoon, J.A.L., Galura, F.S., Cordero, P.R.F.,

Leaño, E.M., Dedeles, G.R. 2011. Cultural optimization of thraustochytrids for

biomass and fatty acid production. Mycosphere, 2, 521-531.

Araujo, G.S., Matos, L.J.B.L., Fernandes, J.O., Cartaxo, S.J.M., Gonçalves, L.R.B.,

Fernandes, F.A.N., Farias, W.R.L. 2013. Extraction of lipids from microalgae

by ultrasound application: Prospection of the optimal extraction method.

Ultrasonics Sonochemistry, 20(1), 95-98.

Arnal, J.M., Sancho, M., Iborra, I., Gozálvez, J.M., Santafé, A., Lora, J. 2005.

Concentration of brines from RO desalination plants by natural evaporation.

Desalination, 182(1-3), 435-439.

Asif, M. 2014. Role of polyunsaturated fatty acids in cardiovascular diseases. Journal

of Pharmaceutical Care, 2(1), 37-44.

Azeem, M.W., Hanif, M.A., Al-Sabahi, J.N., Khan, A.A., Naz, S., Ijaz, A. 2016.

Production of biodiesel from low priced, renewable and abundant date seed oil.

Renewable Energy, 86, 124-132.

Azevedo, C., Corral, L. 1997. Some ultrastructural observations of a thraustochytrid

(Protoctista, Labyrinthulomycota) from the clam Ruditapes decussatus

(Mollusca, Bivalvia). Diseases of Aquatic Organisms, 31(1), 73-78.

B. Chabrol, R.C. 1937. Une nouvele reaction pour l’etudee des lipides: L’oleidemie.

Press med, 45, 1713.

Baganz, F., Xu, Y., Hellier, P., Ladommatos, N., Purton, S. 2014. Exploring the

potential of microalgae for bioenergy production. New Biotechnology, 31,

Supplement, S25.

138

Page 160: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Balaji, V., Ebenezer, P. 2008. Optimization of extracellular lipase production in

Colletotrichum gloeosporioides by solid state fermentation. Indian Journal of

Science and Technology, 1(7), 1-8.

Balasundaram, B., Pandit, A.B. 2001. Significance of location of enzymes on their

release during microbial cell disruption. Biotechnology and Bioengineering,

75(5), 607-614.

Balasundaram, B., Skill, S.C., Llewellyn, C.A. 2012. A low energy process for the

recovery of bioproducts from cyanobacteria using a ball mill. Biochemical

Engineering Journal, 69, 48-56.

Baldwin, C., Moo-Young, M. 1991. Disruption of a filamentous fungal organism (N.

sitophila) using a bead mill of novel design II. Increased recovery of cellulases.

Biotechnology Techniques, 5(5), 337-342.

Barrow, C.J. 2010. Marine nutraceuticals: Glucosamine and omega-3 fatty acids: New

trends for established ingredients. Agro Food Industry Hi-Tech, 21(2), 36-41.

Baunillo, K.E., Tan, R.S., Barros, H.R., Luque, R. 2012. Investigations on microalgal

oil production from Arthrospira platensis: Towards more sustainable biodiesel

production. RSC Advances, 2(30), 11267-11272.

Becerra, M., Rodríguez-Belmonte, E., Cerdán, M.E., González Siso, M.I. 2001.

Extraction of Intracellular Proteins from Kluyveromyces lactis. Food

Technology and Biotechnology, 39(2), 135-139.

Bellou, S., Baeshen, M.N., Elazzazy, A.M., Aggeli, D., Sayegh, F., Aggelis, G. 2014.

Microalgal lipids biochemistry and biotechnological perspectives.

Biotechnology Advances, 32(8), 1476-1493.

Bendikiene, V., Kiriliauskaite, V., Juodka, B. 2011. Production of environmentally

friendly biodiesel by enzymatic oil transesterification. Journal of

Environmental Engineering and Landscape Management, 19(2), 123-129.

Bertozzini, E., Galluzzi, L., Penna, A., Magnani, M. 2011. Application of the standard

addition method for the absolute quantification of neutral lipids in microalgae

using Nile red. Journal of Microbiological Methods, 87(1), 17-23.

Bhangle, S., Kolasinski, S.L. 2011. Fish oil in rheumatic diseases. Rheumatic Disease

Clinics of North America, 37(1), 77-84.

Birch, E.E., Garfield, S., Castañeda, Y., Hughbanks-Wheaton, D., Uauy, R., Hoffman,

D. 2007. Visual acuity and cognitive outcomes at 4 years of age in a double-

139

Page 161: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References

blind, randomized trial of long-chain polyunsaturated fatty acid-supplemented

infant formula. Early Human Development, 83(5), 279-284.

Bligh, E.G., Dyer, W.J. 1959. A rapid method of total lipid extraction and purification.

Canadian Journal of Biochemistry and Physiology, 37(8), 911-917.

Blondeau, N. 2015. The nutraceutical potential of omega-3 alpha-linolenic acid in

reducing the consequences of stroke. Biochimie.

BoekeMa, B.K.H.L., Beselin, A., Breuer, M., Hauer, B., Koster, M., Rosenau, F.,

Jaeger, K.-E., Tommassen, J. 2007. Hexadecane and tween 80 stimulate lipase

production in Burkholdetia glumae by different mechanisms. Applied and

Environmental Microbiology, 73(12), 3838-3844.

Bongiorni, L., Pusceddu, A., Danovaro, R. 2005. Enzymatic activities of epiphytic and

benthic thraustochytrids involved in organic matter degradation. Aquatic

Microbial Ecology, 41(3), 299-305.

Borkar, P.S., Bodade, R.G., Rao, S.R., Khobragade, C.N. 2009. Purification and

characterization of extracellular lipase from a new strain: Pseudomonas

aeruginosa SRT 9. Brazilian Journal of Microbiology, 40(2), 358-366.

Bradford, M.M. 1976b. A rapid and sensitive method for the quantitation of microgram

quantities of protein utilizing the principle of protein-dye binding. Analytical

Biochemistry, 72(1–2), 248-254.

Breivik, H. 2007. 'Concentrates', in H Breivik (ed.). Long-chain omega-3 specialty

oils, 21, 114.

Bremer, G. 2000. Isolation and culture of Thraustochytrids. Marine Mycology – A

Practical Approach (Hyde KD & Pointing SB, eds), pp. 49–61. Fungal

Diversity Press, Hong Kong.

Brennan, L., Owende, P. 2010. Biofuels from microalgae-A review of technologies for

production, processing, and extractions of biofuels and co-products.

Renewable and Sustainable Energy Reviews, 14(2), 557-577.

Browning, L.M., Walker, C.G., Mander, A.P., West, A.L., Madden, J., Gambell, J.M.,

Young, S., Wang, L., Jebb, S.A., Calder, P.C. 2012. Incorporation of

eicosapentaenoic and docosahexaenoic acids into lipid pools when given as

supplements providing doses equivalent to typical intakes of oily fish. The

American Journal of Clinical Nutrition, 96(4), 748-758.

140

Page 162: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Bull, A.T., Ward, A.C., Goodfellow, M. 2000. Search and discovery strategies for

biotechnology: the paradigm shift. Microbiology and Molecular Biology

Reviews, 64(3), 573-606.

Buono, S., Langellotti, A.L., Martello, A., Rinna, F., Fogliano, V. 2014. Functional

ingredients from microalgae. Food & Function, 5(8), 1669-1685.

Burdge, G.C., Calder, P.C. 2005. Conversion of alpha-linolenic acid to longer-chain

polyunsaturated fatty acids in human adults. Reprod. Nutr. Dev., 45(5), 581-

597.

Burja, A., Radianingtyas, H., Windust, A., Barrow, C. 2006. Isolation and

characterization of polyunsaturated fatty acid producing Thraustochytrium

species: screening of strains and optimization of omega-3 production. Applied

Microbiology and Biotechnology, 72(6), 1161-1169.

Burr, G.O., Burr, M.M. 1929. A new deficiency disease produced by the rigid

exclusion of fat from the diet. Journal of Biological Chemistry, 82(2), 345-367.

Byreddy, A., Gupta, A., Barrow, C., Puri, M. 2015. Comparison of cell disruption

methods for improving lipid extraction from thraustochytrid strains. Marine

Drugs, 13(8), 5111.

Cabo, J., Alonso, R., Mata, P. 2012. Omega-3 fatty acids and blood pressure. British

Journal of Nutrition, 107(SupplementS2), S195-S200.

Calder, P.C. 2012. The role of marine omega-3 (n-3) fatty acids in inflammatory

processes, atherosclerosis and plaque stability. Molecular Nutrition & Food

Research, 56(7), 1073-1080.

Cao, H., Zhang, Z., Wu, X., Miao, X. 2013. Direct biodiesel production from wet

microalgae biomass of Chlorella pyrenoidosa through in situ

transesterification. BioMed Research International, 6.

Cao, Y., Liu, W., Xu, X., Zhang, H., Wang, J., Xian, M. 2014. Production of free

monounsaturated fatty acids by metabolically engineered Escherichia coli.

Biotechnology for Biofuels, 7(1).

Cardenas, F., Alvarez, E., de Castro-Alvarez, M.-S., Sanchez-Montero, J.-M.,

Valmaseda, M., Elson, S.W., Sinisterra, J.-V. 2001. Screening and catalytic

activity in organic synthesis of novel fungal and yeast lipases. Journal of

Molecular Catalysis B: Enzymatic, 14(4–6), 111-123.

Carrière, F., Withers-Martinez, C., van Tilbeurgh, H., Roussel, A., Cambillau, C.,

Verger, R. 1998. Structural basis for the substrate selectivity of pancreatic

141

Page 163: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References

lipases and some related proteins. Biochimica et Biophysica Acta (BBA) -

Reviews on Biomembranes, 1376(3), 417-432.

Carvalho, P.O., Campos, P.R.B., D'Addio Noffs, M., Fregolente, P.B.L., Fregolente,

L.V. 2009. Enzymatic hydrolysis of salmon oil by native lipases: optimization

of process parameters. Journal of the Brazilian Chemical Society, 20(1), 117-

124.

Casas-Godoy, L., Meunchan, M., Cot, M., Duquesne, S., Bordes, F., Marty, A. 2014.

Yarrowia lipolytica lipase Lip2: An efficient enzyme for the production of

concentrates of docosahexaenoic acid ethyl ester. Journal of Biotechnology,

180, 30-36.

Catal, Xe1, , A. 2013. Five decades with polyunsaturated fatty acids: chemical

synthesis, enzymatic formation, lipid peroxidation and its biological effects.

Journal of Lipids, 2013, 19.

Challagulla, V., Fabbro, L., Nayar, S. 2015. Biomass, lipid productivity and fatty acid

composition of fresh water microalga Rhopalosolen saccatus cultivated under

phosphorous limited conditions. Algal Research, 8, 69-75.

Charbol, B., Charonnat, R. 1937. Une nouvele reaction pour l'etudee des lipides:

L'oleidemie. Presse Med, 45, 1713.

Chen, C.-L., Huang, C.-C., Ho, K.-C., Hsiao, P.-X., Wu, M.-S., Chang, J.-S. 2015a.

Biodiesel production from wet microalgae feedstock using sequential wet

extraction/transesterification and direct transesterification processes.

Bioresource Technology, 194, 179-186.

Chen, G., Zhao, L., Qi, Y. 2015b. Enhancing the productivity of microalgae cultivated

in wastewater toward biofuel production: A critical review. Applied Energy,

137, 282-291.

Chen, W., Ma, L., Zhou, P.-p., Zhu, Y.-m., Wang, X.-p., Luo, X.-a., Bao, Z.-d., Yu,

L.-j. 2015c. A novel feedstock for biodiesel production: The application of

palmitic acid from Schizochytrium. Energy, 86, 128-138.

Chen, W., Sommerfeld, M., Hu, Q. 2011. Microwave-assisted Nile red method for in

vivo quantification of neutral lipids in microalgae. Bioresource Technology,

102(1), 135-141.

Chen, W., Zhang, C., Song, L., Sommerfeld, M., Hu, Q. 2009. A high throughput Nile

red method for quantitative measurement of neutral lipids in microalgae.

Journal of Microbiological Methods, 77(1), 41-47.

142

Page 164: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Chen, Y., Vaidyanathan, S. 2012. A simple, reproducible and sensitive

spectrophotometric method to estimate microalgal lipids. Analytica Chimica

Acta, 724, 67-72.

Cheng, C.-H., Du, T.-B., Pi, H.-C., Jang, S.-M., Lin, Y.-H., Lee, H.-T. 2011a.

Comparative study of lipid extraction from microalgae by organic solvent and

supercritical CO2. Bioresource Technology, 102(21), 10151-10153.

Cheng, Y.-S., Zheng, Y., VanderGheynst, J. 2011b. Rapid quantitative analysis of

lipids using a colorimetric method in a microplate format. Lipids, 46(1), 95-

103.

Chisti, Y. 2007. Biodiesel from microalgae. Biotechnology Advances, 25(3), 294-306.

Chisti, Y., Moo-Young, M. 1986. Disruption of microbial cells for intracellular

products. Enzyme and Microbial Technology, 8(4), 194-204.

Cho, S.-C., Choi, W.-Y., Oh, S.-H., Lee, C.-G., Seo, Y.-C., Kim, J.-S., Song, C.-H.,

Kim, G.-V., Lee, S.-Y., Kang, D.-H., Lee, H.-Y. 2012b. Enhancement of lipid

extraction from marine microalga, scenedesmus associated with high-pressure

homogenization process. Journal of Biomedicine and Biotechnology, 6

Choi, S.P., Nguyen, M.T., Sim, S.J. 2010. Enzymatic pretreatment of Chlamydomonas

reinhardtii biomass for ethanol production. Bioresource Technology, 101(14),

5330-5336.

Choonia, H.S., Lele, S.S. 2013. -galactosidase by permeabilization of

indigenously isolated Lactobacillus acidophilus using lysozyme. Chemical

and Biochemical Engineering Quarterly, 27(4), 449-456.

Christie, W.W., Han, X. 2010. Lipid Analysis: Isolation, Separation, Identification and

Lipidomic Analysis: Fourth Edition.

Cicero, A.F., Morbini, M., Borghi, C. 2015. Do we need ‘new’ omega-3

polyunsaturated fatty acids formulations? Expert Opinion on

Pharmacotherapy, 16(3), 285-288.

Cihangir, N., Sarikaya, E. 2004. Investigation of lipase production by a new isolate of

Aspergillus sp. World Journal of Microbiology and Biotechnology, 20(2), 193-

197.

Cockbain, A.J., Toogood, G.J., Hull, M.A. 2012. Omega-3 polyunsaturated fatty acids

for the treatment and prevention of colorectal cancer. Gut, 61(1), 135-149.

Daroch, M., Geng, S., Wang, G. 2013. Recent advances in liquid biofuel production

from algal feedstocks. Applied Energy, 102, 1371-1381.

143

Page 165: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Dartois, V., Baulard, A., Schanck, K., Colson, C. 1992. Cloning, nucleotide sequence

and expression in Escherichia coli of a lipase gene from Bacillus subtilis 168.

Biochimica et Biophysica Acta (BBA) - Gene Structure and Expression,

1131(3), 253-260.

De Caterina, R. 2011. n-3 Fatty acids in cardiovascular disease. New England Journal

of Medicine, 364(25), 2439-2450.

Dejoye Tanzi, C., Abert Vian, M., Chemat, F. 2013. New procedure for extraction of

algal lipids from wet biomass: A green clean and scalable process. Bioresource

Technology, 134, 271-275.

Delarue, J., Guriec, N. 2014. Opportunities to enhance alternative sources of long-

chain n-3 fatty acids within the diet. Proceedings of the Nutrition Society,

73(03), 376-384.

Delrue, F., Setier, P.A., Sahut, C., Cournac, L., Roubaud, A., Peltier, G., Froment,

A.K. 2012. An economic, sustainability, and energetic model of biodiesel

production from microalgae. Bioresource Technology, 111, 191-200.

Di Caprio, F., Altimari, P., Pagnanelli, F. 2015. Integrated biomass production and

biodegradation of olive mill wastewater by cultivation of Scenedesmus sp.

Algal Research, 9, 306-311.

dos Santos, R.R., Moreira, D.M., Kunigami, C.N., Aranda, D.A.G., Teixeira, C.M.L.L.

2015. Comparison between several methods of total lipid extraction from

Chlorella vulgaris biomass. Ultrasonics Sonochemistry, 22, 95-99.

Doucha, J., Lívanský, K. 2008. Influence of processing parameters on disintegration

of Chlorella cells in various types of homogenizers. Applied Microbiology and

Biotechnology, 81(3), 431-440.

Doulah, M. 1977. Mechanism of disintegration of biological cells in ultrasonic

cavitation. Biotechnol Bioeng, 19(Suppl 5), 649 - 660.

Du, Y., Schuur, B., Kersten, S.R.A., Brilman, D.W.F. 2015. Opportunities for

switchable solvents for lipid extraction from wet algal biomass: An energy

evaluation. Algal Research, 11, 271-283.

Dyerberg, J., Madsen, P., Møller, J.M., Aardestrup, I., Schmidt, E.B. 2010.

Bioavailability of marine n-3 fatty acid formulations. Prostaglandins,

Leukotrienes and Essential Fatty Acids, 83(3), 137-141.

144

Page 166: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Ellulu, M.S., Khaza’ai, H., Abed, Y., Rahmat, A., Ismail, P., Ranneh, Y. 2015. Role

of fish oil in human health and possible mechanism to reduce the inflammation.

Inflammopharmacology, 23(2-3), 79-89.

-

2014. Carbon source impact on Yarrowia lipolytica KKP 379 lipase

production. Applied Biochemistry and Microbiology, 50(4), 404-410.

Fajardo, A.R., Cerdán, L.E., Medina, A.R., Fernández, F.G.A., Moreno, P.A.G.,

Grima, E.M. 2007. Lipid extraction from the microalga Phaeodactylum

tricornutum. European Journal of Lipid Science and Technology, 109(2), 120-

126.

Fan, K.W., Chen, F. 2007. Production of high-value products by marine microalgae

thraustochytrids. Bioprocessing for Value-Added Products from Renewable

Resources: New Technologies and Applications, 293-323.

Farkade, V.D., Harrison, S.T.L., Pandit, A.B. 2006. Improved cavitational cell

-galactosidase from

Kluveromyces lactis. Biochemical Engineering Journal, 31(1), 25-30.

Fernández-Lorente, G., Betancor, L., Carrascosa, A., Guisán, J. 2011. Release of

omega-3 fatty acids by the hydrolysis of fish oil catalyzed by lipases

immobilized on hydrophobic supports. Journal of the American Oil Chemists'

Society, 88(8), 1173-1178.

Fernandez-Lorente, G., Filice, M., Terreni, M., Guisan, J.M., Fernandez-Lafuente, R.,

Palomo, J.M. 2008. Lecitase® ultra as regioselective biocatalyst in the

hydrolysis of fully protected carbohydrates: Strong modulation by using

different immobilization protocols. Journal of Molecular Catalysis B:

Enzymatic, 51(3–4), 110-117.

Ferreira-Dias, S., Sandoval, G., Plou, F., Valero, F. 2013. The potential use of lipases

in the production of fatty acid derivatives for the food and nutraceutical

industries. Electronic Journal of Biotechnology, 16(3).

Florentino de Souza Silva, A.P., Costa, M.C., Colzi Lopes, A., Fares Abdala Neto, E.,

Carrhá Leitão, R., Mota, C.R., Bezerra dos Santos, A. 2014. Comparison of

pretreatment methods for total lipids extraction from mixed microalgae.

Renewable Energy, 63, 762-766.

Folch, J., Ascoli, I., Lees, M., Meath, J.A., Lebaron, F.N. 1951. Preparation of lipide

extracts from brain tissue. Journal of Biological Chemistry, 191(2), 833-841.

145

Page 167: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Folch, J., Lees, M., Stanley, G.H.S. 1957. A simple method for the isolation and

purification of total lipides from animal tissues. Journal of Biological

Chemistry, 226(1), 497-509.

Frings, C.S., Dunn, R.T. 1970. A colorimetric method for determination of total serum

lipids based on the sulfo-phospho-vanillin reaction. American Journal of

Clinical Pathology, 53(1), 89-91.

Fu, C.C., Hung, T.C., Chen, J.Y., Su, C.H., Wu, W.T. 2010. Hydrolysis of microalgae

cell walls for production of reducing sugar and lipid extraction. Bioresource

Technology, 101(22), 8750-8754.

Gámez-Meza, N., Noriega- -Juárez, L.A., Ortega-

Monroy-Rivera, J., Toro- -Guerrero, O.

2003. Concentration of eicosapentaenoic acid and docosahexaenoic acid from

fish oil by hydrolysis and urea complexation. Food Research International,

36(7), 721-727.

Gao, L.-g., Cao, J., Mao, Q.-x., Lu, X.-c., Zhou, X.-l., Fan, L. 2013. Influence of

omega-3 polyunsaturated fatty acid-supplementation on platelet aggregation in

humans: A meta-analysis of randomized controlled trials. Atherosclerosis,

226(2), 328-334.

Gebauer, S.K., Psota, T.L., Harris, W.S., Kris-Etherton, P.M. 2006. n-3 Fatty acid

dietary recommendations and food sources to achieve essentiality and

cardiovascular benefits. The American Journal of Clinical Nutrition, 83(6),

S1526-1535S.

Gerde, J.A., Montalbo-Lomboy, M., Yao, L., Grewell, D., Wang, T. 2012. Evaluation

of microalgae cell disruption by ultrasonic treatment. Bioresource Technology,

125, 175-181.

Gogate, P.R., Pandit, A.B. 2008. Application of cavitational reactors for cell disruption

for recovery of intracellular enzymes. Journal of Chemical Technology and

Biotechnology, 83(8), 1083-1093.

Gopinath, S.C.B., Anbu, P., Lakshmipriya, T., Hilda, A. 2013. Strategies to

characterize fungal lipases for applications in medicine and dairy industry.

BioMed Research International, 2013, 154549.

Gotor-Fernández, V., Brieva, R., Gotor, V. 2006. Lipases: Useful biocatalysts for the

preparation of pharmaceuticals. Journal of Molecular Catalysis B: Enzymatic,

40(3–4), 111-120.

146

Page 168: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Govender, T., Ramanna, L., Rawat, I., Bux, F. 2012. BODIPY staining, an alternative

to the Nile Red fluorescence method for the evaluation of intracellular lipids in

microalgae. Bioresource Technology, 114, 507-511.

Griffiths, M.J., Harrison, S.T.L. 2009. Lipid productivity as a key characteristic for

choosing algal species for biodiesel production. Journal of Applied Phycology,

21(5), 493-507.

Grima, E.M., Medina, A.R., Giménez, A.G., Sánchez Pérez, J.A., Camacho, F.G.,

García Sánchez, J.L. 1994. Comparison between extraction of lipids and fatty

acids from microalgal biomass. Journal of the American Oil Chemists’ Society,

71(9), 955-959.

Grimi, N., Dubois, A., Marchal, L., Jubeau, S., Lebovka, N.I., Vorobiev, E. 2014.

Selective extraction from microalgae Nannochloropsis sp. using different

methods of cell disruption. Bioresource Technology, 153, 254-259.

Guckert, J.B., Cooksey, K.E., Jackson, L.L. 1988. Lipid solvent systems are not

equivalent for analysis of lipid classes in the microeukaryotic green alga,

Chlorella. Journal of Microbiological Methods, 8(3), 139-149.

Guil-Guerrero, J.L., Belarbi, E.-H. 2001. Purification process for cod liver oil

polyunsaturated fatty acids. Journal of the American Oil Chemists' Society,

78(5), 477-484.

Guldhe, A., Singh, B., Rawat, I., Ramluckan, K., Bux, F. 2014. Efficacy of drying and

cell disruption techniques on lipid recovery from microalgae for biodiesel

production. Fuel, 128, 46-52.

Günerken, E., D'Hondt, E., Eppink, M.H.M., Garcia-Gonzalez, L., Elst, K., Wijffels,

R.H. 2015. Cell disruption for microalgae biorefineries. Biotechnology

Advances, 33(2), 243-260.

Gupta, A., Abraham, R.E., Barrow, C.J., Puri, M. 2015. Omega-3 fatty acid production

from enzyme saccharified hemp hydrolysate using a novel marine

thraustochytrid strain. Bioresource Technology, 184, 373-378.

Gupta, A., Barrow, C.J., Puri, M. 2012. Omega-3 biotechnology: Thraustochytrids as

a novel source of omega-3 oils. Biotechnology Advances, 30(6), 1733-1745.

Gupta, A., Singh, D., Barrow, C.J., Puri, M. 2013. Exploring potential use of

Australian thraustochytrids for the bioconversion of glycerol to omega-3 and

carotenoids production. Biochemical Engineering Journal, 78, 11-17.

147

Page 169: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Gupta, A., Vongsvivut, J., Barrow, C.J., Puri, M. 2012a. Molecular identification of

marine yeast and its spectroscopic analysis establishes unsaturated fatty acid

accumulation. Journal of Bioscience and Bioengineering, 114(4), 411-417.

Gupta, N., Sahai, V., Gupta, R. 2007. Alkaline lipase from a novel strain Burkholderia

multivorans: Statistical medium optimization and production in a bioreactor.

Process Biochemistry, 42(4), 518-526.

Gupta, R., Gupta, N., Rathi, P. 2004. Bacterial lipases: an overview of production,

purification and biochemical properties. Applied Microbiology and

Biotechnology, 64(6), 763-781.

Guschina, I.A., Harwood, J.L. 2006. Lipids and lipid metabolism in eukaryotic algae.

Progress in Lipid Research, 45(2), 160-186.

Halim, R., Danquah, M.K., Webley, P.A. 2012a. Extraction of oil from microalgae for

biodiesel production: A review. Biotechnology Advances, 30(3), 709-732.

Halim, R., Gladman, B., Danquah, M.K., Webley, P.A. 2011. Oil extraction from

microalgae for biodiesel production. Bioresource Technology, 102(1), 178-

185.

Halim, R., Harun, R., Danquah, M.K., Webley, P.A. 2012b. Microalgal cell disruption

for biofuel development. Applied Energy, 91(1), 116-121.

Halim, R., Rupasinghe, T.W.T., Tull, D.L., Webley, P.A. 2013. Mechanical cell

disruption for lipid extraction from microalgal biomass. Bioresource

Technology, 140, 53-63.

Halldorsson, A., Kristinsson, B., Haraldsson, G.G. 2004. Lipase selectivity toward

fatty acids commonly found in fish oil. European Journal of Lipid Science and

Technology, 106(2), 79-87.

Han, W.W.C.a.X. 2010. Lipid analysis: Lipid isolation, separation, identification and

lipidomic analysis The Oily press, Bridgewater, England.

Han, Y., Wen, Q., Chen, Z., Li, P. 2011. Review of methods used for microalgal lipid-

content analysis. Energy Procedia, 12, 944-950.

Hao, G., Yang, L., Mazsaroff, I., Lin, M. 2007. Quantitative determination of lipase

activity by liquid chromatography-mass spectrometry. Journal of the American

Society for Mass Spectrometry, 18(9), 1579-1581.

Haraldsson, G.G. 1999. The application of lipase for preparing various lipids enriched

with omega-3 fatty acids. Rit Fiskideildar, 16, 97-105.

148

Page 170: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Harun, R., Danquah, M.K. 2011. Enzymatic hydrolysis of microalgal biomass for

bioethanol production. Chemical Engineering Journal, 168(3), 1079-1084.

Hasan, F., Shah, A.A., Hameed, A. 2006. Industrial applications of microbial lipases.

Enzyme and Microbial Technology, 39(2), 235-251.

Haskins, S.D., Kelly, D.G., Weir, R.D. 2010. Novel pressurized solvent extraction

vessels for the analysis of polychlorinated biphenyl congeners in avian whole

blood. Analytica Chimica Acta, 677(1), 19-23.

Hoekman, S.K., Broch, A., Robbins, C., Ceniceros, E., Natarajan, M. 2012. Review of

biodiesel composition, properties, and specifications. Renewable and

Sustainable Energy Reviews, 16(1), 143-169.

Hoen, P.W., Denollet, J., De Jonge, P., Whooley, M.A. 2013. Positive affect and

survival in patients with stable coronary heart disease: Findings from the heart

and soul study. Journal of Clinical Psychiatry, 74(7), 716-722.

Honda, D., Yokochi, T., Nakahara, T., Raghukumar, S., Nakagiri, A., Schaumann, K.,

Higashihara, T. 1999. Molecular phylogeny of labyrinthulids and

thraustochytrids based on the sequencing of 18S ribosomal RNA gene. Journal

of Eukaryotic Microbiology, 46(6), 637-647.

Huang, Y., Hong, A., Zhang, D., Li, L. 2014. Comparison of cell rupturing by

ozonation and ultrasonication for algal lipid extraction from Chlorella

vulgaris. Environmental Technology (United Kingdom), 35(8), 931-937.

Huo, Y.-X., Cho, K.M., Rivera, J.G.L., Monte, E., Shen, C.R., Yan, Y., Liao, J.C.

2011. Conversion of proteins into biofuels by engineering nitrogen flux.

Nature Biotechnology, 29(4), 346-351.

Iftikhar, T., Niaz, M., Hussain, Y., Abbas, S.Q., Ashraf, I., Zia, M.A. 2010.

Improvement of selected strains through gamma irradiation for enhanced

lipolytic potential. Pakistan Journal of Botany, 42(4), 2257-2267.

Iida, Y., Tuziuti, T., Yasui, K., Kozuka, T., Towata, A. 2008. Protein release from

yeast cells as an evaluation method of physical effects in ultrasonic field.

Ultrasonics Sonochemistry, 15(6), 995-1000.

Innis, S.M. 2007. Dietary (n-3) Fatty acids and brain development. The Journal of

Nutrition, 137(4), 855-859.

Islam, M.A., Magnusson, M., Brown, R.J., Ayoko, G.A., Nabi, M.N., Heimann, K.

2013. Microalgal species selection for biodiesel production based on fuel

properties derived from fatty acid profiles. Energies, 6(11), 5676-5702.

149

Page 171: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Izard, J., Limberger, R.J. 2003. Rapid screening method for quantitation of bacterial

cell lipids from whole cells. Journal of Microbiological Methods, 55(2), 411-

418.

Jahanshahi, M., Sun, Y., Santos, E., Pacek, A., Franco, T.T., Nienow, A., Lyddiatt, A.

2002. Operational intensification by direct product sequestration from cell

disruptates: Application of a pellicular adsorbent in a mechanically integrated

disruption-fluidised bed adsorption process. Biotechnology and

Bioengineering, 80(2), 201-212.

Jakobsen, A., Aasen, I., Josefsen, K., Strøm, A. 2008. Accumulation of

docosahexaenoic acid-rich lipid in thraustochytrid Aurantiochytrium sp. strain

T66: effects of N and P starvation and O2 limitation. Applied Microbiology and

Biotechnology, 80(2), 297-306.

Jalalirad, R. 2013. Selective and efficient extraction of recombinant proteins from the

periplasm of Escherichia coli using low concentrations of chemicals. Journal

of Industrial Microbiology and Biotechnology, 40(10), 1117-1129.

Jeon, J.-m., Choi, H.-W., Yoo, G.-C., Choi, Y.-K., Choi, K.-Y., Park, H.-Y., Park, S.-

H., Kim, Y.-G., Kim, H.J., Lee, S.H., Lee, Y.K., Yang, Y.-H. 2013. New

mixture composition of organic solvents for efficient extraction of lipids from

Chlorella vulgaris. Biomass and Bioenergy, 59, 279-284.

Jermsuntiea, W., Aki, T., Toyoura, R., Iwashita, K., Kawamoto, S., Ono, K. 2011.

Purification and characterization of intracellular lipase from the

polyunsaturated fatty acid-producing fungus Mortierella alliacea. New

Biotechnology, 28(2), 158-164.

Johnson, M.B., Wen, Z. 2009. Production of biodiesel fuel from the microalga

Schizochytrium limacinum by direct transesterification of algal biomass.

Energy and Fuels, 23(10), 5179-5183.

Kadkhodaee, R., Povey, M.J.W. 2008. Ultrasonic inactivation of Bacillus -amylase.

I. effect of gas content and emitting face of probe. Ultrasonics Sonochemistry,

15(2), 133-142.

Kahveci, D., Falkeborg, M., Gregersen, S., Xu, X. 2010. Upgrading of farmed salmon

oil through lipase-catalyzed hydrolysis. The Open Biotechnology Journal, 4,

47-55.

150

Page 172: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Kahveci, D., Xu, X. 2011. Repeated hydrolysis process is effective for enrichment of

omega 3 polyunsaturated fatty acids in salmon oil by Candida rugosa lipase.

Food Chemistry, 129(4), 1552-1558.

Kale, A. 2013. Method of extracting neutral lipids with two solvents, pp. US patent

WO 2013142694 A1.

Kamal, M.Z., Barrow, C.J., Rao, N.M. 2015. A computational search for lipases that

can preferentially hydrolyze long-chain omega-3 fatty acids from fish oil

triacylglycerols. Food Chemistry, 173, 1030-1036.

Kanchana, R., Muraleedharan, U., Raghukumar, S. 2011. Alkaline lipase activity from

the marine protists, thraustochytrids. World Journal of Microbiology and

Biotechnology, 27(9), 2125-2131.

Kapturowska, A.U., Stolarzewicz, I.A., Krzyczko -

2012. Studies on the lipolytic activity of sonicated enzymes from Yarrowia

lipolytica. Ultrasonics Sonochemistry, 19(1), 186-191.

Kar, J.R., Singhal, R.S. 2015. Investigations on ideal mode of cell disruption in

extremely halophilic Actinopolyspora halophila (MTCC 263) for efficient

release of glycine betaine and trehalose. Biotechnology Reports, 5, 89-97.

Karatay, S.E., Dönmez, G. 2010. Improving the lipid accumulation properties of the

yeast cells for biodiesel production using molasses. Bioresource Technology,

101(20), 7988-7990.

Kennedy, J., Marchesi, J., Dobson, A. 2008. Marine metagenomics: strategies for the

discovery of novel enzymes with biotechnological applications from marine

environments. Microbial Cell Factories, 7(1), 27.

Kim, J., Yoo, G., Lee, H., Lim, J., Kim, K., Kim, C.W., Park, M.S., Yang, J.-W. 2013.

Methods of downstream processing for the production of biodiesel from

microalgae. Biotechnology Advances, 31(6), 862-876.

Kim, T.-H., Suh, W.I., Yoo, G., Mishra, S.K., Farooq, W., Moon, M., Shrivastav, A.,

Park, M.S., Yang, J.-W. 2015. Development of direct conversion method for

microalgal biodiesel production using wet biomass of Nannochloropsis salina.

Bioresource Technology, 191, 438-444.

Klimek-Ochab, M., Brzezinska-Rodak, M., Zymanczyk-Duda, E., Lejczak, B.,

Kafarski, P. 2011. Comparative study of fungal cell disruption-scope and

limitations of the methods. Folia Microbiol, 56, 469 - 475.

151

Page 173: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Knight, J.A., Anderson, S., Rawle, J.M. 1972. Chemical basis of the sulfo-phospho-

vanillin reaction for estimating total serum lipids. Clinical Chemistry, 18(3),

199-202.

Kosugi, Y., Azuma, N. 1994. Synthesis of triacylglycerol from polyunsaturated fatty

acid by immobilized lipase. Journal of the American Oil Chemists’ Society,

71(12), 1397-1403.

Kou, Z., Bei, S., Sun, J., Pan, J. 2013. Fluorescent measurement of lipid content in the

model organism Chlamydomonas reinhardtii. Journal of Applied Phycology,

25(6), 1633-1641.

Kralovec, J.A., Wang, W., Barrow, C.J. 2010. Production of omega-3 triacylglycerol

concentrates using a new food grade immobilized Candida antarctica lipase

B. Australian Journal of Chemistry, 63(6), 922-928.

Kralovec, J.A., Zhang, S., Zhang, W., Barrow, C.J. 2012. A review of the progress in

enzymatic concentration and microencapsulation of omega-3 rich oil from fish

and microbial sources. Food Chemistry, 131(2), 639-644.

Kumar, M., Sharma, M.P. 2015. Assessment of potential of oils for biodiesel

production. Renewable and Sustainable Energy Reviews, 44, 814-823.

Kumar, P.S., Pandit, A.B. 1999. Modeling hydrodynamic cavitation. Chemical

Engineering and Technology, 22(12), 1017-1027.

Kuratko, C., Abril, J.R., Hoffman, J.P., Salem Jr, N. 2013. 13 - Enrichment of infant

formula with omega-3 fatty acids. in: Food Enrichment with Omega-3 Fatty

Acids, (Eds.) C. Jacobsen, N.S. Nielsen, A.F. Horn, A.-D.M. Sørensen,

Woodhead Publishing, pp. 353-386.

Kuratko, C.N., Salem, N. 2013. Docosahexaenoic acid from algal oil. European

Journal of Lipid Science and Technology, 115(9), 965-976.

Kurtovic, I., Marshall, S.N. 2013. Potential of fish by-products as a source of novel

marine lipases and their uses in industrial applications. Lipid Technology,

25(2), 35-37.

Lam, K.S. 2006. Discovery of novel metabolites from marine actinomycetes. Current

Opinion in Microbiology, 9(3), 245-251.

Lavie, C.J., Milani, R.V., Mehra, M.R., Ventura, H.O. 2009. Omega-3 polyunsaturated

fatty acids and cardiovascular diseases. Journal of the American College of

Cardiology, 54(7), 585-594.

152

Page 174: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Leano, E.M., Gapasin, R.S.J., Polohan, B., Vrijmoed, L.L.P. 2003. Growth and fatty

acid production of thraustochytrids from Panay mangroves, Philippines.

Fungal Diversity, 12, 111-122.

Lee, A.K., Lewis, D.M., Ashman, P.J. 2012. Disruption of microalgal cells for the

extraction of lipids for biofuels: Processes and specific energy requirements.

Biomass and Bioenergy, 46, 89-101.

Lee Chang, K., Dumsday, G., Nichols, P., Dunstan, G., Blackburn, S., Koutoulis, A.

2013. High cell density cultivation of a novel Aurantiochytrium sp. strain TC

20 in a fed-batch system using glycerol to produce feedstock for biodiesel and

omega-3 oils. Applied Microbiology and Biotechnology, 97(15), 6907-6918.

Lee Chang, K., Nichols, C., Blackburn, S., Dunstan, G., Koutoulis, A., Nichols, P.

2014. Comparison of thraustochytrids Aurantiochytrium sp., Schizochytrium

sp., Thraustochytrium sp., and Ulkenia sp. for production of biodiesel, long-

chain omega-3 oils, and exopolysaccharide. Marine Biotechnology, 16(4), 396-

411.

Lee, H.S., Kwon, K.K., Kang, S.G., Cha, S.-S., Kim, S.-J., Lee, J.-H. 2010a.

Approaches for novel enzyme discovery from marine environments. Current

Opinion in Biotechnology, 21(3), 353-357.

Lee, I., Han, J.-I. 2015. Simultaneous treatment (cell disruption and lipid extraction)

of wet microalgae using hydrodynamic cavitation for enhancing the lipid yield.

Bioresource Technology, 186, 246-251.

Lee, J.-Y., Yoo, C., Jun, S.-Y., Ahn, C.-Y., Oh, H.-M. 2010b. Comparison of several

methods for effective lipid extraction from microalgae. Bioresource

Technology, 101(1, Supplement), S75-S77.

Lee, S., Yoon, B.-D., Oh, H.-M. 1998. Rapid method for the determination of lipid

from the green alga Botryococcus braunii. Biotechnology Techniques, 12(7),

553-556.

Lee Ying Yeng, A., Kadir, M.S., Ghazali, H., Raja Abd Rahman, R.N., Saari, N. 2013.

A comparative study of extraction techniques for maximum recovery of

glutamate decarboxylase (GAD) from Aspergillus oryzae NSK. BMC Research

Notes, 6(1), 526.

Leite, G.B., Paranjape, K., Abdelaziz, A.E.M., Hallenbeck, P.C. 2015. Utilization of

biodiesel-derived glycerol or xylose for increased growth and lipid production

by indigenous microalgae. Bioresource Technology, 184, 123-130.

153

Page 175: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Lenihan-Geels, G., Bishop, K.S., Ferguson, L.R. 2013. Alternative Sources of Omega-

3 Fats: Can We Find a Sustainable Substitute for Fish? Nutrients, 5(4), 1301-

1315.

Lewis, T., Nichols, P.D., McMeekin, T.A. 2000. Evaluation of extraction methods for

recovery of fatty acids from lipid-producing microheterotrophs. Journal of

Microbiological Methods, 43(2), 107-116.

Li, C.-Y., Cheng, C.-Y., Chen, T.-L. 2001. Production of Acinetobacter radioresistens

lipase using Tween 80 as the carbon source. Enzyme and Microbial

Technology, 29(4–5), 258-263.

Li, J., Liu, R., Chang, G., Li, X., Chang, M., Liu, Y., Jin, Q., Wang, X. 2015. A strategy

for the highly efficient production of docosahexaenoic acid by

Aurantiochytrium limacinum SR21 using glucose and glycerol as the mixed

carbon sources. Bioresource Technology, 177, 51-57.

Li, M., Yang, L.-R., Xu, G., Wu, J.-P. 2014a. Identification of organic solvent-tolerant

lipases from organic solvent-sensitive microorganisms. Journal of Molecular

Catalysis B: Enzymatic, 99, 96-101.

Li, X., Xu, H., Wu, Q. 2007. Large-scale biodiesel production from microalga

Chlorella protothecoides through heterotrophic cultivation in bioreactors.

Biotechnology and Bioengineering, 98(4), 764-771.

Li, Y., Ghasemi Naghdi, F., Garg, S., Adarme-Vega, T., Thurecht, K., Ghafor, W.,

Tannock, S., Schenk, P. 2014b. A comparative study: the impact of different

lipid extraction methods on current microalgal lipid research. Microbial Cell

Factories, 13(1), 14.

Liang, M.-H., Jiang, J.-G. 2013. Advancing oleaginous microorganisms to produce

lipid via metabolic engineering technology. Progress in Lipid Research, 52(4),

395-408.

Lie, Ø., Lambertsen, G. 1985. Digestive lipolytic enzymes in cod (Gadus morrhua):

Fatty acid specificity. Comparative Biochemistry and Physiology -- Part B:

Biochemistry and, 80(3), 447-450.

Lin, L., Cunshan, Z., Vittayapadung, S., Xiangqian, S., Mingdong, D. 2011.

Opportunities and challenges for biodiesel fuel. Applied Energy, 88(4), 1020-

1031.

Lin, Y.-S., Ginsberg, G., Lin, J.-W., Sonawane, B. 2014. Mercury exposure and

omega-3 fatty acid intake in relation to renal function in the US population.

154

Page 176: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References

International Journal of Hygiene and Environmental Health, 217(4–5), 465-

472.

Linder, M., Matouba, E., Fanni, J., Parmentier, M. 2002. Enrichment of salmon oil

with n-3 PUFA by lipolysis, filtration and enzymatic re-esterification.

European Journal of Lipid Science and Technology, 104(8), 455-462.

Liu, D., Zeng, X.-A., Sun, D.-W., Han, Z. 2013. Disruption and protein release by

ultrasonication of yeast cells. Innovative Food Science & Emerging

Technologies, 18, 132-137.

Liu, N., Wang, Y., Zhao, Q., Cui, C., Fu, M., Zhao, M. 2012. Immobilisation of

lecitase® ultra for production of diacylglycerols by glycerolysis of soybean oil.

Food Chemistry, 134(1), 301-307.

Liu, N., Wang, Y., Zhao, Q., Zhang, Q., Zhao, M. 2011. Fast synthesis of 1,3-DAG by

Lecitase® ultra-catalyzed esterification in solvent-free system. European

Journal of Lipid Science and Technology, 113(8), 973-979.

Liu, Y., Singh, P., Sun, Y., Luan, S., Wang, G. 2014a. Culturable diversity and

biochemical features of thraustochytrids from coastal waters of Southern

China. Applied Microbiology and Biotechnology, 98(7), 3241-3255.

Liu, Y., Tang, J., Li, J., Daroch, M., Cheng, J. 2014b. Efficient production of

triacylglycerols rich in docosahexaenoic acid (DHA) by osmo-heterotrophic

marine protists. Applied Microbiology and Biotechnology, 98(23), 9643-9652.

Lucia, B., Fernando, D. 2002. Distribution and abundance of thraustochytrids in

different Mediterranean coastal habitats. Aquatic Microbial Ecology, 30(1),

49-56.

Lyberg, A.-M., Adlercreutz, P. 2008. Lipase specificity towards eicosapentaenoic acid

and docosahexaenoic acid depends on substrate structure. Biochimica et

Biophysica Acta (BBA) - Proteins and Proteomics, 1784(2), 343-350.

Mahaffey, K.R., Sunderland, E.M., Chan, H.M., Choi, A.L., Grandjean, P., Mariën,

K., Oken, E., Sakamoto, M., Schoeny, R., Weihe, P., Yan, C.H., Yasutake, A.

2011. Balancing the benefits of n-3 polyunsaturated fatty acids and the risks of

methylmercury exposure from fish consumption. Nutrition Reviews, 69(9),

493-508.

Makrides, M., Collins, C.T., Gibson, R.A. 2011. Impact of fatty acid status on growth

and neurobehavioural development in humans. Maternal & Child Nutrition, 7,

80-88.

155

Page 177: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Mandal, S., Patnaik, R., Singh, A.K., Mallick, N. 2013. Comparative assessment of

various lipid extraction protocols and optimization of transesterification

process for microalgal biodiesel production. Environmental Technology,

34(13-14), 2009-2018.

Market watch, Feb 24, 2014 1:59 p.m

Martins, A.B., da Silva, A.M., Schein, M.F., Garcia-Galan, C., Záchia Ayub, M.A.,

Fernandez-Lafuente, R., Rodrigues, R.C. 2014. Comparison of the

performance of commercial immobilized lipases in the synthesis of different

flavor esters. Journal of Molecular Catalysis B: Enzymatic, 105, 18-25.

Martins, D.A., Custódio, L., Barreira, L., Pereira, H., Ben-Hamadou, R., Varela, J.,

Abu-Salah, K.M. 2013. Alternative sources of n-3 long-chain polyunsaturated

fatty acids in marine microalgae. Marine Drugs, 11(7), 2259-2281.

Mata, T.M., Martins, A.A., Caetano, N.S. 2010. Microalgae for biodiesel production

and other applications: A review. Renewable and Sustainable Energy Reviews,

14(1), 217-232.

Maurer, N.E., Hatta-Sakoda, B., Pascual-Chagman, G., Rodriguez-Saona, L.E. 2012.

Characterization and authentication of a novel vegetable source of omega-3

fatty acids, sacha inchi (Plukenetia volubilis L.) oil. Food Chemistry, 134(2),

1173-1180.

Maurich, V., Zacchigna, M., Pitotti, A. 1991. p-nitrophenyllaurate: a substrate for the

high-performance liquid chromatographic determination of lipase activity.

Journal of Chromatography B: Biomedical Sciences and Applications, 566(2),

453-459.

Mawson, R., Gamage, M., Terefe, N., Knoerzer, K. 2011. Ultrasound in enzyme

activation and inactivation. in: Ultrasound Technologies for Food and

Bioprocessing, (Eds.) H. Feng, G. Barbosa-Canovas, J. Weiss, Springer New

York, pp. 369-404.

Mazzucotelli, C.A., Agüero, M.V., Moreira, M.d.R., Ansorena, M.R. 2015.

Optimization of medium components and physicochemical parameters to

simultaneously enhance microbial growth and production of lypolitic enzymes

by Stenotrophomonas sp. Biotechnology and Applied Biochemistry, DOI:

10.1002/bab.1378.

McKenzie, H.A., White Jr, F -lactalbumin: Structure,

function, and interrelationships. Advances in Protein Chemistry, 41, 173-315.

156

Page 178: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References McMillan, J.R., Watson, I.A., Ali, M., Jaafar, W. 2013. Evaluation and comparison of

algal cell disruption methods: Microwave, waterbath, blender, ultrasonic and

laser treatment. Applied Energy, 103, 128-134.

McNeill, G., Ackman, R., Moore, S. 1996. Lipase-catalyzed enrichment of long-chain

polyunsaturated fatty acids. Journal of the American Oil Chemists’ Society,

73(11), 1403-1407.

Medina, A.R., Grima, E.M., Giménez, A.G., González, M.J.I. 1998. Downstream

processing of algal polyunsaturated fatty acids. Biotechnology Advances,

16(3), 517-580.

Meher, L.C., Vidya Sagar, D., Naik, S.N. 2006. Technical aspects of biodiesel

production by transesterification-a review. Renewable and Sustainable Energy

Reviews, 10(3), 248-268.

Mercer, P., Armenta, R.E. 2011. Developments in oil extraction from microalgae.

European Journal of Lipid Science and Technology, 113(5), 539-547.

Middelberg, A.P.J. 1995. Process-scale disruption of microorganisms. Biotechnology

Advances, 13(3), 491-551.

Milledge, J.J. 2012. Microalgae - Commercial potential for fuel, food and feed. Food

Science and Technology, 26(1), 28-30.

Miller, M.R., Nichols, P.D., Carter, C.G. 2008. n-3 Oil sources for use in aquaculture

– alternatives to the unsustainable harvest of wild fish. Nutrition Research

Reviews, 21(02), 85-96.

Mishra, M.K., Kumaraguru, T., Sheelu, G., Fadnavis, N.W. 2009. Lipase activity of

Lecitase® Ultra: characterization and applications in enantioselective

reactions. Tetrahedron: Asymmetry, 20(24), 2854-2860.

Mishra, S.K., Suh, W.I., Farooq, W., Moon, M., Shrivastav, A., Park, M.S., Yang, J.-

W. 2014. Rapid quantification of microalgal lipids in aqueous medium by a

simple colorimetric method. Bioresource Technology, 155, 330-333.

Mogren, H., Lindblom, M., Hedenskog, G. 1974. Mechanical disintegration of

microorganisms in an industrial homogenizer. Biotechnology and

Bioengineering, 16(2), 261-274.

Molfino, A., Gioia, G., Fanelli, F., Muscaritoli, M. 2014. The role for dietary omega-

3 fatty acids supplementation in older adults. Nutrients, 6(10), 4058.

Monks, L.M., Rigo, A., Mazutti, M.A., Vladimir Oliveira, J., Valduga, E. 2013. Use

of chemical, enzymatic and ultrasound-assisted methods for cell disruption to

157

Page 179: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References

obtain carotenoids. Biocatalysis and Agricultural Biotechnology, 2(2), 165-

169.

Mubarak, M., Shaija, A., Suchithra, T.V. 2015. A review on the extraction of lipid

from microalgae for biodiesel production. Algal Research, 7, 117-123.

Murphy, M.G. 1990. Dietary fatty acids and membrane protein function. The Journal

of Nutritional Biochemistry, 1(2), 68-79.

Mutanda, T., Ramesh, D., Karthikeyan, S., Kumari, S., Anandraj, A., Bux, F. 2011.

Bioprospecting for hyper-lipid producing microalgal strains for sustainable

biofuel production. Bioresource Technology, 102(1), 57-70.

Nagle, N., Lemke, P. 1990. Production of methyl ester fuel from microalgae. Applied

Biochemistry and Biotechnology, 24-25(1), 355-361.

Nalder, T.D., Marshall, S., Pfeffer, F.M., Barrow, C.J. 2014. Characterisation of lipase

fatty acid selectivity using novel omega-3 pNP-acyl esters. Journal of

Functional Foods, 6, 259-269.

Namal Senanayake, S.P.J. 2013. 17 - Methods of concentration and purification of

omega-3 fatty acids. in: Separation, Extraction and Concentration Processes

in the Food, Beverage and Nutraceutical Industries, (Ed.) S.S.H. Rizvi,

Woodhead Publishing, pp. 483-505.

Nautiyal, P., Subramanian, K.A., Dastidar, M.G. 2014. Production and

characterization of biodiesel from algae. Fuel Processing Technology, 120, 79-

88.

Neises, B., Steglich, W. 1978. Simple method for the esterification of carboxylic acids.

Angewandte Chemie International Edition in English, 17(7), 522-524.

Nguyen, T.T.T., Le, V.V.M. 2013. Effects of ultrasound on cellulolytic activity of

cellulase complex. International Food Research Journal, 20(2), 557-563.

Nichols, P.D., McManus, A., Krail, K., Sinclair, A.J., Miller, M. 2014. Recent

advances in omega-3: Health benefits, sources, products and bioavailability.

Nutrients, 6(9), 3727-3733.

Nogueira, L.A.H. 2011. Does biodiesel make sense? Energy, 36(6), 3659-3666.

Nunes, E., Cavaco, A., Carvalho, C. 2014. Children's health risk and benefits of fish

consumption: Risk indices based on a diet diary follow-up of two weeks.

Journal of Toxicology and Environmental Health - Part A: Current Issues,

77(1-3), 103-114.

158

Page 180: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Oh, S.H., Han, J.G., Kim, Y., Ha, J.H., Kim, S.S., Jeong, M.H., Jeong, H.S., Kim,

N.Y., Cho, J.S., Yoon, W.B., Lee, S.Y., Kang, D.H., Lee, H.Y. 2009. Lipid

production in Porphyridium cruentum grown under different culture

conditions. Journal of Bioscience and Bioengineering, 108(5), 429-434.

Okada, T., Morrissey, M.T. 2007. Production of n - 3 polyunsaturated fatty acid

concentrate from sardine oil by lipase-catalyzed hydrolysis. Food Chemistry,

103(4), 1411-1419.

Olofsson, M., Lamela, T., Nilsson, E., Bergé, J.-P., del Pino, V., Uronen, P., Legrand,

C. 2014. Combined effects of nitrogen concentration and seasonal changes on

the production of lipids in Nannochloropsis oculata. Marine Drugs, 12(4),

1891-1910.

Orlando Beys Silva, W., Mitidieri, S., Schrank, A., Vainstein, M.H. 2005. Production

and extraction of an extracellular lipase from the entomopathogenic fungus

Metarhizium anisopliae. Process Biochemistry, 40(1), 321-326.

Pahl, S., Lee, A., Kalaitzidis, T., Ashman, P., Sathe, S., Lewis, D. 2013. Harvesting,

thickening and dewatering microalgae biomass. in: Algae for Biofuels and

Energy, (Eds.) M.A. Borowitzka, N.R. Moheimani, Vol. 5, Springer

Netherlands, pp. 165-185.

Papadia, C., Coruzzi, A., Montana, C., Di Mario, F., Franze, A., Forbes, A. 2010.

Omega-3 fatty acids in the maintenance of ulcerative colitis. JRSM Short Rep,

1(1), 15.

Park, J.-Y., Park, M.S., Lee, Y.-C., Yang, J.-W. 2015. Advances in direct

transesterification of algal oils from wet biomass. Bioresource Technology,

184, 267-275.

Peng, Q., Wang, X., Shang, M., Huang, J., Guan, G., Li, Y., Shi, B. 2014. Isolation of

a novel alkaline-stable lipase from a metagenomic library and its specific

application for milkfat flavor production. Microbial Cell Factories, 13(1), 1.

Pérez, D., Martín, S., Fernández-Lorente, G., Filice, M., Guisán, J.M., Ventosa, A.,

García, M.T., Mellado, E. 2011. A novel halophilic lipase, lipbl, showing high

efficiency in the production of eicosapentaenoic acid (EPA). PLoS ONE, 6(8),

e23325.

Phukan, M.M., Chutia, R.S., Konwar, B.K., Kataki, R. 2011. Microalgae Chlorella as

a potential bio-energy feedstock. Applied Energy, 88(10), 3307-3312.

159

Page 181: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Pinzi, S., Leiva, D., López-García, I., Redel-Macías, M.D., Dorado, M.P. 2014. Latest

trends in feedstocks for biodiesel production. Biofuels, Bioproducts and

Biorefining, 8(1), 126-143.

Postma, P.R., Miron, T.L., Olivieri, G., Barbosa, M.J., Wijffels, R.H., Eppink, M.H.M.

2015. Mild disintegration of the green microalgae Chlorella vulgaris using

bead milling. Bioresource Technology, 184, 297-304.

Poudyal, H., Panchal, S.K., Diwan, V., Brown, L. 2011. Omega-3 fatty acids and

metabolic syndrome: Effects and emerging mechanisms of action. Progress in

Lipid Research, 50(4), 372-387.

Prabakaran, P., Ravindran, A.D. 2011. A comparative study on effective cell

disruption methods for lipid extraction from microalgae. Letters in Applied

Microbiology, 53(2), 150-154.

Priji, P., Unni, K.N., Sajith, S., Binod, P., Benjamin, S. 2015. Production,

optimization, and partial purification of lipase from Pseudomonas sp. strain

BUP6, a novel rumen bacterium characterized from Malabari goat.

Biotechnology and Applied Biochemistry, 62(1), 71-78.

Prommuak, C., Pavasant, P., Quitain, A.T., Goto, M., Shotipruk, A. 2012. Microalgal

lipid extraction and evaluation of single-step biodiesel production. Engineering

Journal, 16(5), 157-166.

Pruvost, J., Van Vooren, G., Le Gouic, B., Couzinet-Mossion, A., Legrand, J. 2011.

Systematic investigation of biomass and lipid productivity by microalgae in

photobioreactors for biodiesel application. Bioresource Technology, 102(1),

150-158.

Puri, M., Gupta, S., Pahuja, P., Kaur, A., Kanwar, J.R., Kennedy, J.F. 2010. Cell

-d-galactosidase from

kluyveromyces marxianus YW-1 for lactose hydrolysis in milk. Applied

Biochemistry and Biotechnology, 160(1), 98-108.

Qawasmi, A., Landeros-Weisenberger, A., Bloch, M.H. 2013. Meta-analysis of

LCPUFA supplementation of infant formula and visual acuity. Pediatrics,

131(1), e262-e272.

Qi, K., Fan, C., Jiang, J., Zhu, H., Jiao, H., Meng, Q., Deckelbaum, R.J. 2008. Omega-

3 fatty acid containing diets decrease plasma triglyceride concentrations in

mice by reducing endogenous triglyceride synthesis and enhancing the blood

clearance of triglyceride-rich particles. Clinical Nutrition, 27(3), 424-430.

160

Page 182: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Qin, X.-L., Lan, D.-M., Zhong, J.-F., Liu, L., Wang, Y.-H., Yang, B. 2014. Fatty acid

specificity of T1 lipase and its potential in acylglycerol synthesis. Journal of

the Science of Food and Agriculture, 94(8), 1614-1621.

R, R.K., P, H.R., Muthu, A. 2015. Lipid extraction methods from microalgae: A

comprehensive review. Frontiers in Energy Research, 2.

Rabinowitz, C., Douek, J., Weisz, R., Shabtay, A., Rinkevich, B. 2006. Isolation and

characterization of four novel thraustochytrid strains from a colonial tunicate.

Indian Journal of Marine Sciences, 35(4), 341-350.

Raghukumar, S. 2002. Ecology of the marine protists, the Labyrinthulomycetes

(Thraustochytrids and Labyrinthulids). European Journal of Protistology,

38(2), 127-145.

Raghukumar, S. 2008. Thraustochytrid marine protists: production of PUFAs and

other emerging technologies. Marine Biotechnology, 10(6), 631-640.

Raghukumar, S., Damare, V.S. 2011. Increasing evidence for the important role of

Labyrinthulomycetes in marine ecosystems. Botanica Marina, 54(1), 3-11.

Ramluckan, K., Moodley, K.G., Bux, F. 2014. An evaluation of the efficacy of using

selected solvents for the extraction of lipids from algal biomass by the soxhlet

extraction method. Fuel, 116, 103-108.

Rao, A.V., Rao, L.G. 2007. Carotenoids and human health. Pharmacological

Research, 55(3), 207-216.

Ratledge, C. 2013. Microbial oils: an introductory overview of current status and

future prospects. OCL, 20(6), D602.

Ratnayake, W.M.N., Galli, C. 2009. Fat and fatty acid terminology, methods of

analysis and fat digestion and metabolism: A background review paper. Annals

of Nutrition and Metabolism, 55(1-3), 8-43.

Ren, H.-Y., Liu, B.-F., Kong, F., Zhao, L., Ren, N.-Q. 2015. Improved Nile red

staining of Scenedesmus sp. by combining ultrasonic treatment and three-

dimensional excitation emission matrix fluorescence spectroscopy. Algal

Research, 7, 11-15.

Richardson, A.J., Montgomery, P. 2005. The oxford-durham study: a randomized,

controlled trial of dietary supplementation with fatty acids in children with

developmental coordination disorder. Pediatrics, 115(5), 1360-1366.

Richmond, G.S., Smith, T.K. 2011. Phospholipases A1. International Journal of

Molecular Sciences, 12(1), 588.

161

Page 183: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Rodolfi, L., Chini Zittelli, G., Bassi, N., Padovani, G., Biondi, N., Bonini, G., Tredici,

M.R. 2009. Microalgae for oil: Strain selection, induction of lipid synthesis

and outdoor mass cultivation in a low-cost photobioreactor. Biotechnology and

Bioengineering, 102(1), 100-112.

Rubio-Rodríguez, N., Beltrán, S., Jaime, I., de Diego, S.M., Sanz, M.T., Carballido,

J.R. 2010. Production of omega-3 polyunsaturated fatty acid concentrates: A

review. Innovative Food Science and Emerging Technologies, 11(1), 1-12.

Rumin, J., Bonnefond, H., Saint-Jean, B., Rouxel, C., Sciandra, A., Bernard, O.,

Cadoret, J.P., Bougaran, G. 2015. The use of fluorescent Nile red and BODIPY

for lipid measurement in microalgae. Biotechnology for Biofuels, 1-16.

Ryckebosch, E., Bermúdez, S., Termote-Verhalle, R., Bruneel, C., Muylaert, K.,

Parra-Saldivar, R., Foubert, I. 2014. Influence of extraction solvent system on

the extractability of lipid components from the biomass of Nannochloropsis

gaditana. Journal of Applied Phycology, 26(3), 1501-1510.

Ryckebosch, E., Bruneel, C., Muylaert, K., Foubert, I. 2012a. Microalgae as an

alternative source of omega-3 long chain polyunsaturated fatty acids. Lipid

Technology, 24(6), 128-130.

Ryckebosch, E., Muylaert, K., Foubert, I. 2012b. optimization of an analytical

procedure for extraction of lipids from microalgae. Journal of the American

Oil Chemists' Society, 89(2), 189-198.

Ryu, B.-G., Kim, K., Kim, J., Han, J.-I., Yang, J.-W. 2013. Use of organic waste from

the brewery industry for high-density cultivation of the docosahexaenoic acid-

rich microalga, Aurantiochytrium sp. KRS101. Bioresource Technology, 129,

351-359.

Sadeghinezhad, E., Kazi, S.N., Sadeghinejad, F., Badarudin, A., Mehrali, M., Sadri,

R., Reza Safaei, M. 2014. A comprehensive literature review of bio-fuel

performance in internal combustion engine and relevant costs involvement.

Renewable and Sustainable Energy Reviews, 30, 29-44.

Safi, C., Charton, M., Ursu, A.V., Laroche, C., Zebib, B., Pontalier, P.Y., Vaca-Garcia,

C. 2014a. Release of hydro-soluble microalgal proteins using mechanical and

chemical treatments. Algal Research, 3(1), 55-60.

Safi, C., Ursu, A.V., Laroche, C., Zebib, B., Merah, O., Pontalier, P.-Y., Vaca-Garcia,

C. 2014b. Aqueous extraction of proteins from microalgae: Effect of different

cell disruption methods. Algal Research, 3, 61-65.

162

Page 184: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Salama, E.-S., Kim, H.-C., Abou-Shanab, R.I., Ji, M.-K., Oh, Y.-K., Kim, S.-H., Jeon,

B.-H. 2013. Biomass, lipid content, and fatty acid composition of freshwater

Chlamydomonas mexicana and Scenedesmus obliquus grown under salt stress.

Bioprocess and Biosystems Engineering, 36(6), 827-833.

Salihu, A., Alam, M.Z. 2015. Solvent tolerant lipases: A review. Process

Biochemistry, 50(1), 86-96.

Samori, C., Lopez Barreiro, D., Vet, R., Pezzolesi, L., Brilman, D.W.F., Galletti, P.,

Tagliavini, E. 2013. Effective lipid extraction from algae cultures using

switchable solvents. Green Chemistry, 15(2), 353-356.

Sana, B. 2015. Marine Microbial Enzymes: Current Status and Future Prospects. in:

Hb25_Springer Handbook of Marine Biotechnology, (Ed.) S.-K. Kim, Springer

Berlin Heidelberg, pp. 905-917.

Sangeetha, R., Arulpandi, I., Geetha, A. 2011. Bacterial lipases as potential industrial

biocatalysts: An overview. Research Journal of Microbiology, 6(1), 1-24.

Sathish, A., Marlar, T., Sims, R.C. 2015. Optimization of a wet microalgal lipid

extraction procedure for improved lipid recovery for biofuel and bioproduct

production. Bioresource Technology, 193, 15-24.

Sathish, A., Sims, R.C. 2012. Biodiesel from mixed culture algae via a wet lipid

extraction procedure. Bioresource Technology, 118, 643-647.

Satpati, G., Pal, R. 2014. Rapid detection of neutral lipid in green microalgae by flow

cytometry in combination with Nile red staining-an improved technique.

Annals of Microbiology, 1-13.

Sauer, T., Galinski, E.A. 1998. Bacterial milking: A novel bioprocess for production

of compatible solutes. Biotechnology and Bioengineering, 57(3), 306-313.

Schnurr, P.J., Allen, D.G. 2015. Factors affecting algae biofilm growth and lipid

production: A review. Renewable and Sustainable Energy Reviews, 52, 418-

429.

Schreck, S., Grunden, A. 2014. Biotechnological applications of halophilic lipases and

thioesterases. Applied Microbiology and Biotechnology, 98(3), 1011-1021.

Schütte, H., Kroner, K.H., Hustedt, H., Kula, M.R. 1983. Experiences with a 20 litre

industrial bead mill for the disruption of microorganisms. Enzyme and

Microbial Technology, 5(2), 143-148.

Schutte, H., Kula, M.R. 1990. Pilot-scale and process-scale techniques for cell

disruption. Biotechnology and Applied Biochemistry, 12(6), 599-620.

163

Page 185: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Schwenzfeier, A., Wierenga, P.A., Gruppen, H. 2011. Isolation and characterization

of soluble protein from the green microalgae Tetraselmis sp. Bioresource

Technology, 102(19), 9121-9127.

Scott, S.D., Armenta, R.E., Berryman, K.T., Norman, A.W. 2011. Use of raw glycerol

to produce oil rich in polyunsaturated fatty acids by a thraustochytrid. Enzyme

and Microbial Technology, 48(3), 267-272.

Seabert, T.A., Pal, S., Pinet, B.M., Haman, F., Robidoux, M.A., Imbeault, P.,

Krümmel, E.M., Kimpe, L.E., Blais, J.M. 2014. Elevated contaminants

-3 fatty acids in wild food consumers of

two remote first nations communities in Northern Ontario, Canada. PLoS ONE,

9(3), e90351.

Senanayake, S.P.J.N. 2000. Enzyme-assisted synthesis of structured lipids containing

long-chain omega-3 and omega-6 polyunsaturated fatty acids, Vol. NQ73570,

Memorial University of Newfoundland (Canada). Ann Arbor, pp. 345-345 p.

Shahidi F and Senanayake, S.P.J.N. 2006. Nutraceuticals and specialty lipids,

Nutraceuticals and specialty lipids and their co-products. Boca RAton, FL,

CRC Press, 1-25.

Shahidi, F., Wanasundara, U.N. 1998. Omega-3 fatty acid concentrates: nutritional

aspects and production technologies. Trends in Food Science & Technology,

9(6), 230-240.

Sharma, A., Chaurasia, S.P., Dalai, A.K. 2013. Non-selective hydrolysis of tuna fish

oil for producing free fatty acids containing docosahexaenoic acid. The

Canadian Journal of Chemical Engineering, 92(2), 344-354.

Sharma, D., Kumbhar, B.K., Verma, A.K., Tewari, L. 2014. Optimization of critical

growth parameters for enhancing extracellular lipase production by

alkalophilic Bacillus sp. Biocatalysis and Agricultural Biotechnology, 3(4),

205-211.

Sheelu, G., Kavitha, G., Fadnavis, N. 2008. Efficient immobilization of lecitase in

gelatin hydrogel and degumming of rice bran oil using a spinning basket

reactor. Journal of the American Oil Chemists' Society, 85(8), 739-748.

Shen, Y., Pei, Z., Yuan, W., Mao, E. 2009. Effect of nitrogen and extraction method

on algae lipid yield. International Journal of Agricultural and Biological

Engineering, 2(1), 51-57.

164

Page 186: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Sheng, J., Vannela, R., Rittmann, B.E. 2012. Disruption of Synechocystis PCC 6803

for lipid extraction. Water Science and Technology, 65(3), 567-573.

Sheng, J., Vannela, R., Rittmann, B.E. 2011a. Evaluation of cell-disruption effects of

pulsed-electric-field treatment of Synechocystis PCC 6803. Environmental

Science and Technology, 45(8), 3795-3802.

Sheng, J., Vannela, R., Rittmann, B.E. 2011b. Evaluation of methods to extract and

quantify lipids from Synechocystis PCC 6803. Bioresource Technology,

102(2), 1697-1703.

Shimada, Y., Maruyama, K., Sugihara, A., Baba, T., Komemushi, S., Moriyama, S.,

Tominaga, Y. 1998. Purification of ethyl docosahexaenoate by selective

alcoholysis of fatty acid ethyl esters with immobilized Rhizomucor miehei

lipase. Journal of the American Oil Chemists' Society, 75(11), 1565-1571.

Shimada, Y., Maruyama, K., Sugihara, A., Moriyama, S., Tominaga, Y. 1997.

Purification of docosahexaenoic acid from tuna oil by a two-step enzymatic

method: Hydrolysis and selective esterification. Journal of the American Oil

Chemists' Society, 74(11), 1441-1446.

Sijtsma, L., de Swaaf, M.E. 2004. Biotechnological production and applications of the

-3 polyunsaturated fatty acid docosahexaenoic acid. Applied Microbiology

and Biotechnology, 64(2), 146-153.

Silitonga, A.S., Masjuki, H.H., Mahlia, T.M.I., Ong, H.C., Chong, W.T., Boosroh,

M.H. 2013. Overview properties of biodiesel diesel blends from edible and

non-edible feedstock. Renewable and Sustainable Energy Reviews, 22, 346-

360.

Singh, A., Mukhopadhyay, M. 2012. Overview of fungal lipase: A Review. Applied

Biochemistry and Biotechnology, 166(2), 486-520.

Singh, A., Nigam, P.S., Murphy, J.D. 2011. Renewable fuels from algae: An answer

to debatable land based fuels. Bioresource Technology, 102(1), 10-16.

Singh, D., Gupta, A., Wilkens, S.L., Mathur, A.S., Tuli, D.K., Barrow, C.J., Puri, M.

2015a. Understanding response surface optimisation to the modeling of

Astaxanthin extraction from a novel strain Thraustochytrium sp. S7. Algal

Research, 11, 113-120.

Singh, D., Mathur, A.S., Tuli, D.K., Puri, M., Barrow, C.J. 2015b. Propyl gallate and

butylated hydroxytoluene influence the accumulation of saturated fatty acids,

165

Page 187: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References

omega-3 fatty acid and carotenoids in thraustochytrids. Journal of Functional

Foods, 15, 186-192.

Singh, D., Puri, M., Wilkens, S., Mathur, A.S., Tuli, D.K., Barrow, C.J. 2013.

Characterization of a new zeaxanthin producing strain of Chlorella

saccharophila isolated from New Zealand marine waters. Bioresource

Technology, 143, 308-314.

Singh, R., Gupta, N., Goswami, V., Gupta, R. 2006. A simple activity staining protocol

for lipases and esterases. Applied Microbiology and Biotechnology, 70(6), 679-

682.

Skinner, E.R., Watt, C., Besson, J.A.O., Best, P.V. 1993. Differences in the fatty acid

composition of the grey and white matter of different regions of the brains of

patients with Alzheimer's disease and control subjects. Brain, 116(3), 717-725.

Slade, R., Bauen, A. 2013. Micro-algae cultivation for biofuels: Cost, energy balance,

environmental impacts and future prospects. Biomass and Bioenergy, 53, 29-

38.

Soliev, A.B., Hosokawa, K., Enomoto, K. 2011. Bioactive pigments from marine

bacteria: applications and physiological roles. Evidence-Based

Complementary and Alternative Medicine, 2011, 17.

B. 2012. Growth,

lipid extraction and thermal degradation of the microalga Chlorella vulgaris.

New Biotechnology, 29(3), 325-331.

Sprague, M., Walton, J., Campbell, P.J., Strachan, F., Dick, J.R., Bell, J.G. 2015.

Replacement of fish oil with a DHA-rich algal meal derived from

Schizochytrium sp. on the fatty acid and persistent organic pollutant levels in

diets and flesh of Atlantic salmon (Salmo salar, L.) post-smolts. Food

Chemistry, 185, 413-421.

Steriti, A., Rossi, R., Concas, A., Cao, G. 2014. A novel cell disruption technique to

enhance lipid extraction from microalgae. Bioresource Technology, 164, 70-

77.

Stoytcheva, M., Montero, G., Zlatev, R., León, J.Á., Gochev, V. 2012. Analytical

methods for lipases activity determination: A review. Current Analytical

Chemistry, 8(3), 400-407.

Supriya, G., Ramachandra, T.V. 2012. Optimal extraction of lipids from microalgae,

Microcystis. Nature Environment and Pollution Technology, 11(2), 213-218.

166

Page 188: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Swanson, D., Block, R., Mousa, S.A. 2012. Omega-3 fatty acids EPA and DHA:

Health benefits throughout life. Advances in Nutrition: An International

Review Journal, 3(1), 1-7.

Sztajer, H., Maliszewska, I., Wieczorek, J. 1988. Production of exogenous lipases by

bacteria, fungi, and actinomycetes. Enzyme and Microbial Technology, 10(8),

492-497.

Tai, E.K.K., Wang, X.B., Chen, Z.-Y. 2013. An update on adding docosahexaenoic

acid (DHA) and arachidonic acid (AA) to baby formula. Food & Function,

4(12), 1767-1775.

Talebi, A.F., Mohtashami, S.K., Tabatabaei, M., Tohidfar, M., Bagheri, A.,

Zeinalabedini, M., Hadavand Mirzaei, H., Mirzajanzadeh, M., Malekzadeh

Shafaroudi, S., Bakhtiari, S. 2013. Fatty acids profiling: A selective criterion

for screening microalgae strains for biodiesel production. Algal Research, 2(3),

258-267.

Taoka, Y., Nagano, N., Okita, Y., Izumida, H., Sugimoto, S., Hayashi, M. 2011. Effect

of Tween 80 on the growth, lipid accumulation and fatty acid composition of

Thraustochytrium aureum ATCC 34304. Journal of Bioscience and

Bioengineering, 111(4), 420-424.

Taubert, J., Krings, U., Berger, R. 2000. A comparative study on the disintegration of

filamentous fungi. Journal of Microbiological Methods, 42, 225 - 232.

Tengku-Rozaina, T.M., Birch, E.J. 2014. Positional distribution of fatty acids on hoki

and tuna oil triglycerides by pancreatic lipase and 13C NMR analysis.

European Journal of Lipid Science and Technology, 116(3), 272-281.

Thompson Jr, G.A. 1996. Lipids and membrane function in green algae. Biochimica

et Biophysica Acta (BBA) - Lipids and Lipid Metabolism, 1302(1), 17-45.

Tiesinga, J.J.W., Pouderoyen, G.v., Nardini, M., Ransac, S., Dijkstra, B.W. 2007.

structural basis of phospholipase activity of Staphylococcus hyicus lipase.

Journal of Molecular Biology, 371(2), 447-456.

Tiquia, S.M. 2002. Evolution of extracellular enzyme activities during manure

composting. Journal of Applied Microbiology, 92(4), 764-775.

wska, B., Herold, F. 2010. Effect of selenium enrichment on

antioxidant activities and chemical composition of Lentinula edodes (Berk.)

Pegl. mycelial extracts. Food and Chemical Toxicology, 48(4), 1085-1091.

167

Page 189: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Udani, J.K., Ritz, B.W. 2013. High potency fish oil supplement improves omega-3

fatty acid status in healthy adults: An open-label study using a web-based,

virtual platform. Nutrition Journal, 12(1).

US Department of Agriculture. National nutrient database. (Retrieved 12 April 2015)

Valverde, L.M., Moreno, P.A.G., Callejón, M.J.J., Cerdán, L.E., Medina, A.R. 2013.

Concentration of eicosapentaenoic acid (EPA) by selective alcoholysis

catalyzed by lipases. European Journal of Lipid Science and Technology, 115,

990-1004.

van Eijck, J., Batidzirai, B., Faaij, A. 2014. Current and future economic performance

of first and second generation biofuels in developing countries. Applied

Energy, 135, 115-141.

Van Handel, E. 1985. Rapid determination of total lipids in mosquitoes. Journal of the

American Mosquito Control Association, 1(3), 302-304.

Van Kampen, M.D., Simons, J.W.F.A., Dekker, N., Egmond, M.R., Verheij, H.M.

1998. The phospholipase activity of Staphylococcus hyicus lipase strongly

depends on a single Ser to Val mutation. Chemistry and Physics of Lipids,

93(1-2), 39-45.

Vatassery, G.T., Sheridan, M.A., Krezowski, A.M., Divine, A.S., Bach, H.L. 1981.

Use of the sulfo-phospo-vanillin reaction in a routine method for determining

total lipids in human cerebrospinal fluid. Clinical Biochemistry, 14(1), 21-24.

Vello, V., Phang, S.-M., Chu, W.-L., Abdul Majid, N., Lim, P.-E., Loh, S.-K. 2014.

Lipid productivity and fatty acid composition-guided selection of Chlorella

strains isolated from Malaysia for biodiesel production. Journal of Applied

Phycology, 26(3), 1399-1413.

Venkateshwarlu, G., Let, M.B., Meyer, A.S., Jacobsen, C. 2004. Modeling the sensory

impact of defined combinations of volatile lipid oxidation products on fishy

and metallic off-flavors. Journal of Agricultural and Food Chemistry, 52(6),

1635-1641.

Vester, J., Glaring, M., Stougaard, P. 2014. Discovery of novel enzymes with

industrial potential from a cold and alkaline environment by a combination of

functional metagenomics and culturing. Microbial Cell Factories, 13(1), 72.

Vyas, A.P., Verma, J.L., Subrahmanyam, N. 2010. A review on FAME production

processes. Fuel, 89(1), 1-9.

168

Page 190: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Wahidin, S., Idris, A., Shaleh, S.R.M. 2014. Rapid biodiesel production using wet

microalgae via microwave irradiation. Energy Conversion and Management,

84, 227-233.

Wang, C.-h., Guo, R.-f., Yu, H.-w., Jia, Y.-m. 2008. Cloning and sequence analysis of

a novel cold-adapted lipase gene from strain lip35 (Pseudomonas sp.).

Agricultural Sciences in China, 7(10), 1216-1221.

Wang, D., Li, Y., Hu, X., Su, W., Zhong, M. 2015. Combined enzymatic and

mechanical cell disruption and lipid extraction of green alga Neochloris

oleoabundans. International Journal of Molecular Sciences, 16(4), 7707-7722.

Wang, Y., Zhao, M., Song, K., Wang, L., Tang, S., Riley, W.W. 2010. Partial

hydrolysis of soybean oil by phospholipase A1 (Lecitase Ultra). Food

Chemistry, 121(4), 1066-1072.

Wawrik, B., Harriman, B.H. 2010. Rapid, colorimetric quantification of lipid from

algal cultures. Journal of Microbiological Methods, 80(3), 262-266.

Wijffels, R.H., Barbosa, M.J., Eppink, M.H.M. 2010. Microalgae for the production

of bulk chemicals and biofuels. Biofuels, Bioproducts and Biorefining, 4(3),

287-295.

Wilkens, S.L., Maas, E.W. 2012. Development of a novel technique for axenic

isolation and culture of thraustochytrids from New Zealand marine

environments. Journal of Applied Microbiology, 112(2), 346-352.

William, B., Craig, W., James, M. 2005. Development of a docosahexaenoic acid

production technology using Schizochytrium. in: Single Cell Oils, AOCS

Publishing.

Williams, P.J.l.B., Laurens, L.M.L. 2010. Microalgae as biodiesel & biomass

feedstocks: Review & analysis of the biochemistry, energetics & economics.

Energy & Environmental Science, 3(5), 554-590.

Winkler, U.K., Stuckmann, M. 1979. Glycogen, hyaluronate, and some other

polysaccharides greatly enhance the formation of exolipase by Serratia

marcescens. Journal of Bacteriology, 138(3), 663-670.

Winwood, R.J. 2013. Recent developments in the commercial production of DHA and

EPA rich oils from micro-algae. OCL, 20(6), D604.

Wong, M.K.M., Tsui, C.K.M., Au, D.W.T., Vrijmoed, L.L.P. 2008. Docosahexaenoic

acid production and ultrastructure of the thraustochytrid Aurantiochytrium

169

Page 191: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References

mangrovei MP2 under high glucose concentrations. Mycoscience, 49(4), 266-

270.

Wu, Q., Zhou, T., Ma, L., Yuan, D., Peng, Y. 2015. Protective effects of dietary

-3 polyunsaturated fatty acids on the visual

acuity of school-age children with lower IQ or attention-deficit hyperactivity

disorder. Nutrition, 31(7–8), 935-940.

WWW.nal.usda.gov/fnic/foodcomp/search/ (Retrieved 17 April 2015)

Xin, L., Hong-ying, H., Jia, Y. 2010. Lipid accumulation and nutrient removal

properties of a newly isolated freshwater microalga, Scenedesmus sp. LX1,

growing in secondary effluent. New Biotechnology, 27(1), 59-63.

Yang, B., Zhou, R., Yang, J.-G., Wang, Y.-H., Wang, W.-F. 2008. Insight into the

enzymatic degumming process of soybean oil. Journal of the American Oil

Chemists' Society, 85(5), 421-425.

Yang, F., Cheng, C., Long, L., Hu, Q., Jia, Q., Wu, H., Xiang, W. 2015. Extracting

lipids from several species of wet microalgae using ethanol at room

temperature. Energy & Fuels, 29(4), 2380-2386.

Yang, F., Xiang, W., Sun, X., Wu, H., Li, T., Long, L. 2014. A novel lipid extraction

method from wet microalga Picochlorum sp. at room temperature. Marine

Drugs, 12(3), 1258.

Yang, H.-L., Lu, C.-K., Chen, S.-F., Chen, Y.-M., Chen, Y.-M. 2010. isolation and

characterization of taiwanese heterotrophic microalgae: screening of strains for

docosahexaenoic acid (DHA) production. Marine Biotechnology, 12(2), 173-

185.

Yang, J.G., Wang, Y.H., Yang, B., Mainda, G., Guo, Y. 2006. Degumming of

vegetable oil by a new microbial lipase. Food Technology and Biotechnology,

44(1), 101-104.

Yao, L., Gerde, J., Wang, T. 2012. Oil extraction from microalga Nannochloropsis sp.

with isopropyl alcohol. Journal of the American Oil Chemists' Society, 89(12),

2279-2287.

Yap, B.H.J., Crawford, S.A., Dumsday, G.J., Scales, P.J., Martin, G.J.O. 2014. A

mechanistic study of algal cell disruption and its effect on lipid recovery by

solvent extraction. Algal Research, 5, 112-120.

170

Page 192: Downstream processing of lipids and lipases from ...dro.deakin.edu.au/eserv/DU:30083099/byreddy... · List of publications 1. Avinesh R. Byreddy, Colin Barrow, Munish Puri (2016),

References Yashodhara, B.M., Umakanth, S., Pappachan, J.M., Bhat, S.K., Kamath, R., Choo,

B.H. 2009. Omega-3 fatty acids: a comprehensive review of their role in health

and disease. Postgraduate Medical Journal, 85(1000), 84-90.

Yen, H.W., Hu, I.C., Chen, C.Y., Ho, S.H., Lee, D.J., Chang, J.S. 2013. Microalgae-

based biorefinery - From biofuels to natural products. Bioresource Technology,

135, 166-174.

Yoo, G., Park, W.-K., Kim, C.W., Choi, Y.-E., Yang, J.-W. 2012. Direct lipid

extraction from wet Chlamydomonas reinhardtii biomass using osmotic shock.

Bioresource Technology, 123, 717-722.

You, L.L., Baharin, B.S. 2006. Effects of enzymatic hydrolysis on crude palm olein

by lipase from candida rugosa. Journal of Food Lipids, 13(1), 73-87.

Yuan, D., Lan, D., Xin, R., Yang, B., Wang, Y. 2014. Biochemical properties of a new

cold-active mono- and diacylglycerol lipase from marine member Janibacter

sp. strain HTCC2649. International Journal of Molecular Sciences, 15(6),

10554-10566.

Zhang, C., Kim, S.-K. 2010. research and application of marine microbial enzymes:

status and prospects. Marine Drugs, 8(6), 1920-1934.

Zheng, H., Yin, J., Gao, Z., Huang, H., Ji, X., Dou, C. 2011. Disruption of Chlorella

vulgaris cells for the release of biodiesel-producing lipids: a comparison of

grinding, ultrasonication, bead milling, enzymatic lysis, and microwaves.

Applied Biochemistry and Biotechnology, 164(7), 1215-1224.

Zhou, P.-P., Lu, M.-B., Li, W., Yu, L.-J. 2010. Microbial production of

docosahexaenoic acid by a low temperature-adaptive strain Thraustochytriidae

sp. Z105: screening and optimization. Journal of Basic Microbiology, 50(4),

380-387.

Zhu, L.D., Hiltunen, E., Antila, E., Zhong, J.J., Yuan, Z.H., Wang, Z.M. 2014.

Microalgal biofuels: Flexible bioenergies for sustainable development.

Renewable and Sustainable Energy Reviews, 30, 1035-1046.

Zuta, C., Simpson, B., Chan, H., Phillips, L. 2003. Concentrating PUFA from mackerel

processing waste. Journal of the American Oil Chemists' Society, 80(9), 933-

936.

171