in-situ observation and quantification of microalgae ......the direct observation of these complex...

133
In-situ observation and quantification of microalgae downstream processing on a microfluidic platform by Xiang Cheng A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Mechanical & Industrial Engineering University of Toronto © Copyright by Xiang Cheng 2018

Upload: others

Post on 24-Feb-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

In-situ observation and quantification of microalgae

downstream processing on a microfluidic platform

by

Xiang Cheng

A thesis submitted in conformity with the requirements

for the degree of Doctor of Philosophy

Mechanical & Industrial Engineering

University of Toronto

© Copyright by Xiang Cheng 2018

Page 2: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

ii

In-situ observation and quantification of microalgae downstream

processing on a microfluidic platform

Xiang Cheng

Doctor of Philosophy

Mechanical & Industrial Engineering

University of Toronto

2018

Abstract

Producing biofuels and bioproducts from microalgae is a promising path for low-carbon energy

and products. Microalgal biomass is an attractive feedstock for the generation of carbon neutral

biofuels and high-value bioproducts because of the high growth rate and lipid content of many

microalgae species. Understanding the downstream processing of converting microalgal biomass

to valuable products is a critical step in the biofuel industry. In this thesis, a novel microfluidic

platform capable of precise control of processing parameters and providing optical access to

reactions at high temperature and pressure was developed and applied to observe and quantify the

biomass-to-bioproducts conversions in three distinct studies.

First, for bioenergy application, hydrothermal liquefaction of microalgae was performed on this

microfluidic platform monitored using fluorescence microscopy. A strong shift in the fluorescence

signature from the algal slurry at 675 nm (chlorophyll peak) to a post-HTL stream at 510 nm is

observed for reaction temperatures at 260°C, 280°C, 300°C and 320°C (P = 12 MPa), and occurs

over a timescale on the order of 10 min. Biocrude formation and separation from the aqueous phase

into immiscible droplets is directly observed and occurs over the same timescale.

Second, many algal bioproduct efforts currently focus on high-value products such as astaxanthin

due to the much-improved economics over producing fuels. Hydrothermal disruption of the cell

wall for astaxanthin extraction from wet biomass using high temperature and pressure was

demonstrated and studied using this microfluidic platform. Hydrothermal disruption at a

Page 3: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

iii

temperature of 200 °C was shown to be highly effective, resulting in near-complete astaxanthin

extraction from wet biomass - a significant improvement over traditional methods.

Third, supercritical CO2 has relatively low critical temperature and pressure (31.1 °C and 7.4 MPa)

is considered a greener solvent for bioactive compounds extraction. Supercritical CO2 extractions

of astaxanthin with and without co-solvents (ethanol and olive oil) were performed on the

microfluidic platform to study the extraction mechanism in each case. Astaxanthin extraction using

ScCO2 achieved 92% recovery at 55 °C and 8 MPa applied over 15 hours. With the addition of co-

solvents, ethanol and olive oil, the timescales of extraction process are reduced significantly from

15 hours to a few minutes, representing the fastest complete astaxanthin extraction at such low

pressures.

The direct observation of these complex reaction processes was made possible for the first time

here, allowing visual characterization, fluorescence spectroscopy, and quantitative imaging of the

conversion at the single-cell scale during all stages. This level of insight has simply not been

possible with previous conventional reactors. Although batch reactors have advantages in, for

instance, quantifying yields requiring large volumes of products, microfluidic reactors have

advantages with respect to process control and visualization at cellular level - providing high

resolution, real-time data on complex reactions. The innovative platform and results presented in

this thesis provide new insight in the challenging area of biomass-to-bioproduct conversion, and

provide insight that can inform larger scale operations.

Page 4: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

iv

Acknowledgments

I would like to extend my gratitude and thanks to individuals who have supported me both

professionally and personally throughout my doctoral journey.

First, I would like to express my thanks to my thesis supervisor, Prof. David Sinton, for his

guidance, support and encouragement during the Ph.D. program. I am grateful for giving me the

opportunity to join his research group and encouraging me to strike out in a new direction of my

choice. I am also grateful for all I have learned from and inspired by him, not only the knowledge,

but also the professionalism, management and leadership that he shares on a daily basis.

I would also like to thank my examination committee: Prof. Grant Allen, Prof. Murray Thomson,

Prof. Shulin Chen and Prof. Axel Guenther for their time and effort in evaluating my work and

providing valuable and insightful comments. Also, thank you to those who wrote letters of

recommendation – Behraad Bahreyni, Farid Golnaraghi, Ahmad Rad, Scott Morgan and David

Sinton, your support and confidence have impacted me in so many ways. Moreover, I would like

to thank my lab mates that we shared amazing experience together over the last four years.

Although not exhaustive, I would like to thank Matthew Ooms, Bo Bao, Percival Graham, Brian

Nguyen, Jason Riordon, Seven Qi, Reza Nosrati, Pushan Lele and Tom Burdyny for your help,

encouragement, and inspiration.

Finally, I would like to thank my family for their love and support that was worth much more than

I can express on paper. A big thanks to my parents, Feng Cong and Qingzhong Cheng, for their

unconditional love and support along the way. Special thank you to my aunt Lam Cong who

encouraged me to study abroad and supported me all the way through. My deepest gratitude and

appreciation goes to my beautiful and lovely wife, Xia Zhang, for supporting me throughout my

career and my life in general.

Page 5: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

v

Table of Contents

Acknowledgments.......................................................................................................................... iv

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

List of Figures .............................................................................................................................. viii

List of Tables ............................................................................................................................... xiv

List of Appendices .........................................................................................................................xv

Chapter 1. Forward ....................................................................................................................1

1.1 Motivation ............................................................................................................................1

1.2 Thesis Overview ..................................................................................................................4

Chapter 2. Introduction ..............................................................................................................7

2.1 Microalgae for biofuel production .......................................................................................7

2.1.1 Assessments for downstream conversions ...............................................................8

2.1.2 Water under subcritical conditions ........................................................................13

2.1.3 Characterization of microalgae and post-HTL products ........................................16

2.1.4 HTL of biomass and effects of processing conditions ...........................................19

2.2 Microalgae for high-value bioproducts ..............................................................................22

2.2.1 Cell structure of Haematococcus pluvialis ............................................................23

2.2.2 Cell wall disruption and extraction techniques ......................................................25

2.3 High temperature and pressure microfluidics ....................................................................26

Chapter 3. Current Downstream Processing and Emerging Technologies for Microalgae .....29

3.1 Optimizing downstream processing ...................................................................................29

3.2 Harvesting and dewatering processes ................................................................................30

3.2.1 Biomass thickening ................................................................................................30

3.2.2 Dehydration to dry biomass ...................................................................................31

3.3 Processing for dry microalgae ...........................................................................................32

3.3.1 Lipids extraction ....................................................................................................32

Page 6: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

vi

3.3.2 Extracted lipids and other components ..................................................................33

3.3.3 Processing of whole dry microalgae ......................................................................34

3.4 Processing for wet microalgae biomass .............................................................................35

3.4.1 Hydrothermal liquefaction .....................................................................................36

3.4.2 Supercritical water gasification ..............................................................................39

3.5 Processing for microalgae in culture..................................................................................40

3.5.1 Fermentation ..........................................................................................................40

3.5.2 Anaerobic digestion ...............................................................................................41

3.5.3 Direct secretion ......................................................................................................42

3.5.4 Microalgae-microbial fuel cells .............................................................................42

Chapter 4. Hydrothermal Liquefaction of Microalgae on a Chip for Biocrude Production ....44

4.1 Introduction ........................................................................................................................45

4.2 Experimental ......................................................................................................................47

4.3 Results ................................................................................................................................50

4.4 Conclusion .........................................................................................................................54

4.5 Supplementary Information ...............................................................................................55

4.5.1 Fabrication of microfluidic chip ............................................................................55

4.5.2 Experimental apparatus ..........................................................................................56

4.5.3 High temperature and pressure packaging .............................................................57

Chapter 5. Hydrothermal Disruption of Algae Cells for Astaxanthin Extraction ...................59

5.1 Introduction ........................................................................................................................59

5.2 Experimental setup.............................................................................................................62

5.2.1 Device design and fabrication ................................................................................62

5.2.2 Cell culture and trapping ........................................................................................62

5.2.3 Cell wall disruption and extraction ........................................................................62

5.2.4 Quantifying astaxanthin content ............................................................................63

Page 7: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

vii

5.2.5 HPLC analysis of extracted astaxanthin ................................................................64

5.3 Results and discussion .......................................................................................................64

5.4 Conclusion .........................................................................................................................71

5.5 Supplemental material .......................................................................................................71

Chapter 6. Astaxanthin Extraction from Algae using Supercritical CO2 with Co-solvent ......75

6.1 Introduction ........................................................................................................................75

6.2 Experimental Section .........................................................................................................78

6.3 Results and discussion .......................................................................................................79

6.4 Conclusions ........................................................................................................................86

6.5 Supplementary Information ...............................................................................................86

Chapter 7. Conclusions ............................................................................................................89

7.1 Summary ............................................................................................................................89

7.2 Future Outlook ...................................................................................................................90

References ......................................................................................................................................92

Appendices ...................................................................................................................................110

A1. The full thermodynamic phase envelope of a mixture in 1000 microfluidic chambers ..110

Page 8: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

viii

List of Figures

Figure 1-1. Microalgae based biorefinery integrated with related industries to produce numerous

sustainable deliverables. Reproduced from ref9, © 2010, with permission from Elsevier. ............ 2

Figure 1-2. Sunlight-to-biomass conversion efficiency and losses along the path, as well as value

and market size of related microalgae products. Reproduced from ref15, © 2016, under CC-BY

license. ............................................................................................................................................ 4

Figure 2-1. A summary of life cycle assessments for biofuel production form microalgae. a) Net

energy ratio for microalgae biomass production and b) Illustrative estimates for carbon dioxide

emissions from algal biomass production. Reproduced from ref34, © 2013, with permission from

Elsevier. ........................................................................................................................................ 10

Figure 2-2. a) Energy balance and b) Global Warming Potential for biofuel production using wet

and dry lipid extraction. Reproduced from ref8, © 2013, with permission from American Chemical

Society........................................................................................................................................... 11

Figure 2-3. Techno-economic analysis of liquid fuel produced from hydrothermal liquefaction of

woody biomass a) Effect of improvement on MFSP from state-of-technology (SOT) case to goal

case. b) Sensitivity analysis of parameter variation on the MFSP of the goal case. a) and b) are

reproduced from ref37, © 2014, with permission from Elsevier. c) Sensitivity analysis of the MFSP

of biofuel produced from defatted microalgae via HTL. Reproduced from ref38, © 2014, with

permission from Elsevier. ............................................................................................................. 12

Figure 2-4. Density, static dielectric constant and ion dissociation constant (Kw) of water at 30

MPa as a function of temperature. Reproduced from ref46, © 2008, with permission from Royal

Society of Chemistry..................................................................................................................... 15

Figure 2-5. Characteristics for algae used for HTL. The green, black and red dashed arrows indicate

the mass fraction of lipids, proteins and carbohydrates, respectively. Reproduced from ref10, ©

2014, with permission from Elsevier. ........................................................................................... 17

Figure 2-6. HTL of microalgae procedure and product separation process. Reproduced from ref54,

© 2012, with permission from Elsevier. ....................................................................................... 18

Page 9: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

ix

Figure 2-7. Yields of products from hydrothermal liquefaction of different types of feedstock in:

a) water, b) sodium carbonate, c) formic acid. Reproduced from ref12, © 2011, with permission

from Elsevier. ................................................................................................................................ 19

Figure 2-8. Biocrude yield of HTL as a function of heating rate. a) Temperature profiles with

corresponding biocrude yields from HTL of Nannochloropsis sp. for different set-point

temperatures of 300, 500, 400, and 600 °C. Reproduced from ref65, © 2013, with permission from

American Chemical Society. b) The effect of heating rate on the products yield from HTL of

macroalgae. (T = 350 °C; holding time = 15 min; and biomass/water ratio = 1/10, w/w).

Reproduced from ref66, © 2014, with permission from Elsevier. ................................................. 21

Figure 2-9. Market value of the microalgal components and total selling price of the biomass for

different market scenarios. Reproduced from ref69, © 2016, under CC-BY license. ..................... 23

Figure 2-10. a) Microscopic images of Haematococcus pluvialis in life cycle. (A) Green vegetative

motile cell; (B) Green vegetative palmella cell; (C) Astaxanthin accumulating palmella cell in

transition to aplanospore; (D) Astaxanthin accumulated aplanospore cell. Scale bar: 10 μm.

Reproduced from ref72, © 2016, under CC-BY license. b) Illustration of life cycle of H. pluvialis.

Reproduced from ref73, © 2013, under CC-BY license. ............................................................... 24

Figure 2-11. Summary of five developmental states of the cell wall during aplanospore morphogenesis

in Haematococcus pluvialis: I, 1-week-old flagellates ; II, flagellates at least 2 weeks old just rounding

off; III, 2- to 3-week-old aplanospores; IV, at least 3-week-old aplanospores; V, aplanospores in their

final state. CYP, cytoplasm; IS, interspace; PL, plasmalemma; PW, primary wall; SV, secretory

vesicles; SW, secondary wall; TCL, tripartite crystalline layer; TLS, trilaminar sheath; W1±W7, layers

of the extracellular matrix. Reproduced from ref74, © 2002, with permission from Taylor & Francis.

....................................................................................................................................................... 25

Figure 2-12. Examples of (a) metal, (b) glass and (c) silicon/glass microreactors. Reproduced from

ref20, © 2011, with permission from Elsevier. .............................................................................. 27

Figure 3-1. Summary of current downstream processing techniques for wet microalgae biomass (a-f)

and cells in culture (g-j). (a-b) Hydrothermal liquefaction156,157. Image (b) has been reproduced with

permission from Elsevier157, Copyright 2013. (c) Supercritical water gasification159. Reproduced with

permission from Elsevier, Copyright 2016. (d-e) Ionic liquid treatment for wet extraction194,195. Image

Page 10: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

x

(d) has been reproduced with permission from Royal Society of Chemistry194, Copyright 2014. Image

(e) has been reproduced with permission from Royal Society of Chemistry196, Copyright 2015. (f)

Astaxanthin extraction for hydrothermal disrupted cells197. (g) Fermentation of pretreated wet

biomass161. Image reproduced under CC-BY license, Copyright 2014. (h-i) Anaerobic digestion of algal

biomass169,170. Image (h) has been reproduced with permission from Elsevier170, Copyright 2017. Image

(i) has been reproduced with permission from Elsevier169, Copyright 2016. (j) Microalgae-microbial

fuel cells198. Reproduced with permission from Elsevier, Copyright 2015. This figure is reproduced

from ref91 with permission from The Royal Society of Chemistry. ................................................... 38

Figure 4-1. a) Schematic of hydrothermal liquefaction of microalgae in the microfluidic chip with

in-situ observation of biocrude production using fluorescence microscopy. b) Distinct fluorescence

signatures of algae slurry at the inlet and biocrude at the outlet. Reproduced by permission of The

Royal Society of Chemistry. ......................................................................................................... 46

Figure 4-2. Schematic representation of the assembly of the water-cooled manifold, temperature

controlled heating chuck and the microfluidic chip using a separation glass and double seal O-ring

to prevent cracking from hard contact. Reproduced by permission of The Royal Society of

Chemistry. ..................................................................................................................................... 48

Figure 4-3. a) Normalized fluorescence intensity of algae slurry observed at viewing points along

the channel under 320°C indicating the formation of biocrude over time. b) The progression of

normalized fluorescence intensity of the 510nm peak at the reaction temperature of 260°C, 280°C,

300°C and 320°C indicating higher reaction temperature has higher reaction rate. Solid lines

included as a guide for the eye. Reproduced by permission of The Royal Society of Chemistry.51

Figure 4-4. a) Fluorescence images obtained at viewing points along the channel with increase in

reaction time indicating the progression of biocrude formation. b) Microscopic observation of

fluids at the inlet and outlet via both fluorescence and dark-field imaging. Scale bars: 50 μm.

Reproduced by permission of The Royal Society of Chemistry................................................... 53

Figure 4-5. a) Schematic illustration of achieving high temperature and pressure on a chip by

separating high pressure compression from high temperature area. b) Si/glass chip fabrication

process (from top to bottom, left to right)..................................................................................... 56

Page 11: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

xi

Figure 4-6. Schematic diagram of experimental setup with flow direction indicated by arrows

along the processing path. The flow path of the switching valve at two positions is shown in green

lines. .............................................................................................................................................. 57

Figure 4-7. Detailed drawing of the compression sealing of the microfluidic chip (MF) with algae

slurry fluid indicated in green. O-rings and spacers are used to ensure a quality seal between the

manifold and the chip and to prevent overtightening and damage. .............................................. 58

Figure 5-1. Simplified schematic of on-chip astaxanthin extraction from H. pluvialis. Enlarged

schematics of the cell capture area show initial cell trapping, cell wall disruption and astaxanthin

extraction. Reproduced by permission of The Royal Society of Chemistry. ................................ 61

Figure 5-2. a) Dark field images of mature red cysts of H. pluvialis at both the initial stage and

after solvent-based extraction, for each of five tested cases. All images have the same scale bar,

and were obtained with identical settings using darkfield microscopy. b) Normalized extracted red

content for each case. Reproduced by permission of The Royal Society of Chemistry. .............. 66

Figure 5-3. Time-course images of red cysts treated with hydrothermal processes of 150 oC for 30

min, 200 oC for 30 min, 200 oC for 10 min and 200 oC for 5 min, respectively. The scale bar is

identical for all images. Reproduced by permission of The Royal Society of Chemistry. ........... 67

Figure 5-4. Normalized red content during acetone extraction for six cells treated by hydrothermal

processing at 200 °C for 10 min. Inset images shows corresponding images of one representative

cell at four times during the procedure. Reproduced by permission of The Royal Society of

Chemistry. ..................................................................................................................................... 68

Figure 5-5. HPLC analysis of cell extract products for (a) mechanical extraction using a mortar

and pestle and (b) hydrothermal extraction for 20 min at 200 °C treatment. Reproduced by

permission of The Royal Society of Chemistry. ........................................................................... 69

Figure 5-6. Schematic diagram of the experimental setup with flow direction indicated by arrows

along the processing path. The flow path of the switching valve at two positions is shown using

green lines. .................................................................................................................................... 71

Page 12: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

xii

Figure 5-7. Irradiation spectrum of light used to induce astaxanthin accumulation in

Haematococcus pluvialis. ............................................................................................................. 72

Figure 5-8. Microscope images of red cysts, indicating coloration change during the heating up

phase in the hydrothermal disruption process. The scale bar applies to all images...................... 72

Figure 5-9. Dark field, bright field and fluorescence images of initial cells and after acetone

extraction. Fluorescence 1 and 2 images were taken using FITC (excitation filter: 475/50 nm;

emission filter: 540/50 nm) and TxRed (excitation filter: 559/34 nm; emission filter: 630/69 nm)

filter cubes respectively. The scale bar of 50 µm applies to all images. ...................................... 73

Figure 5-10. Cell wall deformation of red cysts under hydrothermal processes at 200 °C in 10 min

with and without flow. Green dashed lines indicate cell wall boundaries. The scale bar of 50 µm

applies to all images. ..................................................................................................................... 74

Figure 6-1. Single-cellular visualization and quantification of astaxanthin extraction in

Haematococcus pluvialis using supercritical CO2 and co-solvents. Post ScCO2 and ethanol

extraction using acetone provides an overall extraction efficiency metric. Images were taken from

pure ScCO2 extraction experiment at 70 °C. ................................................................................ 80

Figure 6-2. a) Darkfield images H.p cells before and after ScCO2 extraction at 40 °C and 70 °C. b)

The progression of normalized red content for ScCO2 extraction at 40 °C, 55 °C and 70 °C

indicating higher extraction temperatures resulted in higher extraction rates. Solid lines represent

1st order trendline fits to the experimental data with equation given by the side. c) The normalized

red content for ScCO2 extraction process at 55 °C for a complete extraction process over 900

minutes with dark-field snapshots along the process. ................................................................... 82

Figure 6-3. The progression of normalized red content for ScCO2 extraction with ethanol over

1000 seconds at 40 °C, 55 °C and 70 °C. Zoom-in plot for the first 30 s extraction time indicates

a rapid extraction of astaxanthin from ScCO2 with ethanol at 55 °C and 70 °C. ......................... 83

Figure 6-4. a) The comparison of normalized red content for ScCO2 extraction with ethanol and

olive oil at 55 °C over 200 seconds. Time-lapsed snapshots of the extraction process with olive oil

are provided. b) Snapshot of ScCO2 extraction with olive oil indicating three phases: ScCO2, olive

Page 13: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

xiii

oil and water. The boundary of olive oil and CO2 phases are illustrated using yellow and green

dash lines. ...................................................................................................................................... 85

Figure 6-5. Schematic diagram of the experimental setup with the flow direction indicated by

arrows along the processing path. The flow of the switching valve at two positions is shown using

red lines. ........................................................................................................................................ 87

Figure 6-6. Illustration of typical colors of H. pluvialis cells with different concentration of

astaxanthin and the calculated RGB-based astaxanthin content. .................................................. 88

Page 14: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

xiv

List of Tables

Table 3-1 Summary of extraction and conversion techniques for different concentrations of

microalgae biomass. ...................................................................................................................... 36

Table 4-1: Elemental composition and Higher Heating Value of dry algae and biocrude from 1

min, 5 min and 10 min reaction times........................................................................................... 52

Table 4-2: List of components in the apparatus and their purpose. .............................................. 57

Table 6-1: Summary of studies on ScCO2 extraction of astaxanthin from H.pluvialis ................ 77

Table 6-2: Summary of supercritical carbon dioxide extraction experiments and results. ........... 80

Page 15: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

xv

List of Appendices

Appendix 1: The full thermodynamic phase envelope of a mixture in 1000 microfluidic

chambers

Page 16: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

1

Chapter 1.

Forward

Microalgae have great potential to help address critical global challenges on mitigating climate

change, producing sufficient fuels, food and chemicals in a sustainable manner. Understanding the

downstream conversions of microalgae biomass to biofuels and bioproducts is a crucial step –

particularly because these downstream conversion processes can be very energy intensive, eroding

both the environmental and economic motivations for microalgae. In this thesis, an innovative

microfluidic platform will be introduced with three industry-relevant applications demonstrated in

Chapter 4 to 6.

1.1 Motivation

In 2015, 195 nations have set a binding agreement, known as COP211 to keep global temperature

rise below 2 °C to prevent severe climate effects. In the same year, the planet temperature was

0.9 °C above the 20th-century average – nearly halfway to this globally agreed-on 2°C threshold.

Global temperature increase is believed to cause serious consequences, including sea level rise,

forced displacement, extreme weathers, reduced crop productivity and pandemics.2 In respect of

climate change mitigation, reducing greenhouse gas (GHG) emissions from fossil fuels which

contribute to 85.5% of global energy in 2016 (33.3% Oil, 24.1% Natural gas and 28.1 % Coal)

present the greatest potential. With the increasing energy demand from the expanding population,

along with the goal of mitigating GHG emissions, a significant amount of research in developing

renewable energy is required. Bioenergy presents the largest source of renewable energy today,

used for heat, electricity, and transport fuels.3 The used of biomass power play a key role in

decarbonizing electricity systems by replacing energy source from fossil fuels to carbon-neutral

fuels. More importantly, liquid biofuels can be used to replace petroleum-based transport fuels

without significant change of current infrastructure. It is projected that modern bioenergy could

contribute over 20 % of global final energy supply in 2030, doubling its share from 10% in 2010.4

Among all available sources for bioenergy, microalgae as the third generation of biomass have

several advantages including fast growing rate, rich lipid content, reduced land-use, synergy effect

in wastewater treatment and do not compete directly with food. The fast-growing microalgae is a

Page 17: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

2

sustainable source to produce fuels, food, pharmaceuticals, chemicals and animal feed (Figure 1-

1). However, despite these advantages biofuel production from algae still faces serious challenges

especially the high energy and economic costs of current downstream processing techniques.

Unlike terrestrial plants, microalgae are harvested as wet biomass which contains a significant

amount of water than crop-based biomass. Conventional conversion process used for terrestrial

plants involving lipids extraction and transesterification to biodiesel requires an energy intensive

drying process which increases the carbon intensity of this approach. From a life cycle assessment

of biofuel production, drying algal biomass sufficiently for conventional lipids extraction

consumes more than 90% of the energy content in the algal oils.5–7 Therefore, the future algae-to-

biofuel conversion techniques must be adaptable to wet biomass to obtain a net positive energy

balance.8

Figure 1-1. Microalgae based biorefinery integrated with related industries to produce

numerous sustainable deliverables. Reproduced from ref9, © 2010, with permission from

Elsevier.

Similar to how nature generates fossil fuels from buried biomass under high temperature and high

pressure in a geological timescale, an emerging conversion approach adaptable to wet biomass

Page 18: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

3

called Hydrothermal Liquefaction (HTL) has attracted great attention in recent years. HTL shares

the same mechanism to convert biomass into biocrude but with a duration as short as a few minutes.

At elevated temperature (200-380 °C) and pressure (5-28 MPa) with presence of water, HTL

chemically and physically cracks down large biomolecules into small fractions and simultaneously

transforms them into biocrude.10 HTL shows two prominent advantages over conventional

conversion technologies: drying treatment for feedstock is eliminated since a wet slurry is directly

input into the HTL process; and higher biocrude yields are usually found in HTL experiments

because HTL converts not only lipids but also other biomass such as carbohydrates and proteins

into biocrude.9–11 These advantages of HTL have made it a unique and promising path to produce

liquid biofuels from microalgae feedstock at a favorable energy return on investment. However,

HTL experiments have been conducted in laboratory scale batch reactors that (1) require

substantial heating times lead to ambiguity in the reported results and (2) lack of in-situ observation

of the reactions. Therefore, in order to understand the reaction mechanism of HTL and reach the

full potential of biocrude yield, precise control of processing conditions and in-situ visualization

of the reaction process are essential.

Producing high-value bioproducts from microalgae has attracted great attention in recent years due

to high processing costs and low market value for biofuels, with current low oil price further

driving this transition. Figure 1-2 indicates related microalgae products with their estimated market

value and size. Natural astaxanthin from microalgae is a highly valuable bioproduct with

tremendous health benefits and has become the primary source for the nutraceutical industry. The

extraction of astaxanthin from the microalgae is hindered by a thick cell wall that prevents solvent

extraction and digestion from direct consumption. This barrier is extremely robust to chemical and

physical disruption, making astaxanthin extraction difficult. Hydrothermal processes utilizing high

temperature and pressure have shown promise for extraction of bioactives, woody biomass

decomposition and biocrude formation which has strong potential for cell-wall disruption. In

addition, supercritical carbon dioxide (ScCO2) has attracted attention recently as a green solvent

for the extraction of bioactive compounds. The overall extraction efficiency of ScCO2 is highly

dependent on the characteristics of the feedstock, operating conditions (temperature, pressure,

duration) and addition of modifiers/co-solvents. Therefore, the development of effective cell wall

disruption techniques and understanding of the ScCO2 extraction mechanism are critical to the

design of commercial-scale astaxanthin extraction reactors and processes.

Page 19: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

4

Figure 1-2. Sunlight-to-biomass conversion efficiency and losses along the path, as well as

value and market size of related microalgae products. Reproduced from ref15, © 2016, under

CC-BY license.

Microfluidic approaches with high degree of process control have been applied for bioenergy

applications.16–18 Recent applications of microfluidics in high temperature and pressure

processing, particularly in the context of chemical synthesis19,20 and phase analysis21,22 made it a

promising technique for microalgae downstream processing. In addition to precise control on

temperature and pressure, continuously flowing microfluidic reactors can achieve extremely high

heating rates avoiding ambiguity due to long heating times. Moreover, microfluidic devices with

in-situ observation using fluorescence enables direct, real-time monitoring of processes at the

cellular level. The ultimate goals of these works are to (1) develop a high temperature and pressure

microfluidic platform with in-situ observation capability and (2) perform microalgae to biofuels

and bioproducts conversions.

1.2 Thesis Overview

This thesis is focused on the development of high temperature and pressure microfluidic platform

and microalgae downstream processing techniques particularly biomass-to-biocrude conversion

via hydrothermal liquefaction, hydrothermal disruption of cell wall, and supercritical CO2

extraction of astaxanthin. The results and the microfluidic platform on which they were collected

Page 20: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

5

represent the first of their kind in the field of microalgae downstream conversions. These

researches provide unprecedented insight into biomass to biofuels and bioproducts conversions

that will guide the design of large scale reactors and processes. Supplementary contributions

include projects which I contributed to as co-author and are included in appendices.

Chapter 1 briefly describes the research motivations of building a sustainable future by producing

energy and high-value bioproducts from microalgae biomass. This chapter discusses the key

aspects of these two downstream processes and briefly describes the demand for novel high

temperature and pressure microfluidic platform in that context.

Chapter 2 provides a background of related work in the field of hydrothermal processes for

biomass, supercritical CO2 extractions of bioactive compounds and microfluidics in general. These

concepts introduced in this chapter include processing evaluations using LCA and TEA, water

characteristics under subcritical conditions, biomass composition and conversion, supercritical

CO2 extraction and high temperature and pressure microfluidics.

Chapter 3 reviews the current microalgae biomass downstream processes and emerging techniques

mainly for bioenergy applications. This chapter is part of the review manuscript published in

Sustainable Energy and Fuels, and portions of this chapter have been reproduced with permission

from The Royal Society of Chemistry.

Chapter 4 demonstrates microalgae biomass-to-biocrude conversion on a chip via hydrothermal

liquefaction. A high temperature and pressure microfluidic platform with optical access and

precise control of processing parameters was first demonstrated in the research of HTL. This

chapter was published in Lab on a chip and featured as HOT article – reproduced by permission of

The Royal Society of Chemistry.

Chapter 5 presents the hydrothermal disruption of cell wall for astaxanthin extraction on a

microfluidic platform. Hydrothermal disruption at a temperature of 200 °C was shown to be highly

effective, resulting in near-complete astaxanthin extraction from wet biomass – a significant

improvement over traditional methods. This chapter was published in Green Chemistry –

reproduced by permission of The Royal Society of Chemistry.

Page 21: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

6

Chapter 6 presents astaxanthin extraction from Haematococcus pluvialis using supercritical carbon

dioxide with and without co-solvents. The associated manuscript is in preparation. This chapter

was submitted for publication.

Chapter 7 summarizes the author’s contribution in the downstream conversion of biomass to

biofuels and bioproducts and provides potential directions for future studies.

Page 22: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

7

Chapter 2.

Introduction

To reach a sustainable future, integrated biorefineries have been suggested as a means to produce

food, energy, chemicals and materials.23 Microalgae is particularly well positioned due to the many

benefits over other sources of biomass such as high growth rate, low land requirement, high oil

content and recycling of nutrients from wastewater. In this chapter, two major microalgae

applications are introduced, liquid biofuels and high-value bioproducts.

In context, current microfluidic devices used for high temperature and pressure reactions are

reviewed. There have been efforts to scale microfluidic methods for processing and scale

microfluidic reactors for emulsion production24–26 and the bulk production of chemicals under

challenging experimental conditions19,27–29. These reactors generally approach the challenge of

producing bulk product by 1) parallelizing (multiple parallel processes on a single chip) and 2)

numbering out (having many chips operating in parallel). There are several companies in this space

making commercial products such as Corning, Ehrfeld Mikrotechnik, Alfa Laval and Chemtrix,

who are developing modular microfluidic reactors. Other companies such as Uniqsis, Microinnova

and FutureChemistry are developing integrated continuous flow reactors for chemical synthesis

without using a chip. While some of the above-described methods may be applied to increase

output from methods developed in this thesis, production of bulk product is not the focus here.

Rather, this thesis focuses on developing and demonstrating microfluidic technologies to study

biomass conversion processes, with the goal to inform industry and the academic community by

resolving previously opaque processes at the single-cell scale. The insights herein can be applied

to improve large conversion systems already operating at scale.

2.1 Microalgae for biofuel production

With increasing energy demands resulting from accelerating population growth and global

development, the need to curb anthropogenic greenhouse gas emissions is increasingly important.

Biofuels from microalgae are both renewable and potentially carbon neutral and are therefore a

compelling alternative to fossil fuels for transportation, which accounts for 28% of global energy

Page 23: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

8

consumption30. For biofuel production, microalgae have many benefits over other sources of

biomass. For instance, microalgae can be cultivated in brackish/wastewater environments which

do not interfere with food supplies or contribute to greenhouse gas emissions through land-use

change31. Perhaps the most significant advantage is the rapid growth rate of microalgae which can

double in as few as two hours.32 Moreover, microalgae cultivation can be integrated directly into

wastewater treatment plants and flue gas producing facilities to recycle wasted nitrogen,

phosphorus and CO2 as nutrients.7,33 Although these benefits make microalgae an outstanding

candidate for biofuel production, a major obstacle in producing carbon neutral and energy

favourable biofuel is the poor efficiency of converting the raw biomass into liquid biofuels.

2.1.1 Assessments for downstream conversions

Conventional methods to produce biofuel from biomass requires the extraction of lipids from cells

which can then separately be converted to biodiesel, also known as fatty acid methyl esters

(FAME). During this transesterification process, the major components in algal lipids,

triacylglycerides (TGA) are converted to FAME through a reaction involving alcohol catalyzed by

acids or alkalis at elevated temperature. This process is mature and commonly used in the

conversion of vegetable oils to biodiesel. In this process, lipids are the only intracellular

components contributing to the biofuel yield whereas carbohydrates and proteins are unused. As a

result, the overall biodiesel yield of this process is strongly correlated to the growth rate of

microalgae, lipid content of the cell, the efficiency of extraction method, and the lipids-to-biodiesel

conversion rate. Generally, for a specific algal strain, the growth rate and lipid content have an

inverse relationship, where the oil-rich microalgae tend to grow slower than most low-oil algal

strains. Even within the same species, the proportion of individual cell constituents largely depends

on environmental conditions. More importantly, the challenge for this method is that the biomass

must first be dried, which significantly increases the carbon intensity of the final fuel as well as its

cost. A more recent conversion approach, hydrothermal liquefaction (HTL), has attracted a great

deal of attention due to its ability to directly convert wet biomass to biofuels. Other downstream

conversion methods and emerging techniques are reviewed in Chapter 3. In this section, Life cycle

Assessment (LCA) and Techno-economic Analysis (TEA) are used to evaluate different

downstream processing routes.

Page 24: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

9

Recently, a life cycle assessment (LCA) has been widely used to investigate energy balance and

greenhouse emissions on algae-derived fuels. Slade and Bauen34 reviewed several recent LCA

studies on algal biofuel production including cultivation, harvesting and oil extraction and found

that only open pond systems were able to achieve a positive net energy (Figure 2-1a). For raceway

pond systems, more than 80% of overall energy consumed was due to biomass drying and

dewatering with the exception of the LCA performed by “[11] Stephenson”. In this case the authors

assume the use of an effective oil extraction process adaptable to wet biomass which does not exist

yet. Analogous to these energy balance analyses, the carbon emission chart (Figure 2-1b) indicates

that only raceway pond systems can reduce emissions to less than 84g CO2e/MJ (similar to

petroleum-derived diesel) and the majority of those emissions are again associated with biomass

drying and dewatering. In agreement with other LCA studies, drying algal biomass sufficiently for

conventional lipids extraction consumes more than 90% of the energy content in the algal oils5–7.

Page 25: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

10

Figure 2-1. A summary of life cycle assessments for biofuel production form microalgae. a)

Net energy ratio for microalgae biomass production and b) Illustrative estimates for carbon

dioxide emissions from algal biomass production. Reproduced from ref34, © 2013, with

permission from Elsevier.

Page 26: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

11

A LCA from Sills et al.8 performed a direct comparison of the energy balance and GHG emissions

of biofuel production from microalgae using HTL and lipid extraction. The results indicated that

HTL approach generates a higher energy returns on investment (EROI) and lower global warming

potential than dry extraction methods (Figure 2-2). Although the EROI ratio of HTL is just slightly

higher than 1 in the base case, it is expected to increase over time with the development of

associated technologies. For petroleum-derived fuels the EROI used to be on the order of 100 but

has fallen to 4-5 nowadays35. It is then foreseeable that the EROI of algae-derived fuel could

surpass petroleum-derived fuel over time and the incentives to mitigate GHG emissions could

accelerate this transformation. Moreover, integrating HTL with wastewater treatments in algal

biofuel production brings additional environmental benefits into this assessment.36

Figure 2-2. a) Energy balance and b) Global Warming Potential for biofuel production using

wet and dry lipid extraction. Reproduced from ref8, © 2013, with permission from American

Chemical Society.

Page 27: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

12

Figure 2-3. Techno-economic analysis of liquid fuel produced from hydrothermal

liquefaction of woody biomass a) Effect of improvement on MFSP from state-of-technology

(SOT) case to goal case. b) Sensitivity analysis of parameter variation on the MFSP of the

goal case. a) and b) are reproduced from ref37, © 2014, with permission from Elsevier. c)

Sensitivity analysis of the MFSP of biofuel produced from defatted microalgae via HTL.

Reproduced from ref38, © 2014, with permission from Elsevier.

A techno-economic analysis (TEA) is also commonly used to guide the decision making process

when commercializing new technologies. In terms of algal biofuel production, it is crucial to select

the pathways that show the most economic promise. The Pacific Northwest National Laboratory

Page 28: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

13

(PNNL) has used a TEA to investigate the economic value of producing biofuels from woody

biomass via HTL37. Based on the laboratory scale testing results of current technologies, the

minimum fuel-selling price (MFSP) from HTL is $4.44/gallon of gasoline-equivalent value.

However, the technology improvement of HTL process alone is expected to reduce the MFSP by

$1.34/gallon or 31% of the MFSP (Figure 2-3a). A sensitivity analysis performed on an HTL

process indicates that the cost of feedstock and the efficiency of HTL process itself have the largest

effects on the MFSP (Figure 2-3b&c). With technology breakthroughs in strain selection, genetic

modification, cultivation and effective harvesting, however, it is possible that the price of

feedstocks could be dramatically reduced. Current price for dry ash free algae is estimated at

$430/ton including cultivation, harvesting and dewatering to 20 wt% dry solids39. This price is

much high than that of corn of $170/ton while the mass productivity of microalgae is about 10

times higher that of corn8. Since the cost reduction of microalgae feedstock is a foreseeable trend,

the performance of HTL as the largest influencing factor in the downstream processing is critical

to the commercialization of microalgae biofuel production.

Overall, results from TEA are generally in agreement with the LCA in suggesting that downstream

conversion processes using HTL is a promising approach due to its ability to directly convert wet

biomass. The development of this technology could then have a significant impact on the next

generation of biofuel production.

2.1.2 Water under subcritical conditions

Hydrothermal liquefaction, which directly converts biomass into biocrude at elevated temperatures

and pressure, has attracted much attention. This technique was originally called “pressurized hot

water extraction” and later known as “subcritical water extraction” because under subcritical

conditions, water is less polar, and more soluble to organic compounds. Later, water at subcritical

conditions was found to function as more than just a solvent and to classify the process better it

was divided into three categories: hydrothermal carbonization (< 200 °C), hydrothermal

liquefaction (200 - 380 °C) and hydrothermal gasification (> 380 °C). In these processes, water

plays an essential role and understanding the fundamentals of water under these subcritical

conditions (below 374 °C and 22MPa) is critical to the development of HTL technique for

producing biofuels from biomass. In this thesis, carbonization and gasification are not discussed

in detail as they are used mainly for solid and gaseous products respectively.

Page 29: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

14

Under ambient conditions water is a polar solvent excellent at dissolving polar compounds and

salts. As temperature increases, the characteristics of water change substantially. In HTL, water is

not considered an inert medium but an active participant in the reaction, either as a reactant or

catalyst. The characteristics of water and the role of water in high-temperature reactions are well

reviewed by Akiya and Savage.40 At subcritical conditions, the dramatic changes in the properties

of water are mainly displayed in the changes to: 1) hydrogen bonding, 2) dielectric constant, 3)

ion product, and 4) density.

The strong hydrogen bonding of water is the source of its many unique properties. With increasing

temperature and decreasing density, the hydrogen bonding in water becomes weaker and less

persistent. For example, water at 300 °C and 0.5 g/cm3 retains only 30-45 % of the hydrogen

bonding that exists at ambient conditions.41 In addition, the hydrogen bond network in water at

high temperature exists in the form of small clusters, rather than the infinite percolating network

of hydrogen bonds found in ambient liquid water. In general, the average cluster size decreases

with increasing temperature and decreasing density which leads to a high “local concentration” of

H+ and OH- ions42 that enhance ionic reactions.

The dielectric constant (relative permittivity) is the ratio of the permittivity of a substance to the

permittivity of free space, which indicates a relative measure of its chemical polarity. The density,

static dielectric constant and ion dissociation constant (Kw) of water at 30 MPa as a function of

temperature is shown in Figure 2-4. In general, the static dielectric constant of water decreases

with increasing temperature and decreasing density. For example, the dielectric constant of water

decreases from 78.85 to 19.66 as temperature increases from 25 °C to 300 °C, resulting in water

molecules changing from highly polar to fairly nonpolar.43 As a result, small organic compounds

are highly soluble in subcritical water and completely miscible in supercritical water.44,45 In HTL,

water can dissolve organic compounds and enhance reaction with organic compounds.

Page 30: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

15

Figure 2-4. Density, static dielectric constant and ion dissociation constant (Kw) of water at

30 MPa as a function of temperature. Reproduced from ref46, © 2008, with permission

from Royal Society of Chemistry.

The ion dissociation constant (Kw) is defined as the product of the concentration of H+ and OH- in

water and it also dramatically increases with the increasing temperature to a maximum of about

10-11 mol2/kg2 at around 300 °C (Figure 2-4). For the temperature in the HTL range (200-380 °C),

water is an effective medium for both acid- and base-catalyzed reactions due to the high

concentration of H+ and OH- ions. These effects are further enhanced due to smaller water cluster

sizes resulted from aforementioned weaker hydrogen bonding.42

Water density also varies greatly with temperature and pressure. The decrease in water density

contributes to chemical reactions mainly in two ways: 1) increase in diffusivity of water40 and 2)

increase in ion dissociation constant47. The increase in ion dissociation constant and diffusivity

enhance reactions by increasing the concentration of ions and mass transport, respectively. With

the change in density from 1 to 0.1 g/cm3, the diffusivity increases by an order of magnitude. These

changes in transport properties can affect reactions influenced by diffusion time scales and solvent

dynamics.

Page 31: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

16

The dramatic change in the physiochemical properties of water at high temperature suggests the

promising path of converting biomass to biofuels via hydrothermal liquefaction. In addition,

subcritical water can serve as a hydrogen source in reactions which can greatly affect the product

distribution.40 In summary, water at high temperature has the ability to carry out a unique set of

chemical reactions and it may help to explain why HTL is more likely to convert biomass into oil

than pyrolysis. Since water is the cheapest and most environmentally benign solvent and is present

with the microalgae feedstock, HTL of biomass with other additives such as alcohols and catalysts

are avoided and discussed elsewhere.48

2.1.3 Characterization of microalgae and post-HTL products

Characterizing the physical and chemical properties of microalgae is critical to the design of

appropriate downstream processes as well as an understanding of reaction mechanism in HTL.

Also, classifying and quantifying the post-HTL products is essential for the understanding of HTL

processes. In this section, major algal components and analytical methods to quantify them will be

introduced.

The major organic components of algae including microalgae and macroalgae are lipids, proteins

and carbohydrates. Unlike woody biomass, most species of algae do not contain lignin. This is

considered a benefit for algae because lignin is relatively resistant to chemical degradation and

produces a significant amount of solid residue in HTL.49 Lipids are non-polar aliphatic compounds

namely triacylglycerides which is the main screening criterion in biodiesel production.32 However,

algae with high lipid contents usually have slow growth rates and low biomass productivity.50

Since HTL converts all organic components of biomass into biocrude, rather than being purely

limited to lipids, the algae used for HTL can have low lipids content as indicated in Figure 2-5.

Nevertheless, high-lipid algae tend to have higher biocrude yields because the conversion rate

follows the trend of lipids > proteins > carbohydrates.12,51 Determination and quantification of

lipids are commonly performed through solvent extraction followed by Gas Chromatography –

Mass Spectroscopy (GC-MS). For lipids extraction, a modified Bligh-Dyer extraction method

utilizing chloroform and methanol is preferred due to the higher extraction efficiency as oppose to

Soxhlet extraction, which utilizes only non-polar organic solvent such as hexane. Acid hydrolysis

of lipids to produce isolated fatty acids is usually used prior to the GC-MS.

Page 32: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

17

Figure 2-5. Characteristics for algae used for HTL. The green, black and red dashed arrows

indicate the mass fraction of lipids, proteins and carbohydrates, respectively. Reproduced

from ref10, © 2014, with permission from Elsevier.

Proteins are major components of algae and are the primary component containing sulfur and

nitrogen. The proteins content in the dry biomass is therefore estimated based on the nitrogen

composition. A thermogravimetric analysis (TGA) based on continuous measurement of biomass

can be used to determine the moisture and ash content before HTL. The first-stage mass loss (up

to 110 °C) generally refers to the moisture content of the mix while the mass leftover at

temperatures greater than 900 °C is considered as ash. Dry biomass generated after moisture

removal is subjected to an elemental analysis to determine the amounts of constituent elements:

carbon (C), hydrogen (H), nitrogen (N) and sulfur (S). In the elemental analysis, a complete

combustion process with sufficient oxygen is used to convert biomass into gases such as CO2, SO2,

N2 and NOx for quantification. Oxygen (O) content is usually calculated by difference (%O =

100% − %C − %H − %S − %N − %𝐴𝑠ℎ ). The protein contents are estimated from the

nitrogen content with a conversion factor of 6.25.52

Page 33: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

18

The carbohydrates in biomass are mainly polysaccharides, starch, cellulose and hemicellulose. The

carbohydrate contents in biomass are determined by subtracting the lipid, protein and ash content

from 100%.53 Trace amounts of minerals are also contained in biomass and can be determined by

inductively coupled plasma atomic emission spectrometer (ICP-AES).

Figure 2-6. HTL of microalgae procedure and product separation process. Reproduced from

ref54, © 2012, with permission from Elsevier.

The previously described analysis techniques describe the contents of wet biomass. It is similarly

important to be able to quantify the outputted products after biomass has undergone HTL. These

post-HTL products consist of liquid products, gaseous products and solid residue. The portion of

liquid products that is dissolvable in dichloromethane is defined as biocrude and the rest is

considered as aqueous products. Within the biocrude, the portion that is dissolvable in hexane is

considered as light biocrude while the remaining is heavy biocrude. The separation of each

component is shown in Figure 2-6. Biocrude is a complex mixture. The easiest and most interesting

characteristic of biocrude is the higher heating value (HHV) which is estimated from the elemental

composition using a modified Dulong’s formula.10,54 A more detailed analysis of biocrude can be

performed using GC-MS and a large number of compounds are found mainly consist of cyclic

nitrogenates, cyclic oxygenates and cyclic nitrogen and oxygen compounds.55 Products in the

aqueous phase have been quantified using elemental analysis on water-evaporated residue.

Gaseous products are also quantifiable using gas chromatography with a thermal conductivity

detector (TCD).56 The solid residue is separable by centrifuge and mainly contains ash and

inorganics.

Page 34: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

19

2.1.4 HTL of biomass and effects of processing conditions

Hydrothermal liquefaction has been studied on multiple species of algae and woody biomass as

well as individual biomass components such as lipids, proteins and carbohydrates. In general, the

biocrude yield varies significantly (20% to 87%) with different types of feedstock and processing

conditions10, but follows the trend according to the composition in biomass: lipids > proteins >

carbohydrates (Figure 2-7).12,51,57

Figure 2-7. Yields of products from hydrothermal liquefaction of different types of feedstock

in: a) water, b) sodium carbonate, c) formic acid. Reproduced from ref12, © 2011, with

permission from Elsevier.

Page 35: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

20

Lipids can be readily hydrolyzed by HTL at greater than 90% conversion rates to produce fatty

acids that are relatively stable in subcritical water58. King et al. obtained free fatty acid yields of

90-100 % from hydrolysis of soybean oil in 10-15 min with subcritical water at 330-340 °C and

13.1 MPa.59 Proteins are polymers of amino acids that are linked through peptide bonds connecting

carboxyl and amine groups. In HTL, proteins are hydrolyzed slowly to produce amino acids which

will further be degraded in subcritical water via decarboxylation and deamination reactions.60 The

hydrolysis rate of proteins and the decomposition rate of amino acids highly depends on the

reaction temperature. Rogalinski et al. performed hydrolysis of bovine serum albumin and found

the amino acid decomposition rate is higher than the hydrolysis rate at temperatures above 250 °C

with the highest amino acid yield obtained at 290 °C in 65 s.61 HTL of several carbohydrates have

also been well studied in the past including monosaccharides, such as glucose and fructose, and

polysaccharides, such as starch and lignocellulose.60 The destruction of monosaccharides in

particular is drastic under hydrothermal conditions.62 For instance, Kabyemela et al.63 observed a

55% conversion of glucose after 2 s at 300 °C and 90% conversion after 1 s at 350 °C.

Polysaccharides including starch, cellulose and hemicellulose are fundamentally polymers of

monosaccharides suggesting that the HTL process consists of two stages: polysaccharide

depolymerization and monosaccharide degradation. Starch and cellulose are polymers of glucose

that are linked by different bonds. Rogalinski et al.64 found that 100 % conversion of cellulose was

achieved within 2 min at 280 °C while starch hydrolysis was faster than cellulose hydrolysis in

hydrothermal conditions. Hemicellulose is a heteropolymer composed of monosaccharides

including xylose, mannose, glucose, galactose and others. It was found that nearly 100% of

hemicellulose could be hydrolyzed at 230 °C and 34.5 MPa in 2 min.13 The timescales of the above

HTL of carbohydrates were mainly over a span of a few minutes, suggesting that rapid heating is

crucial to study the specific effects of hydrolysis and degradation under hydrothermal conditions.

Page 36: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

21

Figure 2-8. Biocrude yield of HTL as a function of heating rate. a) Temperature profiles with

corresponding biocrude yields from HTL of Nannochloropsis sp. for different set-point

temperatures of 300, 500, 400, and 600 °C. Reproduced from ref65, © 2013, with permission

from American Chemical Society. b) The effect of heating rate on the products yield from

HTL of macroalgae. (T = 350 °C; holding time = 15 min; and biomass/water ratio = 1/10,

w/w). Reproduced from ref66, © 2014, with permission from Elsevier.

Page 37: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

22

Several research sources indicate that biocrude yield is tightly related to the processing

conditions67,68 such as temperature, pressure, residence time and especially heating rate65,66 (Figure

2-8). Under current system optimizing the many parameters involved in HTL remains a challenge.

For instance, Faeth et al.65 reported that their highest biocrude yield (66 wt%) from HTL was

produced when operating at their maximum heating rate of 230 °C/min for a duration of 1 min,

meaning the temperature over the course of the experiment was constantly changing. Higher

heating rates were not investigated due to the limited heating capability of their experimental setup.

To date, HTL experiments have been conducted in laboratory scale batch reactors using sandbaths,

ovens or heating coils for temperature control. These approaches have two critical issues. First,

due to the large fluid volumes and large physical size of the apparatus, substantial heating times

are required, which are often ignored in the subsequent analyses and reported reaction times. These

long heating times lead to ambiguity in the reported optimal values for temperature, heating rate,

and reaction time since the heating and pressurization delays blur the results. Second, in situ

monitoring of the reaction process is not possible in current reactors, precluding real-time

quantification of the reaction process. In order to understand the reaction mechanism of HTL and

reach the full potential of biocrude yield, precise control of processing conditions and in-situ

visualization of the reaction process are essential.

2.2 Microalgae for high-value bioproducts

The potential commercialization of microalgae mainly falls into two categories: 1) large-volume,

low-value bulk commodities such as biofuels and food/feed and 2) low-volume, high-value special

products such as pigments, health care and additives. Although HTL shows promise in biofuel

production, further development of this technology is required to lower the production cost to

compete with fossil fuels. On the other hand, the production of high-value products from

microalgae could be currently profitable69 and the maturation of associated technology will also

benefit the development of biofuel production. The market value of microalgae products is

indicated in Figure 2-9.

Page 38: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

23

Figure 2-9. Market value of the microalgal components and total selling price of the

biomass for different market scenarios. Reproduced from ref69, © 2016, under CC-BY

license.

Among all high-value products, natural astaxanthin from microalgae – Haematococcus pluvialis

is one of the most promising candidates showing high market value and large market volume.

Astaxanthin is a strong antioxidant, a pigment and has important applications in the nutraceuticals,

cosmetics, food and aquaculture industries.70,71 However, critical bottlenecks and major challenges

also exist in the commercial scale production of astaxanthin especially in astaxanthin recovery

from biomass. The accumulation of astaxanthin is strongly associated with the formation of a rigid

cell wall that prevents solvent extraction. Also, for astaxanthin extraction, a chemically inert and

greener solvent is required in terms of product quality and environment concerns. Therefore,

development of effective cell-wall disruption and ScCO2 extraction techniques are crucial to the

commercial implementation of astaxanthin production at large scale.

2.2.1 Cell structure of Haematococcus pluvialis

H. pluvialis is a freshwater unicellular green microalgae with four distinguishable stages in the life

cycle: macrozooid (during cell division), microzooid (after germination), palmella, and cyst

(aplanospore) (Figure 2-10b). The first three stages are generally referred as green vegetative phase

((A) and (B) in Figure 2-10a) and the cyst (aplanospore) stage is called red phase ((C) and (D) in

Figure 2-10a) suggesting an accumulation of astaxanthin in the cell.

Page 39: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

24

Figure 2-10. a) Microscopic images of Haematococcus pluvialis in life cycle. (A) Green

vegetative motile cell; (B) Green vegetative palmella cell; (C) Astaxanthin accumulating

palmella cell in transition to aplanospore; (D) Astaxanthin accumulated aplanospore cell.

Scale bar: 10 μm. Reproduced from ref72, © 2016, under CC-BY license. b) Illustration of

life cycle of H. pluvialis. Reproduced from ref73, © 2013, under CC-BY license.

Harvesting of astaxanthin is performed on mature aplanospores resulting in the maximum amount

of astaxanthin during the life cycle (2-5 % of dry weight). At this stage, cells become resistant to

extreme environmental conditions and show two distinct structures: 1) a thick and right trilaminar

sheath and 2) a secondary cell wall. The development of the rigid cell wall and its structure during

the life cycle is shown in Figure 2-11. Hagen et al.74 reported the aplanospore cell wall contained

70 % carbohydrates (66% hexoses), 3% cellulose, 6% proteins, and 3% acetolysis-resistant

material. Due to the fast reaction rates of carbohydrates under hydrothermal conditions, subcritical

water could be an effective solution to disrupt cell walls.

Page 40: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

25

Figure 2-11. Summary of five developmental states of the cell wall during aplanospore

morphogenesis in Haematococcus pluvialis: I, 1-week-old flagellates ; II, flagellates at least

2 weeks old just rounding off; III, 2- to 3-week-old aplanospores; IV, at least 3-week-old

aplanospores; V, aplanospores in their final state. CYP, cytoplasm; IS, interspace; PL,

plasmalemma; PW, primary wall; SV, secretory vesicles; SW, secondary wall; TCL,

tripartite crystalline layer; TLS, trilaminar sheath; W1±W7, layers of the extracellular

matrix. Reproduced from ref74, © 2002, with permission from Taylor & Francis.

2.2.2 Cell wall disruption and extraction techniques

Astaxanthin accumulates inside a thick, rigid cell wall of H.p cysts and the performance of

recovery is highly dependent on the cell-wall disruption and subsequent solvent extraction

processes. Intact cysts of H.p without any cell-wall disruption treatment only releases ∼20% of

the internal astaxanthin using an acetone extraction solvent over a 16 hour period.75 Both chemical

disruptions using acid or base and biological disruption using enzyme followed by acetone

extraction recovered 25-40 % of astaxanthin, a slight improvement over the control. Therefore,

having an effective cell-wall disruption treatment is critical to the subsequent solvent extraction

process. Kim et al.76 recently reviewed different astaxanthin recovery processes and separated

them into four groups based on cell-wall disruption methods: chemical, physical, physico-

chemical, and biological methods. Physical disruptions use grinding, bead beating and French-

Page 41: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

26

pressure-cell while physico-chemical disruptions are a combination of physical (bead beating,

grinding, milling or ultrasound) and chemical (organic solvent, acid or breaking buffer) processes.

They are reported with higher yields than chemical disruptions but there is ambiguity in the

comparison due to three issues: 1) most physical disruptions use dry biomass which might be

processed differently; 2) different solvents are used in the subsequent extraction processes such as

methanol/dichloromethane mixture, acetone and ethyl acetate; 3) recovery rates were reported as

mg/g cell which is strongly influenced by the feedstock as opposed to the percentage of total

astaxanthin extracted. Therefore, a platform that can quickly and visually assess the effectiveness

of different cell-wall disruption methods and extraction conditions is desired to provide a fair

comparison and insights into this important process.

Among all extraction solvents, supercritical carbon dioxide (ScCO2) has attracted much attention

as a green solvent for the extraction of bioactive compounds.77,78 Supercritical CO2 features three

major advantages over organic solvents: 1) it is abundant and benign to human health and the

environment, 2) a solvent-free extract can easily be produced and extraction by CO2 evaporation

at room temperature and pressure, 3) bioactive compounds are well-preserved due to the inert

chemical property of CO2 and relatively low critical temperature (31.1 °C). However, the overall

extraction efficiency of ScCO2 is highly dependent on the characteristics of the feedstock75,76,79,80,

operating conditions (temperature, pressure, duration)81,82 and addition of modifiers/co-solvents83–

85. Understanding the extraction mechanism of ScCO2 with and without co-solvents is critical to

the development of astaxanthin production from microalgae. In addition, to avoid extremely high-

pressure operation while remaining the high extraction rate and efficiency, a platform with optical

access is desired to bring insights into the extraction mechanism and the relationship of co-solvents

to ScCO2.

In summary, the extraction efficiency, extraction rate, scalability along with the energy

consumption should be considered in the development of cell-wall disruption and solvent

extraction techniques for industrial-scale astaxanthin production form H.pluvialis biomass.

2.3 High temperature and pressure microfluidics

Microfluidic technologies have been used widely in chemical synthesis, health care, fundamental

physics and bioenergy production mainly due to the following advantages: 1) they require less

reagent as compared to bulk systems; 2) they provide a high degree of control over processing

Page 42: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

27

parameters; and 3) they provide safer working environments when performing dangerous

chemistry.19,86 In addition, microfluidic reactors have often claimed to produce higher yields than

bulk reactor due to the scale-dependent processes of heat and mass transfer and large surface-area-

to-volume ratio. As a result, there have been some examples of people using parallel microfluidic

reactors (numbering up) to increase productivity rather than increasing the characteristic

dimension of the channel (scaling up). Given the low volume size of microfluidic channels but

high throughput on information from a rapid screening of reaction conditions, the actual product

of most microfluidic reactors is the information.

Figure 2-12. Examples of (a) metal, (b) glass and (c) silicon/glass microreactors. Reproduced

from ref20, © 2011, with permission from Elsevier.

The advantages of using microfluidic reactors are even more pronounced in chemical synthesis

involving harsh process conditions such as high temperature and pressure.87,88 Commonly used

microfluidic reactor made of polymers such as polydimethylsiloxane (PDMS) and

polymethyacrylate (PMMA) cannot withstand high temperature and pressure. Metal-based

microreactors fabricated using conventional machining are good for harsh condition processing

but are limited by the lack of optical access. Glass microfluidic reactors have the advantage of

utilizing in-situ optical characterization techniques but are limited by the low thermal conductivity.

Microfluidic reactors fabricated out of silicon and glass have advantages in both high temperature

and pressure reactions by allowing: 1) optical access through the glass, 2) high thermal

Page 43: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

28

conductivity through the silicon, and 3) both high pressure and temperature. Examples of different

microfluidic reactors are shown in Figure 2-12.

Silicon/glass microfluidic reactors have been used for materials synthesis88, fuel conversions21,

phase behavior studies89 and supercritical fluid extractions90. The silicon/glass chip is fabricated

by silicon etching using solutions or deep reactive ion etching (DRIE) followed by anodic bonding.

By matching the thermal expansion coefficient of glass to silicon, the chip can withstand high

pressure at various temperatures. The high-pressure sealing is achieved using modular packaging

at room temperature by cutting a silicon window to isolate the high-temperature region. Achieving

high temperature and pressure conditions in microfluidic reactors suggests promising applications

of water under hydrothermal conditions and supercritical fluid extractions.

Page 44: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

29

Chapter 3.

Current Downstream Processing and

Emerging Technologies for Microalgae

This chapter is a portion of a review published in Sustainable Energy and Fuels and reproduced

from ref91 – reproduced by permission of The Royal Society of Chemistry. The candidate was one of

the equally contributing authors in this work and played the primary role in reviewing the emerging

techniques in downstream processing. Additional authors for the work include Dr. Scott C.

Pierobon, Dr. Percival Graham, Mr. Brian Nguyen, Mr. Evan G. Karakolis, and Prof. David

Sinton.

3.1 Optimizing downstream processing

Unlike terrestrial plants, microalgae are harvested as wet biomass which requires different

downstream processing from crop-based biomass. To complete the microalgae-to-biofuel

conversion route, pretreatment of biomass such as harvesting, thickening and dehydration are

typically applied to simplify conversion processes92–96. However, given the high latent heat

vaporization for water (2,265kJ/kg) the energy intensity and associated cost of drying is a

fundamental challenge for microalgal biofuels. Given the fact that most microalgae have a total

energy content on the order of 18,000 kJ/kg97,98, drying a microalgae slurry at a concentration

below ~20 wt% (which contains ~3600 kJ/kg of energy) is of limited practicality for biofuel

application. As such, the strategy to operate a microalgae operation is to either avoid massive

drying or produce high-value products to offset the high processing costs. In this section, we will

first discuss the dewatering and drying processes, then describe how dry biomass, wet biomass,

microalgae in culture and direct secretion can be implemented. Microalgae in culture involve

directly using cells, wet biomass is typically concentrated from harvested cells and dry biomass is

Page 45: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

30

obtained by drying wet biomass. Each area will be briefly reviewed while highlighting emerging

technologies.

3.2 Harvesting and dewatering processes

Initial harvesting is typically accomplished by either screening or sedimentation, where the

concentration is increased from 0.1% of total suspended solids (TSS) to a slurry of about 2-7%

TSS95,96,99,100. In terms of screening, microstraineres and vibrating screens are the most commonly

used. Standalone sedimentation by gravity is highly energy efficient for large cells99,101 but is

unsuitable for many types of microalgae. Sedimentation can be enhanced by forming larger

aggregates of cells, typically by chemical flocculation and coagulation. Ultrasound-assisted

harvesting forces cell aggregation at acoustic nodes.102 This technique is free of chemical additives,

operated continuously and has the flexibility of adopting different types and sizes of cells by

adjusting the operating parameters such as acoustic energy density, contrast factor, and ultrasound

frequency. Lab-scale experiments have demonstrated a concentration factor of 11.6 can be

achieved at a flow rate of 25 mL/min.102 Other initial harvesting techniques includes air

flotation103,104, and electric field assisted harvesting95,96. As discussed, biofilm cultivation105,106 has

unique harvesting advantages but further development is required to make them cost and energy

effective for biofuel production.

3.2.1 Biomass thickening

After initial harvesting, thickening techniques can be applied to concentrate the algae suspension

to above 15% of TSS95,96,99. Two major thickening techniques are centrifugation and filtration.

Thickening saves a significant amount of energy compared to directly drying the harvested

biomass6,107,108 but requires extra capital investment. The energy cost for this process depends on

the characteristics of the cells, system design, and desired output concentration. Moreover, as the

Page 46: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

31

desired output concentration increases, the energy cost associated with incremental percentage of

dry biomass climb steeply97,108. Centrifugation is commonly used in the lab and provides rapid

water removal but it is very energy intensive, thus not practical at industrial scale for bioenergy

applications, but is widely used to produce high-value products such as pigments, polyunsaturated

fatty acids (PUFAs), phycobiliproteins, enzymes and toxins96,109,110. Filtration is conceptually

simple but potentially difficult to operate due to the two major issues: 1) the pore size needs to be

small to increase efficiency but not too small to cause clogging issues and reduce flow rates; 2)

easy recovery of algal biomass from the filter is desired without using the backwash which will

lead to re-dilution of the product95,97,111. Advantages of alternative filtration techniques such as

dead-end filtration, tangential flow filtration, cross-flow membrane filtration, axial vibration

membrane filtration and ultrafiltration have been demonstrated for specific applications95,99,112–114.

3.2.2 Dehydration to dry biomass

Dehydration is used to achieve the final high level of biomass concentration so that conventional

extraction methods and infrastructure used for terrestrial plants can be used effectively. Typical

dehydration techniques include solar drying, spray drying, freeze drying and belt drying. Solar

drying either directly uses sunlight, or converts sunlight into heat for drying. This approach is

relatively easy to implement but it is hindered by the lack of control (overheating, weather

dependency), long drying time and possible biomass degradation100,115–118. Spray drying is

commonly used to dry microalgae for food production but it still requires a significant amount of

energy which make it impractical for biofuel production99,118. Freeze drying is used to produce

pharmaceuticals from microalgae to maintain the chemical. physical, and biological characteristics

of intracellular components119–121. However, this process is far too costly and energy intensive to

be applicable to less valuable chemical outputs. Belt drying transports biomass on a conveyor belt

and dehydrate the biomass either by passing through a heating chamber or by the heat conducted

Page 47: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

32

from the belt. Other drying techniques such as rotary drying, cross-flow drying, vacuum shelf

drying and flash drying also been used in the industry but they haven’t reached commercial scale

for biofuel production due to high energy consumption118,122.

3.3 Processing for dry microalgae

Once microalgae biomass is completely dried through dehydration techniques, conventional

physical disruption followed by chemical extraction and conversion techniques can be directly

applied. Effective physical disruption can offset the need for harsh processing conditions required

in the process of chemical extraction93,110,118. Cell homogenization and bead milling are commonly

used in the industry, and emerging disruption techniques such as microwave, pulsed electric field

and ultrasonic are still under development101,118,123. Among these techniques, microwave-assisted

extraction (MAE) has attracted lots of attention due to its advantage of easy operation, high energy

transfer efficiency, rapid heating and relatively low cost. MAE has been investigated to be highly

effective for pigments124 and lipids125 extraction from microalgae. MAE of fucoxanthin from

Cylindrotheca closterium for 5 min enabled maximal extraction equivalent to 60 min of

conventional solvent extraction method.124 With the addition of ionic liquids, MAE has been

applied to extract lipids from wet microalgal biomass where extraction rates were increased by an

order of magnitude in most cases.126 However, the cost associated with these disruption techniques

is still too high for biofuel production. The requirements for physical disruption depend on the

characteristics of the microalgae and specific extraction techniques could be simplified, or even

eliminated, by genetic engineering or extraction process enhancement.

3.3.1 Lipids extraction

The most commonly used lipids extraction technique for dry microalgal biomass is organic solvent

extraction which is found to be highly effective and cost efficient101,123. Supercritical CO2

Page 48: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

33

extraction is not suitable for biofuel production due to its high initial and maintenance cost, but its

application has increased in the nutraceutical and biochemical industry in recent years78,85,127,128.

The major advantage of using supercritical CO2 as solvent is to produce contamination-free

products and the associated cost can be offset by the high marginal value bioproducts. Other

extraction techniques such as accelerated solvent, two-phase solvents and switchable-solvent

extractions have been proposed and demonstrated to improve the extraction efficiency of

intracellular components from microalgae129–131.

3.3.2 Extracted lipids and other components

Lipids extracted from microalgae can be further converted to biodiesel through several pathways:

chemical transesterification, enzymatic conversion and catalytic upgrading. Among these,

chemical transesterification is relatively mature and has been used widely for biodiesel production

with conversion efficiencies above 90%132–134. However, algal oils are typically highly complex

in contrast with vegetable oils, containing fatty acids, phospholipids, carotenoids, chlorophyll and

other components in various composition97,135. Therefore, in order to accelerate this reaction,

obtain higher conversion rate and reach the full potential of commercialization, a full

understanding of strain specific oil composition is required. In contrast, enzymatic conversion uses

biological catalyst (lipases) instead of acids or bases to convert lipids into biodiesel with less

processing energy, easier removal of glycerol and catalysts and less alkaline wastewater pollution.

Although, enzymatic approaches have advantages over conventional methods, other challenges

associated with the cost of lipase, operational life, tolerance of the environment for enzymes and

harvesting strategies for products need to be addressed for this path to run at a commercial scale136–

138. Catalytic upgrading of algal lipids into renewable gasoline, jet fuel and diesel can also be

achieved by hydrotreating processes which react oils with hydrogen at high temperature and

pressure in the presence of catalysts139–141. This conversion process is commonly used in the

Page 49: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

34

petroleum industry to upgrade crude oil to produce a wide multitude of performance specified

fuels. Ideally, this process leverages existing techniques and infrastructure albeit with catalysts

and process parameters tuned for algal feedstocks.

Recent research in algal downstream processing, has shifted from a lipid-centric approach to a

more holistic approach. If one can obtain value from all feedstock components the

commercialization of microalgae technology will be more economic viable. For example, residual

carbohydrates and proteins from microalgae can be used to enrich the nutrition in animal feed or

further processed to produce biofuels through other conversions such as fermentation, anaerobic

digestion and hydrothermal liquefaction. Another approach to achieve commercially feasibility in

microalgae industry is to produce high-value bioproducts.69,142–145. These bioactive components

present a great interest in pharmaceutical, cosmetic and nutraceutical industries and can easily

offset the high costs associated with biomass processing. The understanding of microalgae and

development of technology gained in this approach will accelerate the accomplishment of

producing sustainable energy from microalgae.

3.3.3 Processing of whole dry microalgae

Without extracting lipids, direct combustion and pyrolysis have been investigated to directly

produce energy from whole algae. Direct combustion of dry microalgae biomass releases the

largest amount of energy. However, multiple life cycle assessments5,6,8 indicate drying the

microalgae into powders costs more energy than the biomass contained, plus the challenges

associated with ash content and emission control make direct combustion of microalgae is

impractical on a commercial scale146–148. Moreover, the energy released from direct combustion is

in the form of heat which is less desirable than liquid biofuels used for transportation. Pyrolysis is

a thermal decomposition process in absence of oxygen at high temperature (300 °C – 1000 °C),

Page 50: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

35

converting organic components into a wide range of products with hydrocarbon rich liquid (bio-

oil) as the desired product.149–151 Recent research152 indicated that fast or flash pyrolysis of

microalgae with reaction time of 2-3 s, reaction temperature around 500 °C with heating rate of

600 °C/s is capable of achieving 18 – 24 % liquid yields with higher heating value (HHV) of 29

MJ/kg. Pyrolysis bio-oil from microalgae has lower oxygen content and viscosity and higher

heating value (HHV) than that from woody biomass but still requires extensive refining for use in

conventional fuel engines. More importantly, similar to direct combustion, the major roadblock in

pyrolysis is removing the moisture content in the biomass which cause the overall process to be

energy negative.

3.4 Processing for wet microalgae biomass

Avoiding the energy and cost intensive thickening and drying processes dramatically alleviates the

energy burden for microalgae biofuel applications. Two promising conversion processes

applicable to wet biomass at a concentration above 15% TSS are hydrothermal liquefaction (HTL)

and supercritical water gasification (SCWG). A summry of extraction and conversion techniques

for different concentration of microalgal biomass is shown in Table 2-1.

Page 51: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

36

Table 3-1 Summary of extraction and conversion techniques for different concentrations of

microalgae biomass.

Feedstock Type Conversion/Extraction

Techniques

Main Products Reference

Dry powder

Mechanical disruption and

chemical extraction

Antioxidants, pigments,

PUFA, additives 153,154

Direct Combustion Heat 155

Pyrolysis Bio-oil 152

Microalgae slurry

(> 15% TSS)

Hydrothermal Liquefaction Biocrude 7,65,156,157

Supercritical water

gasification Syngas 158–160

Microalgae in culture

(< 10% TSS)

Fermentation Bioethanol 161–167

Anaerobic digestion Methane 168–172

Direct secretion Hydrogen, alcohols or

alkanes 173–175

Microbial fuel cells Electricity 176–184

3.4.1 Hydrothermal liquefaction

HTL is one of the most promising conversion pathway for biofuel production due to the ability to

employ raw wet feedstock, a fast reaction rate, a high energy return on investment, good

characteristics of biocrude, a low production of char, and the potential for recycling of nutrients

(Fig 3-1a&b). Similar to natural formation of petroleum-based fossil fuels, HTL converts biomass

into biocrude at high temperature (200-380 °C) and pressure (5-28 MPa) in the presence of water,

albeit on a timescale of minutes by chemically and physically cracking down large biomolecules

into small fractions7,10,65,156,185. Compared to other thermochemical conversion techniques such as

pyrolysis, the higher heating value of HTL biocrude is about 35 MJ/kg, much higher than typical

pyrolysis bio-oil with a value of 20-25 MJ/kg186.

A fundamental challenge in HTL of biomass is optimizing the processing conditions (temperature,

pressure, residence time and heating rate) to obtain optimal yield and efficiency10,12,187. To this

Page 52: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

37

end, HTL65,66,188 with a fast heating rate was investigated and resulted in higher biocrude

productivity. However, the heating rate in this experiment was limited at 230 °C/min which was

not able to explore the peak productivity with respect to heating rate. By eliminating the limits

from conventional batch reactors, microdevices156 that allow direct observation of HTL (Figure 3-

1a) have been developed to precisely control processing parameters and perform real-time

monitoring. The results indicate the higher heating values of biocrude approached saturation within

1 min due to early mechanical disruption of cells that enabled solvent extraction. Therefore, the

HHV alone does not present the quality of biocrude whereas chemical composition and physical

characteristics should also be considered.

A sequential HTL process189 was also performed to maximize the biocrude productivity and

minimize the bio-char formation. A low temperature (140 – 200 °C) HTL was used as the first step

to disrupt the cell and release intracellular products, polysaccharides in this case. The second step

utilize a higher temperature (220 – 300 °C) HTL that converts the remaining biomass to biocrude

which resulted in overall higher biocrude productivity and less bio-char. Lower polysaccharide

content resulting in decreased bio-char production also agrees with HTL of individual categories

of biomass feedstocks which indicates polysaccharides generates more solid products than protein

and lipids.12 Similar to HTL, Hydrothermal carbonization (HTC) uses reaction temperatures in the

lower range (180 – 250 °C) with slightly elevated pressure (2 – 10 MPa) which tends to produce

more solid products instead of liquid products190–192. A recent comparison of torrefaction and HTC

of lignocellulosic biomass indicates that HTC as the wet torrefaction method is the more favorable

process in terms of energy content, hydrophobicity, and inorganic components of the solid

products193.

Page 53: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

38

Figure 3-1. Summary of current downstream processing techniques for wet microalgae biomass

(a-f) and cells in culture (g-j). (a-b) Hydrothermal liquefaction156,157. Image (b) has been

reproduced with permission from Elsevier157, Copyright 2013. (c) Supercritical water

gasification159. Reproduced with permission from Elsevier, Copyright 2016. (d-e) Ionic liquid

treatment for wet extraction194,195. Image (d) has been reproduced with permission from Royal

Society of Chemistry194, Copyright 2014. Image (e) has been reproduced with permission from

Royal Society of Chemistry196, Copyright 2015. (f) Astaxanthin extraction for hydrothermal

disrupted cells197. (g) Fermentation of pretreated wet biomass161. Image reproduced under CC-

BY license, Copyright 2014. (h-i) Anaerobic digestion of algal biomass169,170. Image (h) has been

reproduced with permission from Elsevier170, Copyright 2017. Image (i) has been reproduced

with permission from Elsevier169, Copyright 2016. (j) Microalgae-microbial fuel cells198.

Reproduced with permission from Elsevier, Copyright 2015. This figure is reproduced from ref91

with permission from The Royal Society of Chemistry.

Page 54: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

39

3.4.2 Supercritical water gasification

Analogous to HTL, SCWG reacts at higher temperature (400 - 700 °C) and pressure (> 22 MPa)

to decompose biomolecules to produce syngas containing hydrogen, carbon monoxide and

methane with a small quantity of solid and liquid products158–160,199(Figure 3-1c). Gasification is

commonly combined with Fischer-Tropsch Synthesis (FTS) to convert syngas into liquid fuels200–

202. The major advantage of this pathway is the flexibility to produce a wide variety of fuels and

products with known properties. Recent research indicates the issues for developing this process

are: precipitation of inorganic salts, an unclear reaction mechanism associated with reaction

temperature, pressure, heating rate and wall effects, the requirement of effective catalyst and high-

temperature-resistant materials, and high energy costs150,192,199. Conventional gasification (in an

environment of insufficient oxidizer) of microalgae in a temperature range of 800 – 1000 °C was

also studied and due to the requirement of dry biomass, this pathway for biofuel production usually

resulted in negative net energy100,148,203.

Another interesting application of subcritical water is to lyse cells for extraction of high-value

intracellular products197,204,205. A particular challenge is breaking robust cell walls to allow

extraction, particularity with wet biomass197. Lab scale tests have shown that hydrothermal (Figure

3-1f) and ionic liquids (ILs) (Figure 3-1d&e) can achieve comparable efficiencies as mechanical

processing to dry biomass. Hydrothermal disruption of the cell wall at 200 °C effectively enabled

solvent extraction of astaxanthin to achieve more than 95% efficiency as opposed to control with

a 7.5% in extraction efficiency.197 ILs have been used as extracting agents to replace organic

solvents or pre-treatment material to disrupt the cell wall to enhance extraction. An extraction

efficiency of 82% was achieved using 1-Ethyl-3-methylimidazolium ethylsulfate, but a

germination process of H. pluvialis cysts is required to weaken the cell wall.194 ILs are also applied

Page 55: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

40

in a pre-treatment process to achieve > 70% extraction efficiency, but the experiment utilized dry

cell powders as feedstock which normally has weaker cell walls and the reusability of ILs needs

improvement for large scale production.196

There are also other approaches to directly extract intracellular lipids from wet biomass94,206–208,

but due to either high cost or low extraction efficiency none of them have been commercialized

yet. For biofuel productions, since the current wholesale price for diesel is down to about 70 cents

per liter, bringing additional steps or expensive additives into the process is most likely a non-

started.

3.5 Processing for microalgae in culture

To minimize the energy used in removing water from biomass, biologic conversion processes

involving fermentation, anaerobic digestion, and direct secretion have been studied. In these

pathways, biofuels can be either produced directly from microalgae culture or after initial

harvesting with minimal energy cost. Lastly this section includes microalgae-microbial fuel cell

approaches whereby microalgae are employed for direct electricity production.

3.5.1 Fermentation

Fermentation is a well-established method used in alcohol production to convert carbohydrates

into ethanol by yeast. There are two major approaches to produce ethanol from microalgae: 1) like

yeast, some microalgae such as Chlorella and Chlamydomonas can produce alcohols through

heterotrophic fermentation162,163,209,210; 2) microalgae containing a significant amount of

carbohydrates can be used as a sugar source for yeast fermentation164,165,167. In the first approach,

sugars can either be generated internally from the synthesis of microalgae or fed to algae

externally, however the process lacks the productivity of yeast fermentation. The second approach

usually requires pretreatment to promote hydrolysis of the cell wall to: 1) access to intracellular

Page 56: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

41

components; and 2) release fermentable sugars. Although research203 indicated the sugar released

from wet microalgal biomass was lower than that from dried biomass, drying the biomass prior to

conversion does not provide an energy return on investment5,8,148. Recent research161 (Figure 3-

1g) demonstrated that acid-catalyzed pretreatment prior to lipids extraction released more than

90% of fermentable sugars for wet microalgae biomass. Although noticeable achievements have

been made in this pathway, significant breakthroughs are required to produce biofuels

economically at industry scale with this approach.

3.5.2 Anaerobic digestion

Anaerobic digestion used in wastewater treatment to produce methane has the potential to exploit

the entire organic carbon content of microalgae without the requirement for drying (Figure 3-1

h&i). It can also significantly benefit from the direct use of existing infrastructure and experience

in the field of wastewater treatment168–170,211,212. Methane as the main product in the process not

only can be used as fuel but can also be converted to bioplastics to increase the value171,172,213.

However, this pathway is hindered by the low practical methane yields mainly due to the resistance

of microalgal cell walls. Rigid cell walls of microalgae not only reduce the amount of digestible

substrate but also limit the access of microorganisms to intracellular components resulting in low

reaction rate211,212,214,215. Thermal, chemical and mechanical pretreatments of microalgae were

used to improve methane yields but the energy cost involved in the extra steps is higher than the

energy gain from increased methane production168,212,216. By tuning the pH and retention time,

providing heat treatment and addition of methanogen inhibitors, this process can be altered to

produce hydrogen210,217 instead of methane but it requires significant development to be

commercially available. The more likely role for anaerobic digestion is in combination with other

methods to fully harvest the remaining value of biomass and therefore maximize the biofuel

production.

Page 57: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

42

3.5.3 Direct secretion

Direct secretion of products such as hydrogen218,219, alcohols174,220,221 and alkanes173,175,222 from

microalgae culture can be achieved through genetic engineering. Hydrogen production from

Scenedesmus obliquus was first observed by Gaffron and Rubin223 and having an anaerobic

environment was later found to be critical for this process.224 Sulfur deprivation was used to inhibit

the activity of PSII and enhance the activity of hydrogenase enzymes for hydrogen production.

Breakthroughs in genetic engineering processing parameter control on sulfur quantity and

immobilization of cells are expected to increase hydrogen productivity by further inhibiting PSII

activity and increasing the activity of hydrogenase enzymes. Secreted hydrogen can be easily

collected, however the low conversion efficiency remains a major roadblock for this pathway.

Direct secretion of alcohols and alkanes appears to be a promising alternative but most information

in this area is proprietary. Secretion of triterpene from Botryococcus braunii was reported to reach

a volumetric productivity of 22.5 mg/L/photo-h at a high cell concentration of 20 gDW/L.102 These

pathways usually require cheap source of CO2 or sugars as feedstock and development to boost

the feedstock-to-product conversion rates.

3.5.4 Microalgae-microbial fuel cells

Microalgae-microbial fuel cells biologically convert solar energy to electrical energy and while

the technology is maturing, the development of this conversion strategy is early

stage177,178,180,181,198. Microbes at the anode generate electrons, protons and CO2 by digestion of

organics and the microalgae at the cathode takes CO2, light, proton and electrons to grow through

photosynthesis (Figure 3-1j). Also known as biological photovoltaics, or BPVs, this approach has

seen many recent advances. Microfluidic approaches were applied to perform a qualitative

investigation of several key factors including cell density, electron mediator concentration and

light intensity and indicated the major obstacle for power production from BPVs is the transport

Page 58: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

43

of reducing equivalents across the cytoplasmic membrane.176,182 Understanding of electron

transport mechanism is also believed to be essential for selection of photosynthetic microbes to

enhance electrical output.179,184 Furthermore, the overall performance of BPVs is strongly

associated with surface morphology and corresponding material characteristics225, manipulation

of thylakoid terminal oxidases226,227 and types of feedstocks228. One of the challenges with this

approach is the need for electrodes which tend to block light paths in a way similar to the CO2

delivery mechanisms discussed earlier. A proof-of-concept cell was developed whereby both light

and electrons were delivered via a metallic/plasmonic surface183 , however, that approach is not

well suited to large scale production for a number of reasons. In general, there are many challenges

with microalgal-microbial fuel cell approaches including technical obstacles, high operation costs

and low power output that need to be solved for this process to be economically feasible. Similar

to some other approaches above, this strategy can be more readily adopted in a wastewater

treatment context, where the primary objective is remediation and power production is a side

benefit. We note however, that this research area is adapting quickly and may well produce

surprises 229.

Page 59: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

44

Chapter 4.

Hydrothermal Liquefaction of Microalgae on

a Chip for Biocrude Production

Hydrothermal liquefaction uses high temperatures and pressures to break organic compounds into

smaller fractions, and is considered the most promising method to convert wet microalgae

feedstock to biofuel. Although, hydrothermal liquefaction of microalgae has received much

attention, the specific roles of temperature, pressure, heating rate and reaction time remain unclear.

A microfluidic screening platform to precisely control and observe reaction conditions at high

temperature and pressure. In-situ observation using fluorescence enables direct, real-time

monitoring of this process. A strong shift in the fluorescence signature from the algal slurry at 675

nm (chlorophyll peak) to a post-HTL stream at 510 nm is observed for reaction temperatures at

260°C, 280°C, 300°C and 320°C (P = 12 MPa), and occurs over a timescale on the order of 10

min. Biocrude formation and separation from the aqueous phase into immiscible droplets is

directly observed and occurs over the same timescale. The higher heating values for the sample

are observed to increase over shorter timescales on the order of minutes. After only 1 minute at

300°C, the higher heating value increases from an initial value of 21.97 MJ/kg to 33.63 MJ/kg.

The microfluidic platform provides unprecedented control and insight into this otherwise opaque

process, with resolution that will guide the design of large scale reactors and processes.

This chapter was published as technical innovation in Lab on a Chip and reproduced from156.

Copyright © 2016, Rights Managed by Royal Society of Chemistry. This work was featured in

Lab on a Chip as HOT article. The candidate was the first author in this work and played the

primary role in designing the research, performing the experiments, analyzing the data, and writing

the paper. Additional authors for the work include Dr. Matthew D. Ooms, and Prof. David Sinton.

Their contributions were central to the publication of this work and are gratefully acknowledged

and appreciated.

Page 60: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

45

4.1 Introduction

Microalgal biomass is an attractive feedstock for the generation of carbon neutral biofuels

because of the high growth rate and lipid content of many microalgae species.32 Conversion

of the raw algal biomass into biocrude however, remains an energy intensive6 and costly

process230 which is in part why algal biofuels have not yet achieved commercial success.

To produce biodiesel from microalgae using conventional methods, the fatty components

of the cells (lipids), must first be extracted and then converted to biodiesel via

transesterification. Traditionally this is accomplished through mechanical pressing or

solvent extraction. A challenge for both these approaches is that the biomass must first be

dried, which significantly increases the carbon intensity of the final fuel as well as its cost.8

From a life-cycle assessment perspective, drying of the biomass can consume more than

90% of the energy content of the final algal oil.7 Consequently, the potential benefit of

microalgal biofuels is severely undermined by the high energy and financial cost associated

with biomass-to-biofuel conversion.

An emerging conversion approach that does not require pre-drying is hydrothermal

liquefaction (HTL).11,46,48 HTL uses high temperatures and pressures to break organic

compounds into smaller fractions to produce biocrude which can be further upgraded into

a variety of fuels. Optimizing the many parameters involved in HTL processing, however,

remains a challenge. To date, HTL experiments have been conducted in laboratory scale

batch reactors using sand-baths, ovens or heating coils for temperature control. These

approaches have two critical issues. First, due to the large fluid volumes and large physical

size of the apparatus, substantial heating times are required, which are often ignored in the

subsequent analyses. These long heating times lead to ambiguity in the reported optimal

values for temperature, heating rate, and reaction time since the heating and pressurization

delays blur the results. Second, in-situ monitoring of the reaction process is not possible in

current reactors, precluding real-time quantification of the reaction process.

Microfluidic and lab-on-a-chip methods have recently been applied to bioenergy

generation16–18, particularly with respect to microalgae. The cellular scale of microalgae

makes them well suited to manipulation and analysis using microfluidic platforms.183,231

Ensuring cells receive both light and fluids is an optofluidic challenge that has been

addressed, for instance, with integrated waveguides to deliver light to cultures232–234 and

micro-reactor arrays integrated onto individual LCD displays for parallelized illumination

Page 61: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

46

and growth studies.235 These efforts have focused primarily on cultivation of microalgae,

but have not addressed the downstream challenge of converting biomass into useful

products.

In this work, we demonstrate a high temperature and high pressure, continuous flow,

microfluidic reactor to perform controlled HTL on a glass and silicon chip (Fig. 1a). The

small length-scales of the microfluidic chip enable effectively immediate heating of the

algal slurry eliminating the ambiguity associated with conventional reactors. This work

leverages established advantages of using high-pressure high-temperature silicon-glass

microfluidic reactors10,21,86,88,236,237, for bioenergy applications. Our microfluidic chip

makes possible real-time in-situ observation including fluorescence imaging and analysis

as shown in Fig. 1. The fluorescence signatures from both chlorophyll in algae238 and

aromatics in the produced oil239 provide direct indicators of chemical composition during

the reaction.

Figure 4-1. a) Schematic of hydrothermal liquefaction of microalgae in the microfluidic chip

with in-situ observation of biocrude production using fluorescence microscopy. b) Distinct

fluorescence signatures of algae slurry at the inlet and biocrude at the outlet. Reproduced by

permission of The Royal Society of Chemistry.

Page 62: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

47

4.2 Experimental

The chip was fabricated out of glass and silicon because common chip materials such as

polydimethylsiloxane (PDMS) and polymethyacrylate (PMMA) cannot support the high

temperatures and pressures used in HTL and metal-based chips do not permit in-situ observation.

The combination of silicon and glass allowed for (i) optical access through the glass, (ii) high

thermal conductivity through the silicon, and (iii) both high pressure and temperature (over 10

MPa and 300°C respectively).

The square cross-section channel dimensions were 200 μm x 200 μm with a total length of 1320

mm, 1250-mm of which were located in a serpentine heating region. Raw algae slurry was

continuously pumped into the chip at a flow rate of 5 μL/min at 12 MPa. At this rate, the slurry

required 10 min to flow through the length of the heated region. Under steady state flow, the

reaction time (time spent in the heated region) at a given location along the channel could be

calculated. The heating rate of 15 °C/s, the fastest reported heating rate for HTL experiments in

literature, was calculated based on the 35 mm length of channel between the inlet at 50 °C and the

beginning of the heating region at 300 °C. The other pump was running at constant pressure to

provide a reference pressure for the back pressure regulator (Equilibar Inc.). The ultra-low flow,

back pressure regulator was used at the outlet to maintain steady flow at a constant pressure. It was

essential to minimize dead volume downstream of the chip in order to collect representative

samples for off-chip product analysis. Here, a separate fixed-volume sample collector loop

(300μL) was incorporated into the outlet stream, immediately downstream of the chip and

upstream of the backpressure regulator. The outlet line was switched using a 6-port valve with

minimal dead volume (Rheodyne®7030). When isolated, the contents of the collector loop could

be dispensed into a collection vial without depressurizing the entire chip. During the experiment,

the switching valve was in collection position to ensure fluids exiting the microfluidic reactor

quickly entered the fixed-volume sample isolator. Once the sample isolator is filled with

representative samples collected at steady state, the switching valve is quickly turned to the eluting

position. Then a large amount of DI water is pumped in from the syringe to eject the sample into

the sample vial. This enabled inline sampling directly from the output stream. This procedure

ensures sample quality while maintaining high pressure in the chip during collection.

Page 63: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

48

The heated region of the chip was mounted in a temperature controlled stainless steel heating

chuck, as shown in Figure 4-2. Precise temperature control was accomplished through a

proportional-integral-derivative temperature controller (Omega CNI3222) with three cartridge

heaters (Omega CSH-102135/120) maintaining a constant temperature over the entire heating

region of the chip. A thermocouple inserted in the heating chuck provided closed-loop feedback

and kept the temperature variation to less than +/- 1°C during steady state operation at a reaction

temperature of 300°C. The heating chuck was also equipped with a borosilicate glass viewing

window which enabled in-situ observation of the HTL reactions that occurred in the channel. The

chip manifold was maintained at a lower temperature (~ 50°C) to (i) prevent O-ring material failure

and leakage at the ports21 and (ii) to allow rapid and controlled on-chip heating as the fluid

transitioned from the cold to the hot zone (Figure 4-4).

Figure 4-2. Schematic representation of the assembly of the water-cooled manifold,

temperature controlled heating chuck and the microfluidic chip using a separation glass and

double seal O-ring to prevent cracking from hard contact. Reproduced by permission of The

Royal Society of Chemistry.

The biomass injected into the chip consisted of Nannochloropsis oculata with an ash content of

5.9% (Reed Mariculture Inc.). The as-received biomass was cleaned through centrifugation and

suspended in DI water prior to use. Algal slurry at a concentration of 2 wt% was then injected into

the chip using a high pressure pump with constant flow rate. This concentration was chosen in

Page 64: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

49

order to minimize clogging and fouling issues while still providing enough material to allow real-

time observation of the chemical reactions on-chip and for subsequent off-chip analysis.

During operation, the chip was monitored with a fluorescence microscope for changes in the

fluorescence signature of the algal slurry. The sample was excited with ultraviolet light (λ = 375-

400 nm) and fluorescence detected through a long-pass filter (λ > 405 nm) using a

spectrofluorometer (ocean optics USB2000) attached to the microscope.

Produced samples were collected off-chip for analysis of their higher heating values (HHV) by

isolating the reaction products in the sample collector and eluting them to small glass vials. The

product at the outlet had a variety of components including water-soluble compounds and

biocrude, with other smaller amounts of solid particulates and gas. To isolate the biocrude, 2 mL

of dichloromethane was added to the recovery vial followed by vigorous shaking for several

minutes, completely dissolving the biocrude into the dichloromethane. The vials were then set

aside to allow complete phase separation. Once separated the dichloromethane layer was

withdrawn using a glass syringe and stored in a separate vial. This extraction process was

performed multiple times to ensure more than 95% of biocrude was recovered from the sample.

The extracted biocrude was then heated in an oven at 40°C for 8 hours to remove the solvent. The

higher heating values of each sample were calculated using the modified Dulong’s formula as

follow:

𝐻𝐻𝑉 (𝑀𝐽/𝑘𝑔) = 0.335𝐶 + 1.423𝐻 − 0.154𝑂 − 0.145𝑁 (4 − 1)

A carbon-hydrogen-nitrogen elemental analyser was used to measure the carbon (C), hydrogen

(H), nitrogen (N) composition of the initial dry algae sample and the produced biocrude. The

oxygen (O) content was estimated according to:

%O = 100% − %C − %H − %N − %𝐴𝑠ℎ (4 − 2)

where the sulphur content is assumed to be negligible, as is typical for microalgae.57,240 Prior to

elemental analysis, the raw algae was dried in an oven at 105 °C for an hour.

Page 65: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

50

4.3 Results

Algae slurry was continuously pumped through the chip for 15 min to achieve steady state. By

observing the fluorescence signature at different points along the channel (Figure 4-3a), the

progression of the HTL process as a function of time spent in the reactions chamber was quantified.

Here, a series of viewing points along the channel were chosen to correspond to reaction times at

0.4-min intervals, ranging from 0 to 10 min. Near the inlet, raw algal slurry mainly showed

chlorophyll fluorescence with a peak at 675 nm (Figure 4-1b). By 1.2 min, the chlorophyll peak

had significantly dropped and an emerging peak at 510 nm began to rise. Over the next 2 min, the

510 nm peak became dominant and continued to grow. After 10 min, the original chlorophyll peak

was no longer visible and the normalized peak intensity at 510 nm approached a saturation point

as shown in Fig. 3b. The evolution of the peak at 510 nm indicated the formation of aromatic

compounds which are a characteristic component of crude oils as well as other processed plant

based oils. The progression of the fluorescence signature from one dominated by chlorophyll

fluorescence to one resembling conventional crude oils tracked the progression of the HTL

conversion process.

Page 66: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

51

Figure 4-3. a) Normalized fluorescence intensity of algae slurry observed at viewing points

along the channel under 320°C indicating the formation of biocrude over time. b) The

progression of normalized fluorescence intensity of the 510nm peak at the reaction

temperature of 260°C, 280°C, 300°C and 320°C indicating higher reaction temperature has

higher reaction rate. Solid lines included as a guide for the eye. Reproduced by permission

of The Royal Society of Chemistry.

The effect of reaction temperature was investigated by performing identical experiments at 260°C,

280°C, 300°C and 320°C. For each temperature, the normalized peak intensity at 510 nm over the

course of the reaction is shown in Figure 4-3b. The fluorescence signals were normalized to the

Page 67: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

52

projected saturation intensity based on a 1st order exponential curve fit to the experimental data.

The results in Figure 4-3b clearly show that higher reaction temperatures resulted in higher

reaction rates. Specifically, the characteristic times (time required for the fluorescence intensity at

510 nm to reach 63% of its maximum value) were: 6.0 min, 4.6 min, 3.4 min, and 1.9 min for the

reactions run at 260 °C, 280 °C, 300 °C, and 320 °C respectively.

The measured HHVs of dry algae and biocrude are shown in Table 4-1 and correspond well with

previously reported values12,65,241 using similar microalgae species and reaction conditions. As

shown, the most significant increase in HHV occurred within the first few minutes of the reactions.

Beyond reaction times of 1 minute, the variation of HHV was less than 5%.

Table 4-1: Elemental composition and Higher Heating Value of dry algae and biocrude

from 1 min, 5 min and 10 min reaction times.

Element Dry Algae Biocrude (1 min RT)

Biocrude (5 min RT)

Biocrude (10 min RT)

C [%] 51.08 70.11 70.35 70.63

H [%] 7.23 8.75 9.23 9.22

N [%] 8.65 4.46 6.00 6.56

O [%] 27.14 10.78 8.52 7.69

HHV [MJ/kg] 21.97 33.63 34.52 34.65

Combined with the fluorescence data collected, this analysis of HHV suggests that while the

energy content of the biocrude approached saturation at very early times (~ 1 min), other reactions

continued to occur that were not accounted for by the HHV alone. The sharp increase in HHV

during the first minute of the reaction was likely due to rupturing of the cells which made lipid

extraction by dichloromethane more efficient.56 As such, the increase in HHV should be used as

an indicator of mechanical disruption rather than chemical conversion which can be observed more

directly with the fluorescence signal.

Direct observation of the channel during operation provided additional insight into this process.

Most notably, non-fluorescent droplets were observed forming and adhering along the length of

the channel and increased in size towards the outlet (Figure 4-4a). These droplets are expected to

Page 68: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

53

be comprised of aliphatic compounds which are immiscible with water and do not fluoresce. The

formation of these droplets progressed along the length of the channel which suggests a correlation

between the formation of immiscible oil droplets and the parallel change in fluorescence signature

resulting from the formation of aromatic compounds in the aqueous phase. Additionally, the

degradation of the microalgae could be observed between the inlet where whole cells were clearly

visible and the outlet where individual cells were no longer discernable (Figure 4-4b).

Figure 4-4. a) Fluorescence images obtained at viewing points along the channel with

increase in reaction time indicating the progression of biocrude formation. b) Microscopic

observation of fluids at the inlet and outlet via both fluorescence and dark-field imaging.

Scale bars: 50 μm. Reproduced by permission of The Royal Society of Chemistry.

Page 69: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

54

It was also observed that a significant amount of cell debris and solid particulates (and in some

cases, clogging) resulted from operating with short reaction times. Specifically, we observed that

at 320oC when the flow rate was increased such that the maximum reaction time was only two

minutes, a significant amount of solid debris remained in the effluent. In contrast, less solid debris

in the output was observed for longer reaction times achieved with lower flow rates and otherwise

similar conditions. This finding is likely a result of more complete disruption of the biomass during

the first few minutes of the reaction (also indicated by the increase in HHV described earlier).

Furthermore, clogging at the outlet was exacerbated by the rapid cooling of the effluent which

promoted the separation of the oil phase from the aqueous phase, which was directly observed as

an increase in the number and size of oil droplets in the cooled outlet line (Figure 4-4a). Lastly,

the multiphase nature of the generated products, visible as channel-adhered droplets in Fig. 4, will

influence to some extent the residence time, accumulation and ultimate production of different

components. The HHV analysis here, however, is largely unaffected by this issue as the oil is

separated from the produced fluid and based on the relative elemental composition of the biocrude.

These observations, which were made possible only through the direct visualization afforded by

our chip design, have implications for the optimum processing parameters of continuous flow HTL

reactors. Specifically, these findings indicate optimal reaction times between 2 and 10 minutes to

both maximize the conversion of biomass to biocrude, and minimize the amount of debris in the

effluent to prevent fouling. Gradual cooling of outlet stream is also recommended to avoid

clogging.

4.4 Conclusion

In summary, our microfluidic reactor provides unprecedented insight and control over the high

temperature and high pressure cracking of biomass via hydrothermal liquefaction. It allows for in-

situ observation of hydrothermal liquefaction reactions using fluorescence microscopy and

convenient and precise control of reaction temperature, pressure and reaction time in a continuous

flow reactor. These advantages enable the study of high temperature and pressure cracking of

biomass on a platform with a high degree of control which will allow improved understanding of

the reactions taking place during hydrothermal liquefaction. The significant change of

fluorescence signature between the algal slurry (peak at 675 nm) and converted biomass (peak at

510 nm) was observed as an indicator of the progression of hydrothermal liquefaction. Biocrude

formation and separation from the aqueous phase into immiscible droplets was directly observed

Page 70: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

55

and occurred over timescales of ~10 min. The rapid increase of higher heating values was observed

over the timescales of ~1 min and was correlated to observations of particulate matter in the

effluent which manifested as partially clogged channels. These results and the microfluidic

platform on which they were collected represent the first of their kind in the field of hydrothermal

liquefaction research. Lab-on-a-chip methods offer a unique toolset to probe high temperature and

high pressure reaction dynamics and inform large scale reactor design.

4.5 Supplementary Information

4.5.1 Fabrication of microfluidic chip

To achieve both high temperature and pressure on a chip, the high pressure connection between

the chip and the tubing needed to remain at low temperature (Figure 4-5a). A window cut in silicon

was used to reduce heat transfer from high temperature area to high pressure compression ports

and therefore a large temperature gradient was generated along the arms. The chip fabrication

followed the process indicated in Figure 4-5 b. The features on the chip was generated on the

silicon and bonded to glass at the end. The whole fabrication process included two sets of

spincoating, patterning, and Deep reactive-ion etching (DRIE) and therefore two types of

photoresists (S1818 and AZ4620) and two photomasks were used. The silicon wafer was 4” in

diameter,1mm thick, and double side polished. The channel was patterned by photomask #1 and

generated by DRIE for a depth of 200um. The holes and the window cut in silicon were generated

by etching through using DRIE. The pattern on the other side of the channel (called back side in

the figure) was aligned using Back Side Alignment technique with designed markers on both

photomasks. Once the features were completed on the silicon wafer, both silicon and glass were

thoroughly cleaned using Piranha solution before anodic bonding.

Page 71: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

56

Figure 4-5. a) Schematic illustration of achieving high temperature and pressure on a chip

by separating high pressure compression from high temperature area. b) Si/glass chip

fabrication process (from top to bottom, left to right).

4.5.2 Experimental apparatus

The experimental apparatus contains a number of components and the flow path among them is

indicated in the Figure 4-6. All critical components in the apparatus have pressure rating higher

than 20 MPa and are mainly connected using Yor-Lok fittings. Two high pressure pumps (ISCO

260D) were used in this experiment. One was running at constant flow rate to continuously pump

algae slurry in the piston cylinder to the chip for a desired reaction time. The manufacturing

numbers and the application purposes of major components are provided in Table 4-2.

Page 72: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

57

Figure 4-6. Schematic diagram of experimental setup with flow direction indicated by arrows

along the processing path. The flow path of the switching valve at two positions is shown in

green lines.

Table 4-2: List of components in the apparatus and their purpose.

Number Part Name Mfg Number Purpose

1 High Pressure

Pump

ISCO 260D 1a. Pumping algae slurry at constant flow rate

1b. Maintaining constant pressure for BPR

2 Piston cylinder HIP TOC3-10 Algae slurry container

3 Switching valve Rheodyne®7030 Switching flow paths without depressurizing the

whole system

4 Fixed-volume

sample isolator

Radel® R Tubing

1220

Temporarily storing fixed-volume sample in the

loop ensures sample quality

5 Back pressure

regulator

Equilibar

EB1ULF1 - SS316

Maintaining constant pressure for a continuous

flow at ultra-low flow rate

4.5.3 High temperature and pressure packaging

The manifold module is used to provide high pressure sealing between the microfluidic chip and

the rest of the apparatus. This manifold uses a modular design (Figure 4-7) and can interface with

any chip sharing the same port pattern and thickness. Compression between the O-ring (Double-

Seal Viton® 004) and the chip is achieved through screw fasteners and care must be taken when

Page 73: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

58

tightening to avoid fracturing the chip. To assist with tightening, a spacer is placed between the

clamp and the manifold to prevent fracture by overstressing. Also, a layer of polished glass is

placed between the chip and the clamp to provide even clamping pressure. The manifold was

fabricated out of stainless steel (SS316).

Figure 4-7. Detailed drawing of the compression sealing of the microfluidic chip (MF) with

algae slurry fluid indicated in green. O-rings and spacers are used to ensure a quality seal

between the manifold and the chip and to prevent overtightening and damage.

Page 74: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

59

Chapter 5.

Hydrothermal Disruption of Algae Cells for

Astaxanthin Extraction

We demonstrate a hydrothermal method of astaxanthin extraction from wet biomass using a high

temperature and high pressure microfluidic platform. Haematococcus pluvialis cysts are trapped

within the device and visualized in-situ during the cell wall disruption and astaxanthin extraction

processes. The device provides a highly controlled environment and enables direct comparison of

chemical vs. hydrothermal processes at the cellular level. Hydrothermal disruption at a temperature

of 200 °C was shown to be highly effective, resulting in near-complete astaxanthin extraction from

wet biomass - a significant improvement over traditional methods.

This chapter was published as communication in Green Chemistry and reproduced from [ref].

Copyright © 2017, Rights Managed by Royal Society of Chemistry. The candidate was the first

author in this work and played the primary role in designing the research, performing the

experiments, analyzing the data, and writing the paper. Additional authors for the work include

Dr. Jason Riordon, Mr. Brian Nguyen, Dr. Matthew D. Ooms and Prof. David Sinton. Their

contributions were central to the publication of this work and are gratefully acknowledged and

appreciated.

5.1 Introduction

Astaxanthin is a highly valuable microalgal bioproduct with uses ranging from human health to

aquaculture. While synthetic production is currently favoured in industry due to lower production

costs, natural astaxanthin is considered more beneficial than synthetic astaxanthin due to its

superior antioxidant activity.69,70,242 The microalgae Haematococcus pluvialis is by far the richest

natural source of astaxanthin243 and has become the primary source for the nutraceutical industry.72

However, extraction of astaxanthin remains a challenge; the stress conditions that induce

astaxanthin accumulation in H. pluvialis cells also induce cell wall thickening. This barrier is

extremely robust to chemical and physical disruption,76 making astaxanthin extraction difficult,

particularly for mature cysts which are larger and have thicker cell walls compared to cells at other

Page 75: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

60

life-cycle stages. Effective cell wall disruption increases the effectiveness of post-disruption

recovery approaches, which include organic solvents75, ionic liquids,194,196 and supercritical

CO2153,244. Chemical disruption using acids or bases typically results in low (<40 %) extraction

efficiencies.75 Mechanical disruption is therefore the primary method used in industry, but requires

dry biomass.72 Biomass drying in combination with mechanical disruption enhances the extraction

efficiency (>80 %) but also significantly increases the energy and financial costs of

processing245,246.

Hydrothermal processes – where wet biomass is subject to high pressures (5 – 20 MPa) and

temperatures (150 – 350 ºC) – have shown promise for several applications, including pressurized

hot water extraction of bioactives247,248, woody biomass decomposition249 and biocrude

formation57,187,250,251. Hydrothermal processes leverage the physiochemical characteristics of water

at elevated temperature and pressure, including: (i) a significantly higher fraction of water ions

which aids acid or base-catalysed reactions to break cell walls; (ii) a lower dielectric constant,

which enhances the hydrothermal conversion of carbohydrate biomass252; and (iii) energy savings

by avoiding the need for water evaporation.10 Importantly, hydrothermal disruption represents an

environmentally friendly, chemical-free approach, in sharp contrast to traditional chemical

methods.253 However, despite the strong potential of hydrothermal processes for cell wall-

disruption, such an approach has not been applied to astaxanthin extraction.

Current approaches to studying cell disruption and extraction processes of this nature generally

involve relatively large opaque batch reactors.75,254 Significant chemical and thermal gradients are

inherent in such reactors given their size, and pressure vessel requirements generally preclude

direct observation of the process86,88,255. Operating at smaller length scales, as with silicon-glass

microfluidic reactors, can minimize gradients, improve control, and provide access to high

temperatures and pressures, all while allowing direct optical access to the process.21,156 While such

microfluidic systems are themselves impractical for the production of bulk product, they are

ideally suited to screening conditions and quantifying unit process efficiencies – valuable

information for commercial-scale processing.

Page 76: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

61

Figure 5-1. Simplified schematic of on-chip astaxanthin extraction from H. pluvialis.

Enlarged schematics of the cell capture area show initial cell trapping, cell wall disruption

and astaxanthin extraction. Reproduced by permission of The Royal Society of Chemistry.

In this work, we demonstrate a hydrothermal approach to cell wall disruption for astaxanthin

extraction from wet biomass, where extraction is visualized on a microfluidic screening platform.

Extraction was performed using a microfluidic device to provide real-time visualization of the

disruption and extraction processes. The glass/silicon microfluidic chip was designed to allow

optical access, while providing uniform heating and both chemical and thermal resiliency. The

device features parallel reaction channels with trapping posts, where H. pluvialis cysts are trapped

and monitored during wall disruption and astaxanthin extraction (Figure 5-1). Direct visualization

of the process enables quantitative comparison of chemical and hydrothermal methods at the cell

level. It was found that biomass treated hydrothermally, at a temperature of 200 ºC and a pressure

of 6 MPa, demonstrated near-complete astaxanthin extraction efficiency – a significant

improvement over traditional approaches.

Page 77: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

62

5.2 Experimental setup

5.2.1 Device design and fabrication

The silicon-glass microfluidic device featured 40 parallel reaction channels, each 200 µm wide

and 100 µm tall. Each of these channels was lined with cylindrical posts of 20 µm diameter

separated by a distance of 40 µm. These posts served to trap cells while allowing fluid flow. The

device was fabricated by deep reactive ion etching (DRIE) of the silicon, with subsequent bonding

to borosilicate glass of 1.75 mm thickness. The full fabrication procedure for devices of this type

is reported elsewhere.156

5.2.2 Cell culture and trapping

Haematococcus pluvialis cultures were obtained from Algae Analytics (Las Cruces, New Mexico)

and cultured at 25 ºC under continuous white light illumination at ~30 µmol m−2s−1 in media

according to the recipe of Fábregas et al.256. Encystment was induced by exposure to red and blue

light at ~60 µmol m−2s−1 until the cultures developed red coloration indicative of astaxanthin

accumulation (irradiation spectrum shown in Figure 5-7). For each experiment, a suspension of H.

pluvialis cysts in media was injected into the chip using a pump and trapped in the post arrays (a

video of initial cell trapping is presented in ESI† - Video S1). Cysts with thickened cell walls were

selected by using a 70 µm cell strainer. After the cells were loaded, deionised water was flowed

into the chip at a rate of 1 mL min-1 for 5 min to flush the media prior to cell disruption. During

cell loading, and at all stages of astaxanthin recovery, the cells were observed using darkfield

microscopy (Olympus BXFM microscope, 10x objective).

5.2.3 Cell wall disruption and extraction

The cell wall disruption method was varied to test both chemical treatments (HCl 0.1M, NaOH

0.1M) and hydrothermal treatments (30 min @ 150 ºC, 30 min @ 200 ºC, 10 min @ 200 ºC and 5

min @ 200 ºC), and compared to a control case (no treatment). For chemical disruption, solutions

were pumped into the device at a flow rate of 1 mL min-1 for 1 min to rapidly fill the volume of

the chip before stopping the flow for 30 mins. The hydrothermal processes were performed at a

pressure of 6 MPa, chosen to be sufficiently high to maintain water at liquid phase at the selected

temperatures. The device was then heated to the desired temperature at a heating rate of ~60 ºC

min-1 using a custom heating chuck and maintained at the set temperature for 5-30 min by a PID

Page 78: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

63

temperature controller. Once the hydrothermal treatment was completed, the heating chuck was

removed to allow the device to cool to room temperature.

The temperature of 200 °C was chosen to correspond to the approximate temperature where

structural cell wall polysaccharides rapidly depolymerize without degradation of other biomass as

a result of secondary reactions.13,257 After cell wall disruption, acetone was flowed at a rate of 1

mL min-1 for 1 min, before being reduced to a constant rate of 100 µL min-1 for another 20 min to

fully purge the microchannels. The solvent extraction process was identical in all cases.

5.2.4 Quantifying astaxanthin content

To compare the efficiency of astaxanthin extraction methods, the change in coloration of

individual H. pluvialis cells was quantified by image processing for all cases. An equation was

devised which converts the 8-bit RGB values measured for each pixel within the area of a cell,

into a coloration-based global “extracted red content” metric, indicative of the relative astaxanthin

concentration:

𝐴 = (100 −0.59𝐵 + 0.41𝐺

2.55) ∗ (1.36

255−𝑅45 ) (5 − 1)

where R, G and B represents the red, blue and green value at that pixel, respectively. The first

factor was determined using the 475 nm absorbance peak of pure astaxanthin,[258] which

corresponds to R, G and B values of 0, 178, and 255 respectively. Blue and green values are

weighed based on their relative importance to astaxanthin absorption. The second factor was

determined based on the observed drop of the R value in control experiments, without cell-wall

disruption, due to cell shrinkage and associated increase in astaxanthin pigmentation within that

smaller area. Cell shrinkage under exposure to hypertonic media during preliminary control

experiments was used to establish a relationship between R value and astaxanthin concentration –

R value within a cell with a constant quantity of astaxanthin, albeit at different times with different

volumes. Such shrinkage was not induced during hydrothermal disruption experiments presented

herein. It was determined that in this case, a drop of 45 in R value after solvent extraction

corresponded to a 36% decrease in area and corresponding increase in local astaxanthin

concentration. Individual cells in images were selected and tracked to calculate the extraction

efficiency according to (Ai-AF)/Ai where Ai and AF represent the initial and final red content.

Page 79: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

64

Images taken during hydrothermal disruption processes were slightly affected by the heat but the

effect on RGB values was negligible. Notably, our evaluation based on relative red content

provides a rapid, in-situ measure of cell disruption and extraction efficiency. The observed red

pigment is expected to be 80-99% astaxanthin (free form, 4-5%; monoesters, ~70%; and diesters

15-20%) out of overall carotenoids based on previous studies72,259,260. The dark red astaxanthin

pigment within otherwise largely transparent cells provides a high contrast handle for rapid, real-

time screening of conditions at the individual cell level, not possible with off-chip methods. The

method is thus well suited to rapidly screen a large number of conditions for disruption and

extraction, with the best performing conditions candidates for more detailed analysis.

5.2.5 HPLC analysis of extracted astaxanthin

The acetone extracts from the hydrothermally treated (200 ºC for 10 min) cells and mechanically

disrupted cells (by mortar and pestle) were analysed using a high-performance liquid

chromatograph (HPLC) (Shimadzu SPD-10A) equipped with a reversed phase C18 (Supelco, 25

cm × 4.6 mm) column. The detection wavelength was set to 475 nm and the analysis protocol was

identical to previous research.258

5.3 Results and discussion

Initial and post-extraction cell images are presented for each wall disruption method in Figure 5-

2a, enabling direct comparison of different wall disruption methods. Each post-extraction image

was taken immediately after acetone extraction. Figure 5-2b shows normalized extracted red

content, as obtained with equation (1). Notably, hydrothermal disruption methods demonstrated

the largest change in coloration, with cells transitioning from a dark red hue to bright red (at 150

ºC) or white (200 ºC), indicative of near-complete astaxanthin extraction (>95% red content

reduction for the 200 ºC cases). Shorter duration tests at 200 °C indicate that the disruption process

is rapid at this temperature and insensitive to duration. A hydrothermal disruption temperature of

200 °C was sufficient to quickly decompose the cell wall in a span of a few minutes, which is in

agreement with previous research.13 Cells that were not subjected to any wall disruption treatment

(control case) show very little coloration change (7.5% red content reduction). The exchange of

media from deionized water to acetone causes water to exit via the water-permeable cell, resulting

in a 34% reduction in cell area. That cells retain their circular shape is indicative that the cell wall

and plasmalemma were not significantly compromised during acetone exposure. As the cell wall

Page 80: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

65

contracts, however, light scattering at the cell wall - responsible for the white halos seen in the

initial images – is greatly reduced, with only light blue outlines remaining. This change is

attributed to water loss within the cell, and the related contraction of the plasmalemma and outer

cell wall (disappearance of inner bright ring, corresponding to the plasmalemma, confirmed by

brightfield imaging, ESI). The extraction efficiencies for acid and base treatments were higher than

the control, achieving 19.3% and 13.2% recovery, respectively. Cell size reduction was similar to

the control case (36%, 34%). During the HCl disruption process, astaxanthin bleeding was

observed in one of the cells, indicting minor cell wall disruption in isolated cases. For the NaOH

disruption process, no discernible structural changes other than size reduction occurred. The

relative extraction efficiencies achieved are slightly improved and shows similar improvement

(control, 20%; acid treatment, 35%; and base treatment 40%) to those published previously in bulk

reactors with longer extraction time.75 Collectively, these result demonstrate that short duration

treatment at 200 ºC is very effective for subsequent astaxanthin extraction, in sharp contrast to

traditional acid and base disruption methods.

Page 81: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

66

Figure 5-2. a) Dark field images of mature red cysts of H. pluvialis at both the initial stage

and after solvent-based extraction, for each of five tested cases. All images have the same

scale bar, and were obtained with identical settings using darkfield microscopy. b)

Normalized extracted red content for each case. Reproduced by permission of The Royal

Society of Chemistry.

Page 82: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

67

Figure 5-3 shows detailed time-course images of cells during the full extraction protocol, for each

of the four hydrothermal cell wall disruption recipes (detailed intermediate steps in the processes

of Figure 5-2). As shown, a treatment temperature of 150 ºC was sufficient to disrupt the

plasmalemma, but it was insufficient to disrupt the cell wall. In the 150 ºC case run for 30 min, the

cells darkened during heating followed by continuous fading of the outer regions (between the

“initial” and “beginning of the disruption process” images). After a return to room temperature, a

5 µm layer of interspace between the plasmalemma and cell wall was observed. A video of acetone

extraction is presented in ESI† - Video S2.

Figure 5-3. Time-course images of red cysts treated with hydrothermal processes of 150 oC

for 30 min, 200 oC for 30 min, 200 oC for 10 min and 200 oC for 5 min, respectively. The

scale bar is identical for all images. Reproduced by permission of The Royal Society of

Chemistry.

Figure 5-3 also shows the high efficiency of hydrothermal astaxanthin recovery at 200 ºC, for all

tested disruption times – with slight improvements at longer times. For the 30 min at 200 ºC case,

initial fading around the cell periphery was observed during heating (as observed in the 30 min at

150 °C case). However, rapid growth of the interspace and significant astaxanthin bleeding

occurred 1 min into the disruption process (Figure 5-3), indicating disruption of both the

Page 83: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

68

plasmalemma and the cell wall. Rapid cell wall disruption within 1 min of treatment is also

supported by our study of the role of flow rate, presented in Figure 5-9. During the 200 °C

disruption, the polysaccharide wall is expected to slowly depolymerise into furan compounds49,

however, no discernible color change was observed on the cell wall, and furan-based RGB value

modifications are considered negligible. Astaxanthin flow from the cell to the channel was

observed 20 min into the disruption process. The majority of the pigmentation change occurred 1

min into acetone extraction, indicating high cell permeability, in contrast to that observed for other

cell disruption conditions. After 20 min of solvent extraction the cells showed virtually no

pigmentation indicating near complete astaxanthin recovery. The other tested times for the 200 ºC

case (10 min, 5 min) produced similar results, with the 5 min case showing a pigmentation change

during extraction that was less complete. These observations, which were made possible through

the direct visualization enabled by our microfluidic platform, have implications for the optimum

processing parameters of cell wall disruption techniques in larger commercial systems. These

findings indicate an optimal disruption duration between 5 and 10 min to both maximize the

extraction efficiency, and minimize the amount of time and energy required for processing.

Figure 5-4. Normalized red content during acetone extraction for six cells treated by

hydrothermal processing at 200 °C for 10 min. Inset images shows corresponding images of

one representative cell at four times during the procedure. Reproduced by permission of The

Royal Society of Chemistry.

Page 84: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

69

The first 25s of the extraction profile for six individual cells disrupted for 10 min at 200 °C is

shown in Figure 5-4. The plot demonstrates the speed at which cell respond to acetone extraction

after hydrothermal treatment. Six cells were tracked to monitor the astaxanthin content which was

normalized to the post-treatment conditions of each cell. The first 25 seconds of solvent extraction

show a rapid decrease in astaxanthin (> 90%) in each cell (a video showing rapid extraction is

presented in ESI† - Video S3).

Figure 5-5. HPLC analysis of cell extract products for (a) mechanical extraction using a

mortar and pestle and (b) hydrothermal extraction for 20 min at 200 °C treatment.

Reproduced by permission of The Royal Society of Chemistry.

The HPLC analysis in Figure 5-5 demonstrates that the extracted red content for the case of 200 °C

for 10 min is similar in profile to that of cells mechanically disrupted using a mortar and pestle –

and correspond to expected astaxanthin HPLC peaks. Such a similarity demonstrates the

effectiveness of rapid hydrothermal treatment as an attractive alternative to mechanical disruption

methods. There are, however, two notable differences between the spectra. First, diester peaks are

clearly identifiable in the mechanical extraction case, but not in the hydrothermal case. Such a

Page 85: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

70

difference could be due to a portion of diesters being hydrolyzed into monoesters and further

converted to free form astaxanthin. Second, the magnitude of peaks overall is much lower in the

hydrothermal on-chip case due to a level of dilution (a few nanograms of astaxanthin dissolved in

1mL of acetone), and is only an artefact of the visualization method here – trapping a small number

of cells in a flow – and is not a practical limitation of high temperature hydrothermal disruption at

larger scales. Notably, the presence of monoester and carotenoid HPLC peaks suggests the high

temperature, high pressure hydrothermal process did not have a significant adverse effect on

astaxanthin in general, perhaps due to the presence of largely intact rigid cell walls, which have

been shown to protect astaxanthin from thermal degradation at higher temperatures (eg. spray

drying at 220 ºC).261 The absence of oxidising agents within the cells, and inability for external

oxygenating species to enter the reactor, led to effective astaxanthin extraction.

Whereas the presence of strong astaxanthin HPLC peaks in Figure 5-5b demonstrates the

effectiveness of our hydrothermal disruption method, there remains uncertainty as to any

temperature-induced conformational changes, such as trans-cis isomerization of extracted

carotenoids. Kaczor and Baranska performed in-situ Raman spectroscopy to monitor astaxanthin

structural change in a single cell with thermal stress up to 150 °C.262 It was found that astaxanthin

experienced only minor conformational change. While more study is required to elucidate

astaxanthin structural changes under high pressure and high temperature conditions, strong HPLC

peaks here demonstrate the potential of the hydrothermal extraction method.

There is potential to adapt the hydrothermal astaxanthin extraction method presented herein for

commercial-scale reactors. To achieve industrial-scale hydrothermal astaxanthin recovery, the

results of our chip-scale experiments suggest three criteria that must be met: (i) an inert

environment is required to prevent astaxanthin oxidation, (ii) temperature, pressure and residence

time must be precisely controlled, and (iii) low shear conditions must be satisfied to prevent cell

wall rupture and subsequent astaxanthin degradation (as observed in Figure 5-10 in flow

experiments). Such a combination of criteria could be met by with a continuous flow-through

reactor approach, similar in principle to the experiments performed herein, albeit with macroscale

reactor tubes in parallel. In short, while the chip allows unprecedented cell-scale resolution of the

process, the reaction conditions achieved here are in no way unique to the chip-based reactor. The

observations here can be readily applied to engineer scaled processes, with careful engineering to

ensure similar conditions and residence times.

Page 86: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

71

5.4 Conclusion

In this chapter, we have demonstrated a hydrothermal method of cell wall disruption for

astaxanthin extraction uniquely enabled by a high-pressure, high-temperature microfluidic device.

Individual H. pluvialis cells were trapped and visualized throughout one of several procedures,

both chemical and hydrothermal, and extracted red content was quantified optically. Hydrothermal

disruption at 200 ºC was the most effective wall disruption technique, enabling near complete

astaxanthin extraction from wet biomass, a significant improvement over traditional methods.

5.5 Supplemental material

The experimental apparatus is adapted from our previous research with slight modification

upstream of the microfluidic reactor (fabrication details are shown in Chapter 4). The algae

solution, chemicals and solvent stored within piston cylinders [2a & 2b] were separately injected

into the microfluidic reactor by controlling the switching valve [3a]. Switching valve [3b] was

fixed at position 1 until the solvent extraction stage. After 1 min of solvent extraction, the switching

valve [3b] was quickly switched to position 2 to unload the extracted astaxanthin to the fixed-

volume sample isolator.

Figure 5-6. Schematic diagram of the experimental setup with flow direction indicated by

arrows along the processing path. The flow path of the switching valve at two positions is

shown using green lines.

Page 87: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

72

Figure 5-7. Irradiation spectrum of light used to induce astaxanthin accumulation in

Haematococcus pluvialis.

Figure 5-8. Microscope images of red cysts, indicating coloration change during the heating

up phase in the hydrothermal disruption process. The scale bar applies to all images.

Page 88: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

73

Figure 5-9. Dark field, bright field and fluorescence images of initial cells and after acetone

extraction. Fluorescence 1 and 2 images were taken using FITC (excitation filter: 475/50 nm;

emission filter: 540/50 nm) and TxRed (excitation filter: 559/34 nm; emission filter: 630/69

nm) filter cubes respectively. The scale bar of 50 µm applies to all images.

Page 89: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

74

Figure 5-10. Cell wall deformation of red cysts under hydrothermal processes at 200 °C in

10 min with and without flow. Green dashed lines indicate cell wall boundaries. The scale

bar of 50 µm applies to all images.

Video S1: Real-time video showing H. pluvialis cells being trapped by posts in a microfluidic

channel. Cells in this video were not fully transformed into mature red cysts and are here used as

a demonstration.

Video S2: Real-time video showing the in-situ acetone extraction of red cysts of H. pluvialis

treated by hydrothermal disruption at a temperature of 150 °C and pressure of 6 MPa for 30 min.

Video S3: Real-time video showing the in-situ acetone extraction of red cysts of H. pluvialis

treated by hydrothermal disruption at a temperature of 200 °C and pressure of 6 Mpa for 10 min.

Page 90: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

75

Chapter 6.

Astaxanthin Extraction from Algae using

Supercritical CO2 with Co-solvent

Supercritical CO2 is an attractive green-solvent for the extraction of astaxanthin from microalgae,

but current processes require prohibitively high pressures and long extraction times. This study

demonstrates the efficacy of low pressure supercritical CO2 extraction of astaxanthin from

disrupted Haematococcus pluvialis. We employ a microfluidic reactor that enables excellent

control and allows direct monitoring of the whole process at the single cell level, in real time.

Astaxanthin extraction using ScCO2 achieved 92% recovery at 55 °C and 8 MPa applied over 15

hours. With the addition of co-solvents, ethanol and olive oil, the extraction rates in both

experiments were significantly improved reaching full recovery within a few minutes. Notably,

for the ethanol case, the timescales of extraction process are reduced 1800-fold from 15 hours to

30 seconds at 55 °C and 8 MPa, representing the fastest complete astaxanthin extraction at such

low pressures.

This chapter was submitted as original paper to Bioresource Technology. The candidate was the

first author in this work and played the primary role in designing the research, performing the

experiments, analyzing the data, and writing the paper. Additional authors for the work include

Mr. Zhenbang Qi, Dr. Thomas Burdyny, Ms. Tian Kong and Prof. David Sinton. Their

contributions were central to the publication of this work and are gratefully acknowledged and

appreciated.

6.1 Introduction

The production of high-value products from microalgae is currently the most economically viable

option for the industry to offset the high overall costs of growth and extraction.15,69,242,263

Astaxanthin naturally occurs in a number of aquatic species including the microalgae

Haematococcus pluvialis which accumulates the highest levels per cell. Due to its unique structure

and strong antioxidant properties in humans70,242, natural astaxanthin from microalgae has

Page 91: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

76

cultivated a large nutraceutical industry264–266. Unfortunately, the accumulation of astaxanthin

within H. pluvialis is strongly linked to the cell’s rigid wall that is extremely resistant to both

chemical and physical disruptions.73,74,118,267 Also, once the cell wall is breached, organic solvents

are generally needed to extract astaxanthin from the interior. The range of available disruption and

extraction methods and solvents is limited by the requirements for human product consumption.268

A detailed understanding of factors affecting astaxanthin extraction using green solvents at the

cellular level are needed to help inform future large-scale astaxanthin production processes.

Recently, supercritical carbon dioxide (ScCO2) has attracted attention as a green solvent for the

extraction of a variety of bioactive compounds.78,269 Supercritical CO2 features three major

advantages over organic solvents: 1) it is abundant and benign to human health and the

environment; 2) CO2 solvent can be removed easily by evaporation at room temperature and

pressure; and 3) bioactive compounds are well-preserved due to the inertness of CO2 and relatively

low critical temperature (31.1 °C).77,128 However, previous studies using ScCO2 to extract

astaxanthin show large variations in the overall recovery rates, even with similar processing

conditions (Table 6-1). The overall extraction efficiency with ScCO2 depends greatly on the

preparation of the feedstock and the addition of any modifiers/co-solvents, while operating

conditions (temperature, pressure, duration) and extraction cycles are parameters which could also

be adjusted. With regard to commercial feasibility, temperatures for astaxanthin extraction are

limited to a relatively low 70 °C81,83, while the minimum pressure needed for ScCO2 is 7.39 MPa

(although the majority of work to date employs much higher pressures).

Due to the rigid cell wall, direct extraction of wet microalgae biomass using ScCO2 is generally

not feasible. A common strategy is then to dehydrate and mechanically disrupt the feedstock prior

to CO2 extraction, easing the downstream extraction process. However, reported astaxanthin

recovery rates vary widely due to the range of disruption tools used and the ambiguity in

quantifying the degree of cell wall disruption. For example some researchers found the overall

recovery rates were below 50% after cell wall disruption, even when the operating pressure was

as high as 30 MPa.79,80 Furthermore, dehydration and mechanical cell disruption add considerable

time, energy and cost to the process. Intracellular water and disrupted cell debris can also present

challenges such as caking, further reducing extraction rates and/or yield over time.270

Page 92: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

77

Another strategy to aid the extraction process is to employ co-solvents81,84 such as ethanol and

vegetable oils in ScCO2. Although this strategy allowed for a higher recovery rate of 71 %, high

extraction pressures (31 MPa) and multiple cycles of batch processing (8 cycles) were required

(Table 6-1 - #5). The combination of mechanical cell disruption and co-solvent extraction has also

been tested79,84, indicating the potential to greatly speed up extraction and enhance recovery using

ethanol co-solvent. However, the high associated capital and operating costs present barriers to

large-scale production271,272. Large discrepancies between reports in literature present additional

uncertainty. Process control and cell-scale resolution are essential to quantify the effectiveness of

extraction processes, and to inform ScCO2-based extraction of astaxanthin from microalgae at

industrial scales.

Table 6-1: Summary of studies on ScCO2 extraction of astaxanthin from H.pluvialis

# Feedstock Solvent Extraction

conditions Duration

Recovery

rate (%) Reference

1 Wet

Biomass

Acetone STP 16 h 20 Mendes-Pinto et

al., 2001

2 CO2 27.6 Mpa, 60 °C 30 min 1 Pan et al., 2012

3

Dry powder

CO2 40 Mpa, 70 °C 5 h 25 Krichnavaruk et

al., 2008

4 CO2 50 Mpa, 70 °C 4 h 19 Machmudah et

al., 2006

5 Ethanol in CO2

= 9.23 mL/g 31 Mpa, 50 °C

160 min (20 min x 8)

71 Pan et al., 2012

6 Ethanol in CO2

= 2.3 mL/g 43.5 Mpa, 65 °C 3.5 h 87 Wang et al., 2012

7 Ethanol in CO2

= 10 % (v/v) 40 Mpa, 70 °C 5 h 51

Krichnavaruk et

al., 2008

8 Ethanol in CO2

= 5% (v/v) 40 Mpa, 70 °C 4 h 80

Machmudah et

al., 2006

9

Disrupted dry powder

CO2 35 Mpa, 55 °C 2 h 5

Reyes et al., 2014

10 Ethanol in CO2 = 13 % (w/w)

35 Mpa, 55 °C 2 h 56

11 Ethanol in CO2 = 50 % (w/w)

7 Mpa, 45 °C 2 h 124

12 Ethanol in CO2 = 70 % (w/w)

7 Mpa, 45 °C 2 h 65

Page 93: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

78

13

Disrupted dry powder – degree 1

CO2 30 Mpa, 60 °C NA 47

Nobre et al., 2006 14

Disrupted dry powder – degree 1

Ethanol in CO2 = 10 % (v/v)

30 Mpa, 60 °C NA 59

15

Disrupted dry powder – degree 2

Ethanol in CO2 = 10 % (v/v)

30 Mpa, 60 °C NA 92

*NA: not reported

In this paper, we screened the extraction conditions of astaxanthin from H. pluvialis cells using

ScCO2 with in-situ observation on a pressure- and temperature-controlled microfluidic platform.

For the first time, direct monitoring in real time and quantification of the astaxanthin extraction

process on a single cell is realized. Applying this method we demonstrate potential for rapid, low-

pressure astaxanthin extraction with ScCO2 and co-solvent.

6.2 Experimental Section

Haematococcus pluvialis (H. pluvialis) cells were obtained from Iconthin Biotech Corp. and

maintained in aplanospore form at 25 ºC under continuous white light illumination at ~60 µmol

m−2s−1 in media according to the recipe described elsewhere.256 The mature H. pluvialis cysts with

cell diameters from 40 to 70 µm were selected by using two cell strainers. CO2 (99.9%) was

purchased from Praxair. Ethanol (96%) was purchased from Sigma-Aldrich, and the olive oil was

commercially available extra virgin, food grade.

The microfluidic chip was fabricated out of glass and silicon to provide optical access to the

reaction channel and high thermal conductivity for precise temperature control. This device

featured a long single reaction channel, 300 μm wide, 100 μm deep and 680 mm long, with

multiple C-shape traps located in the temperature-controlled region. Each trap has an inner

diameter of 100 μm and outer diameter of 160 μm with a front opening of 80 μm and back opening

of 20 μm. These traps served to immobilize cells while allowing fluid flow. The device was

fabricated by deep reactive ion etching (DRIE) of the silicon, with subsequent anodic bonding to

borosilicate glass of 1.75 mm thickness. The temperature-controlled region of the chip was inserted

into a stainless steel heating chuck with temperature controlled by a PID controller and three

Page 94: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

79

cartridge heaters. The full fabrication procedure for devices of this type was reported in a previous

study.156

For each experiment, a suspension of selected cysts in media was injected into the chip using a

syringe and trapped within C-shape walls. After the cells were loaded, deionized water was flowed

into the chip at a rate of 1 mL min-1 for 5 min to flush the media prior to cell disruption.

Hydrothermal disruption of cell wall at 200 ºC and 8 MPa for 10 mins was used as a standard

disruption process that provided uniform and thorough disruption to all cells. The full

hydrothermal disruption procedure for H. pluvialis cells was reported elsewhere.197

All supercritical CO2 extractions were carried out in the microfluidic chip on disrupted cells. For

all experiments, the heating chuck was adjusted to steady experimental temperatures before

pumping CO2 into the chip. The schematic diagram of the experimental setup and procedure are

illustrated in the supplementary material. The flow path of the system was controlled by a 6-port

valve with minimal dead volume (Rheodyne®7030). When switched, high pressure CO2 could be

slowly pumped into the channel without depressurizing the entire chip and avoid back flow. The

flow rate of CO2 was controlled by a back-pressure regulator (Swagelok KCP1GRA2C1P10000).

During extraction, and at all stages of astaxanthin recovery, the cells were observed using darkfield

microscopy (Olympus BXFM microscope, 10× objective).

To compare the extraction rate and efficiency of different processing conditions, the change in

coloration of individual H. pluvialis cells was quantified by image processing for all cases. The

relative astaxanthin content was determined directly from the colour images through analysis of

RGB pixel values as reported elsewhere197 (additional details in supplementary material).

6.3 Results and discussion

To allow astaxanthin to be extracted from the cell into the extraction medium, the cell wall was

first disrupted. Hydrothermal disruption was chosen here instead of chemical (acid and base),

biologic (enzyme), or mechanical (eg. grinding, milling and bead beating) approaches due to the

uniformness and completeness of the disruption process achieved.75,76,197 As outlined in Fig. 1, all

trapped cells undergo hydrothermal disruption prior to the extraction processes. The disrupted cells

were then extracted with ScCO2 at various temperatures and solvent conditions. At each stage in

the process, the astaxanthin content was approximated from the colour image data.197 To provide

Page 95: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

80

a baseline reference point for 100% extraction, all test processes were followed by acetone

extraction which has been proven as a highly effective approach to remove residual

astaxanthin79,82,84. A summary of experimental conditions and results of this study are shown in

Table 2.

Figure 6-1. Single-cellular visualization and quantification of astaxanthin extraction in

Haematococcus pluvialis using supercritical CO2 and co-solvents. Post ScCO2 and ethanol

extraction using acetone provides an overall extraction efficiency metric. Images were taken

from pure ScCO2 extraction experiment at 70 °C.

Table 6-2: Summary of supercritical carbon dioxide extraction experiments and results.

Experiment Solvents Extraction

conditions Duration

Reduction in

normalized red

content (%)

Red

content

drop rate

(%/min)

1 CO2 8 MPa 40 °C

no flow 120 min 11.9 0.07

2 CO2 8 MPa 55 °C

no flow 120 min 22.4 0.23

3 CO2 8 MPa 70 °C

no flow 120 min 63.1 0.57

4 Ethanol in CO2

= 20 % (v/v)

8 Mpa 40 °C

120 µL min-1 15 min 84.4 31.4*

5 Ethanol in CO2

= 20 % (v/v)

8 MPa 55 °C

120 µL min-1 15 min 98.3 168.5*

Page 96: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

81

6 Ethanol in CO2

= 20 % (v/v)

8 MPa 70 °C

120 µL min-1 15 min 97.8 167.6*

7 Olive oil in CO2

= 20 % (v/v)

8 MPa 55 °C

120 µL min-1 15 min 98.6 168.9*

*This rate is estimated based on the change of normalized red content in 35 seconds.

We first investigated the effects of temperature on pure ScCO2 extraction by setting the

temperature of the reaction chamber to 40 °C, 55 °C and 70 °C. For all pure CO2 extractions it was

necessary to flow CO2 (at atmospheric pressure) to remove the water surrounding the cells.

Without this drying step, water surrounding the cells effectively shielded them from ScCO2,

resulting in no noticeable extraction over long periods. Once the water was removed, we applied

an operating pressure of 8 MPa, just above the critical pressure (7.39 MPa), and a relatively low

pressure as compared to previous work in this area.79,80 No significant change in colour was

observed in the 40 °C case but a noticeable change in the amount and colour of the remaining

astaxanthin in cell was observed in the 70 °C case (the real time images are presented in Figure 6-

1). By calculating the change in normalized red content, the extraction in terms of overall

percentage is plotted over time in Figure 6-2a. In this two-hour interval, the 40 °C case resulted in

very little astaxanthin recovery (12 %) with a slow extraction rate; the 55 °C case showed improved

extraction rates with 22% of the red content being extracted; the 70 °C case showed a significant

improvement by achieving 63% drop over the same time period. The trend observed here for higher

extraction rates at higher temperatures is well supported in the literature.81,83–85

To further characterize the removal over time, the full extraction process was visualized over a 15-

hour period at 55 °C as shown in Figure 6-2b. The near-complete extraction of astaxanthin using

pure ScCO2 was achieved over this period indicating that the reported extraction efficiency is also

tightly related to extraction time. These results show that relatively low-pressure supercritical CO2

is capable of effectively complete astaxanthin extraction, and the rates are higher than those

achieved previously. However, even at the maximum temperature suitable for astaxanthin (70 °C),

the process requires hours to complete.

Page 97: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

82

Figure 6-2. a) Darkfield images H.p cells before and after ScCO2 extraction at 40 °C and 70

°C. b) The progression of normalized red content for ScCO2 extraction at 40 °C, 55 °C and

70 °C indicating higher extraction temperatures resulted in higher extraction rates. Solid

lines represent 1st order trendline fits to the experimental data with equation given by the

side. c) The normalized red content for ScCO2 extraction process at 55 °C for a complete

extraction process over 900 minutes with dark-field snapshots along the process.

Co-solvents present an opportunity to increase extraction rates beyond that of pure ScCO2. A

popular organic co-solvent used is ethanol, which is inexpensive and commonly used in food

processing. Adopting a similar experimental approach as pure ScCO2, we observed the cellular

extraction over time of disrupted H. pluvialis cells using a 4:1 CO2:ethanol flow rate ratio (100 µL

Page 98: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

83

min-1 and 25 µL min-1) with temperatures again varying between 40 °C and 70 °C. Following

disruption of the cell wall (hydrothermally in pure CO2 as before), Figure 6-3 shows the relatively

rapid removal of astaxanthin at the three test process temperatures. With addition of ethanol, the

timescales of extraction process are reduced 1800-fold from 15 hours to 30 seconds for a

ScCO2/ethanol mixture at 55 °C (video 1). Even at 40 °C, the final extraction is over 84% of the

measured red content after 1000 seconds.

Figure 6-3. The progression of normalized red content for ScCO2 extraction with ethanol

over 1000 seconds at 40 °C, 55 °C and 70 °C. Zoom-in plot for the first 30 s extraction time

indicates a rapid extraction of astaxanthin from ScCO2 with ethanol at 55 °C and 70 °C.

The rapid extraction of astaxanthin was observed around 4 minutes after we started pumping

ethanol which is approximately the amount of time required to replace the dead volume (~500 µL)

at the given flow rate (125 µL/min). The enhanced extraction with ethanol is attributed to two

factors. First, water is miscible with the ethanol phase, removing the two-phase interfacial barrier

between the solvent and the water-wet surface of the cell. Secondly, the solubility of astaxanthin

in ethanol is 5 orders of magnitude higher than that in ScCO2 .273

Olive oil is another attractive ScCO2 co-solvent due its potential to be directly combined with

astaxanthin for use in food products, avoiding the need for the solvent separation step required

Page 99: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

84

with ethanol. Olive oil has been shown to achieve extraction rates comparable to ethanol.81 As

shown in Figure 6-4a the addition of olive oil to the extraction fluid instead of ethanol also

performed significantly better than pure ScCO2 and reached complete extraction (~98%) in just

180 seconds. Images at various stages of the extraction process are shown in Figure 6-4a for an

olive oil co-solvent at 55 °C.

In contrast to the ethanol case, the removal of the water barrier by olive oil is observed as oil

immediately displacing the surrounding water phase. In this process, three phases can be observed

within the microchannel: ScCO2, oil and a small amount of water left from hydrothermal process.

Upon interaction with the disrupted cell-wall, the introduction of olive oil results in a number of

emulsions spontaneously appearing immediately around the cell. Rather than the formation of an

oil-water-astaxanthin mixture, the remaining water in the cell instead emulsifies into smaller

droplets and can be seen during the initial extraction phase (Figure 6-4b). Similar to the ethanol

case the increased solubility of astaxanthin versus ScCO2 is readily apparent by the rapid increase

in red content in the co-solvent phase envelope around the C-trap (video 2). While the colour

increases within this region, there is no evidence of colour loss to the surrounding CO2 fluid until

the fluid flow forces the oil droplet forward along the channel, leaving a depleted H. Pluvialis cell.

Page 100: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

85

Figure 6-4. a) The comparison of normalized red content for ScCO2 extraction with ethanol

and olive oil at 55 °C over 200 seconds. Time-lapsed snapshots of the extraction process with

olive oil are provided. b) Snapshot of ScCO2 extraction with olive oil indicating three phases:

ScCO2, olive oil and water. The boundary of olive oil and CO2 phases are illustrated using

yellow and green dash lines.

Page 101: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

86

Some differences between the results of the experiments here and those of previous studies are

noteworthy. First, with respect to the pure ScCO2 test cases, the results here show 92% extraction

in 15 hours at 55 ºC and 8 MPa, with extraction rates increasing with temperature. Other

studies81,84,85 have shown the highest total extraction level plateaued at 25% over 5 hours, at 70 ºC

and 40 MPa. We attribute our improved performance in these pure CO2 cases to the high degree

of disruption achieved prior to extraction and the precise control of conditions. Second, with

respect to co-solvent tests, we achieved full extraction within minutes which is in marked contrast

to other results showing comparable results only after several hours of extraction81,84,85,255. Here we

attribute the improved performance to within the reactor – namely that all cells had access to co-

solvents (in addition to the high degree of prior cell disruption noted earlier). These results point

to the possibility of extraction with significantly lower pressures, provided the reactor achieves a

high degree of cell-to-co-solvent contact. In the context of larger scale reactors, these results point

the potential to reduce pressures and processing time - key drivers of capital and operational costs.

6.4 Conclusions

In this work we demonstrate low pressure supercritical CO2 extraction of disrupted

Haematococcus pluvialis in a microfluidic reactor with optical access to monitor the extraction

process in real time. A near-complete (~98%) rapid extraction of astaxanthin using ScCO2 and an

ethanol co-solvent was achieved in 30 seconds at 8 MPa, representing the fastest complete

astaxanthin extraction at such low pressures. Our results using a single cell reactor provide time

resolved and direct evidence of practical astaxanthin extraction using supercritical CO2 at low

pressures in a matter of minutes, highlighting the potential for fast and full recovery of this valuable

bio-product.

6.5 Supplementary Information

The experimental apparatus was adapted from our previous research with different arrangements

specifically designed for supercritical CO2 extraction. The fluid flow path used is shown in Figure

6-5. At the beginning of the experiment, Haematococcus pluvialis (H.p) cells were loaded into the

microfluidic chip before connecting the chip to the flow system. Prior to the hydrothermal

disruption process, the channels downstream to the microfluidic reactor were pressurized to 8 MPa

with CO2 to avoid abrupt fluid flow due to depressurization when turning the switching valve [3].

Three high pressure pumps were used in the experiment: two water pumps [1a and 1c] and a CO2

Page 102: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

87

pump [1b]. Deionised water in pump [1a] was used to flush the media and pressurize the system

to 8 MPa for the hydrothermal disruption process After hydrothermal disruption, the switching

valve [3] was switched to position 2 to supply supercritical CO2 for the extraction stage. Co-solvent

(ethanol or olive oil) was stored in the piston cylinder [2] and added to the system using a

controlling valve [5a]. Both pumps [1b & 1c] were running at constant flow rates during the

extraction process.

Figure 6-5. Schematic diagram of the experimental setup with the flow direction indicated

by arrows along the processing path. The flow of the switching valve at two positions is shown

using red lines.

To compare the extraction efficiency of astaxanthin for different processes, the change in

coloration of individual H. pluvialis cells was quantified by image processing for all cases. The

area of interest (AOI) was manually selected based on the size of observed cells and used

throughout the entire extraction process. The equation used to convert the 8-bit RGB values of

each pixel in the AOI into a coloration-based global “extracted red content” was adopted from

previous study and shown below:

𝐴 = (100 −0.59𝐵 + 0.41𝐺

2.55) ∗ (1.36

255−𝑅45 )

Page 103: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

88

where R, G and B represents the red, blue and green value at that pixel, respectively. The first and

second factors were determined by the maximum astaxanthin absorbance peak at 475 nm and

empirical correlation between R-value and astaxanthin concentration (a drop of 45 in the R-value

corresponds to a 36% increase in astaxanthin concentration), respectively. The AOI was tracked

to calculate the extraction efficiency with respect to time according to (Ai-AF)/Ai where Ai and

AF represent the initial and final red content. Typical colors of a H. pluvialis cell observed during

extraction were selected and are illustrated in Figure 6-6 with RGB values, assigned astaxanthin

content and calculated astaxanthin content based on the given equation showing on the same row.

The small difference between the assigned astaxanthin content and calculated astaxanthin content

indicates a fair approximation using the given equation.

Figure 6-6. Illustration of typical colors of H. pluvialis cells with different concentration of

astaxanthin and the calculated RGB-based astaxanthin content.

Page 104: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

89

Chapter 7.

Conclusions

7.1 Summary

Maximizing productivity from microalgal biomass requires innovative approaches to address the

central challenges in downstream processing. I have developed and applied a microfluidic platform

capable of performing reactions at high temperature (350 °C) and high pressure (20 MPa), precise

control on processing parameters, and providing optical access for in-situ observation and

quantification. The direct observation of microalgae downstream processes was made possible for

the first time and allowed visual characterization, fluorescence spectroscopy, and quantitative

imaging of the conversion at the single-cell scale during all stages of the reaction.

Besides the engineering achievement of successfully developed this microfluidic platform, a

significant contribution in science was also made to improve microalgae downstream processes.

In the hydrothermal liquefaction of microalgae project, biomass-to-biocrude conversion was

directly observed in a microfluidic channel and the reaction mechanism was indicated by a high

time-resolution change of fluorescence signature in 10 mins. Formation of oil droplets comprised

of aliphatic components was directly observed in a short reaction time suggesting that cell wall

disruption takes place within the early stages of these reactions. Inspired by the findings from HTL

processes, a hydrothermal method was introduced to disrupt rigid cell walls in Haematococcus

pluvialis for astaxanthin recovery. Hydrothermal disruption at 200 °C was the most effective wall

disruption technique, enabling near complete astaxanthin extraction from wet biomass, a

significant improvement over traditional methods. For ScCO2 extraction of astaxanthin, the degree

of disruption, water film surrounding the cells and mass transport showed dominating effects on

the performance of extraction process. Astaxanthin extraction using ScCO2 achieved 92%

recovery at 55 °C and 8 MPa applied over 15 hours. With the addition of co-solvents, ethanol and

olive oil, the timescales of extraction process are reduced dramatically from 15 hours to a few

minutes, representing the fastest complete astaxanthin extraction at such low pressures.

Page 105: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

90

These results and the microfluidic platform on which they were collected represent the first of their

kind in the field of microalgae downstream processing. Although, conventional batch reactors have

competitive advantages on quantifying product yields which requires relatively large volumes of

products, microfluidic methods presented here offer a unique toolset to better understand the

processes with high-resolution information at cellular level, at specific operating conditions such

temperature and pressure. This level of insight has simply not been possible with previous

conventional reactors. The results presented in this thesis provide new insight into important

biomass-to-bioproducts conversion processes – insight that can be applied to improve large scale

operations. Perhaps most notably, the fast reaction rates achieved with this reactor (in some cases

at lower pressures than previously employed) highlight what is possible in larger reactors –

provided that excellent transport is achieved.

7.2 Future Outlook

As discussed earlier, converting biomass into valuable products is a critical step for producing

bioproducts from microalgae. The results presented in this thesis provided unprecedented insights

into these downstream processes but also proposed future challenges. Effectively use this

information to achieve the same mass and thermal transport shown in the microfluidic reactors at

large scale is the next challenge. Also, going down the path of bioproducts production, there are

many other challenges remained. Successfully making biocrude at low cost is just halfway through

and upgrading the biocrude to drop-in fuels at industrial scale still require major breakthroughs.

Perhaps most notably, the biocrude produced in this thesis work is highly oxygenated and thus

requires more downstream processing than, say, conventional crude oil. The purity of high value

products from microalgae is also a rising concern due to different downstream pathways were

used. Standardized testing protocols are expected for future microalgae products.

For both bioenergy and high-value bioproduct production, a more holistic approach over a lipid-

centric approach is needed to maximize all the potential value contained in the biomass. Looking

forward I see two major opportunities for improving the performance of biomass-to-biofuel and

biomass-to-bioproducts conversions respectively.

First, a clearer understanding of the reaction mechanism of HTL will become available by

investigating the effects of processing parameters on individual biomass components such as

carbohydrates and proteins. In particular, the reaction time associated with reaction temperature

Page 106: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

91

and pressure is critical to the design of a sequential process such that each step is finely tuned to

optimize the yield of a targeted product from a specific feedstock. For example, fast hydrothermal

processes could be used as a cell wall disruption technique that enables access to high energy and

high value components – lipids. Once the lipids are recovered, the remaining biomass could then

be used as animal feed or further converted to bioenergy using a more target-specific conversion

method for proteins and carbohydrates. Work presented in Chapter 4 will help provide vital

insights towards optimizing the output of hydrothermal processes, particularly in cases where high

time-resolution information is required to carefully tune the processing parameters.

Second, technology developed during the commercialization of high-value bioproducts from

microalgae will help drive the development of microalgal biofuel. The downstream processes of

both applications share similar challenges – high water content in the feedstock. The breakthroughs

in cell-wall disruption and product extraction made from commercial cases will help to reduce the

costs of biofuel production helping make it an economically viable alternative to fossil fuels.

Further integration and development of utilizing waste streams, such as nutrients from wastewater

and energy from waste heat or geothermal, can further provide value to the competition with fossil

fuels.

Finally, challenges and exciting opportunities both exist for improving the performance of

downstream processing. The markets open to microalgal products are vast and diverse. Low-

volume high-value bioproduct markets are the best initial target with biofuels being the ultimate

long-term high-volume market. Collectively these opportunities motivate near-term advances to

overcome challenges in the downstream processing, and to maximize the overall recoverable value

from microalgae biomass.

Page 107: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

92

References

1. UNEP. United Nations Environment Programm Annual Report 2015.

2. BP. BP Statistical Review of World Energy 2017. (2017).

3. International Energy Agency. How2Guide for Bioenergy Roadmap Development and

Implementation. (2017).

4. International Renewable Energy Agency (IRENA). Global Bioenergy supply and demand

projections. (2014).

5. Xu, L., Wim Brilman, D. W. F., Withag, J. a M., Brem, G. & Kersten, S. Assessment of a

dry and a wet route for the production of biofuels from microalgae: Energy balance

analysis. Bioresour. Technol. 102, 5113–5122 (2011).

6. Lardon, L., Helias, A., Sialve, B., Steyer, J.-P. & Bernard, O. Life-Cycle Assessment of

Biodiesel Production from Microalgae. Environ. Sci. Technol. 43, 6475–6481 (2009).

7. Zhou, Y., Schideman, L., Yu, G. & Zhang, Y. A synergistic combination of algal

wastewater treatment and hydrothermal biofuel production maximized by nutrient and

carbon recycling. Energy Environ. Sci. 6, 3765 (2013).

8. Sills, D. L. et al. Quantitative Uncertainty Analysis of Life Cycle Assessment for Algal

Biofuel Production. Environ. Sci. Technol. 47, 687–694 (2013).

9. Subhadra, B. G. Sustainability of algal biofuel production using integrated renewable

energy park (IREP) and algal biorefinery approach. Energy Policy 38, 5892–5901 (2010).

10. Tian, C., Li, B., Liu, Z., Zhang, Y. & Lu, H. Hydrothermal liquefaction for algal

biorefinery: A critical review. Renew. Sustain. Energy Rev. 38, 933–950 (2014).

11. Yin, S. & Tan, Z. Hydrothermal liquefaction of cellulose to bio-oil under acidic, neutral

and alkaline conditions. Appl. Energy 92, 234–239 (2012).

12. Biller, P. & Ross, A. B. Potential yields and properties of oil from the hydrothermal

liquefaction of microalgae with different biochemical content. Bioresour. Technol. 102,

215–225 (2011).

13. Mok, W. S. L. & Antal, M. J. Uncatalyzed solvolysis of whole biomass hemicellulose by

hot compressed liquid water. Ind. Eng. Chem. Res. 31, 1157–1161 (1992).

14. Kamio, E. et al. Liquefaction kinetics of cellulose treated by hot compressed water under

variable temperature conditions. J. Mater. Sci. 43, 2179–2188 (2008).

15. Ooms, M. D., Dinh, C. T., Sargent, E. H. & Sinton, D. Photon management for augmented

photosynthesis. Nat. Commun. 7, 12699 (2016).

16. Sinton, D. Energy: the microfluidic frontier. Lab Chip 14, 3127 (2014).

17. Uhlen, M. & Andersson Svahn, H. Lab on a chip technologies for bioenergy and

biosustainability research. Lab Chip 11, 3389 (2011).

18. Erickson, D., Sinton, D. & Psaltis, D. Optofluidics for energy applications. Nat. Photonics

5, 583–590 (2011).

19. Günther, A. & Jensen, K. F. Multiphase microfluidics: from flow characteristics to

Page 108: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

93

chemical and materials synthesis. Lab Chip 6, 1487–1503 (2006).

20. Marre, S., Roig, Y. & Aymonier, C. Supercritical microfluidics: Opportunities in flow-

through chemistry and materials science. J. Supercrit. Fluids 66, 251–264 (2012).

21. Marre, S., Adamo, A., Basak, S., Aymonier, C. & Jensen, K. F. Design and Packaging of

Microreactors for High Pressure and High Temperature Applications. Ind. Eng. Chem.

Res. 49, 11310–11320 (2010).

22. Pinho, B. et al. A microfluidic approach for investigating multicomponent system

thermodynamics at high pressures and temperatures. Lab Chip 14, 3843 (2014).

23. Ragauskas, A. J. The Path Forward for Biofuels and Biomaterials. Science (80-. ). 311,

484–489 (2006).

24. Romanowsky, M. B., Abate, A. R., Rotem, A., Holtze, C. & Weitz, D. A. High throughput

production of single core double emulsions in a parallelized microfluidic device. Lab Chip

12, 802 (2012).

25. Shepherd, R. F. et al. Microfluidic assembly of homogeneous and janus colloid-filled

hydrogel granules. Langmuir 22, 8618–8622 (2006).

26. Xu, S. et al. Generation of Monodisperse Particles by Using Microfluidics: Control over

Size, Shape, and Composition. Angew. Chemie Int. Ed. 44, 724–728 (2005).

27. Moore, J. S. & Jensen, K. F. Automated Multitrajectory Method for Reaction

Optimization in a Microfluidic System using Online IR Analysis. Org. Process Res. Dev.

16, 1409–1415 (2012).

28. Marre, S. & Jensen, K. F. Synthesis of micro and nanostructures in microfluidic systems.

Chem. Soc. Rev. 39, 1183 (2010).

29. Chan, E. M., Alivisatos, A. P. & Mathies, R. A. High-Temperature Microfluidic Synthesis

of CdSe Nanocrystals in Nanoliter Droplets. J. Am. Chem. Soc. 127, 13854–13861 (2005).

30. International Energy Agency. Key World Energy Trends 2016. (2016).

31. Searchinger, T. et al. Use of U.S. Croplands for Biofuels Increases Greenhouse Gases

Through Emissions from Land-Use Change. 423, 1238–1240 (2008).

32. Chisti, Y. Biodiesel from microalgae. Biotechnol. Adv. 25, 294–306 (2007).

33. Georgianna, D. R. & Mayfield, S. P. Exploiting diversity and synthetic biology for the

production of algal biofuels. Nature 488, 329–335 (2012).

34. Slade, R. & Bauen, A. Micro-algae cultivation for biofuels: Cost, energy balance,

environmental impacts and future prospects. Biomass and Bioenergy 53, 29–38 (2013).

35. Guilford, M. C., Hall, C. a S., O&apos;Connor, P. & Cleveland, C. J. A new long term

assessment of energy return on investment (EROI) for U.S. oil and gas discovery and

production. Sustainability 3, 1866–1887 (2011).

36. Mu, D. et al. Life Cycle Environmental Impacts of Wastewater-Based Algal Biofuels.

Environ. Sci. Technol. 48, 11696–11704 (2014).

37. Zhu, Y., Biddy, M. J., Jones, S. B., Elliott, D. C. & Schmidt, A. J. Techno-economic

analysis of liquid fuel production from woody biomass via hydrothermal liquefaction

(HTL) and upgrading. Appl. Energy 129, 384–394 (2014).

Page 109: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

94

38. Ou, L., Thilakaratne, R., Brown, R. C. & Wright, M. M. Techno-economic analysis of

transportation fuels from defatted microalgae via hydrothermal liquefaction and

hydroprocessing. Biomass and Bioenergy 72, 45–54 (2015).

39. Elliott, D. C., Biller, P., Ross, A. B., Schmidt, A. J. & Jones, S. B. Hydrothermal

liquefaction of biomass: Developments from batch to continuous process. Bioresour.

Technol. 178, 147–156 (2014).

40. Akiya, N. & Savage, P. E. Roles of Water for Chemical Reactions in High-Temperature

Water. Chem. Rev. 102, 2725–2750 (2002).

41. Hoffmann, M. M. & Conradi, M. S. Are There Hydrogen Bonds in Supercritical Water? J.

Am. Chem. Soc. 119, 3811–3817 (1997).

42. Kruse, A. & Dahmen, N. Water – A magic solvent for biomass conversion. J. Supercrit.

Fluids 96, 36–45 (2015).

43. Zhang, Y. in Biofuels from Agricultural Wastes and Byproducts 201–232 (Wiley-

Blackwell, 2010).

44. Connolly, J. F. Solubility of Hydrocarbons in Water Near the Critical Solution

Temperatures. J. Chem. Eng. Data 11, 13–16 (1966).

45. Rebert, C. J. & Kay, W. B. The phase behavior and solubility relations of the benzene-

water system. AIChE J. 5, 285–289 (1959).

46. Peterson, A. A. et al. Thermochemical biofuel production in hydrothermal media: A

review of sub- and supercritical water technologies. Energy Environ. Sci. 1, 32 (2008).

47. Jin, F., Wang, Y., Zeng, X., Shen, Z. & Yao, G. in Application of Hydrothermal Reactions

to Biomass Conversion (ed. Jin, F.) 3–28 (Springer Berlin Heidelberg, 2014).

doi:10.1007/978-3-642-54458-3_1

48. Xu, C. C. et al. Application of Hydrothermal Reactions to Biomass Conversion. (Springer

Berlin Heidelberg, 2014). doi:10.1007/978-3-642-54458-3

49. Lin, H., Su, J., Liu, Y. & Yang, L. Application of Hydrothermal Reactions to Biomass

Conversion. Application of Hydrothermal Reactions to Biomass Conversion (Springer

Berlin Heidelberg, 2014). doi:10.1007/978-3-642-54458-3

50. Williams, P. J. L. B. & Laurens, L. M. L. Microalgae as biodiesel &amp; biomass

feedstocks: Review &amp; analysis of the biochemistry, energetics &amp; economics.

Energy Environ. Sci. 3, 554 (2010).

51. Teri, G., Luo, L. & Savage, P. E. Hydrothermal Treatment of Protein, Polysaccharide, and

Lipids Alone and in Mixtures. Energy & Fuels 28, 7501–7509 (2014).

52. Licitra, G., Hernandez, T. M. & Van Soest, P. J. Standardization of procedures for

nitrogen fractionation of ruminant feeds. Anim. Feed Sci. Technol. 57, 347–358 (1996).

53. Bi, Z. & He, B. B. Characterization of Microalgae for the Purpose of Biofuel Production.

Trans. ASABE 56, 1529–1539 (2013).

54. Valdez, P. J., Nelson, M. C., Wang, H. Y., Lin, X. N. & Savage, P. E. Hydrothermal

liquefaction of Nannochloropsis sp.: Systematic study of process variables and analysis of

the product fractions. Biomass and Bioenergy 46, 317–331 (2012).

Page 110: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

95

55. Vardon, D. R. et al. Chemical properties of biocrude oil from the hydrothermal

liquefaction of Spirulina algae, swine manure, and digested anaerobic sludge. Bioresour.

Technol. 102, 8295–8303 (2011).

56. Brown, T. M., Duan, P. & Savage, P. E. Hydrothermal Liquefaction and Gasification of

Nannochloropsis sp. Energy & Fuels 24, 3639–3646 (2010).

57. Leow, S. et al. Prediction of microalgae hydrothermal liquefaction products from

feedstock biochemical composition. Green Chem. 17, 3584–3599 (2015).

58. Holliday, R. L., King, J. W. & List, G. R. Hydrolysis of Vegetable Oils in Sub- and

Supercritical Water. Ind. Eng. Chem. Res. 36, 932–935 (1997).

59. King, J. W., Holliday, R. L. & List, G. R. Hydrolysis of soybean oil. Green Chem. 1, 261–

264 (1999).

60. Peterson, A. a. et al. Thermochemical biofuel production in hydrothermal media: A

review of sub- and supercritical water technologies. Energy Environ. Sci. 1, 32 (2008).

61. Rogalinski, T., Herrmann, S. & Brunner, G. Production of amino acids from bovine serum

albumin by continuous sub-critical water hydrolysis. J. Supercrit. Fluids 36, 49–58

(2005).

62. Salak Asghari, F. & Yoshida, H. Acid-Catalyzed Production of 5-Hydroxymethyl Furfural

from <scp>d</scp> -Fructose in Subcritical Water. Ind. Eng. Chem. Res. 45, 2163–2173

(2006).

63. Kabyemela, B. M., Adschiri, T., Malaluan, R. M. & Arai, K. Glucose and Fructose

Decomposition in Subcritical and Supercritical Water: Detailed Reaction Pathway,

Mechanisms, and Kinetics. Ind. Eng. Chem. Res. 38, 2888–2895 (1999).

64. Rogalinski, T., Liu, K., Albrecht, T. & Brunner, G. Hydrolysis kinetics of biopolymers in

subcritical water. J. Supercrit. Fluids 46, 335–341 (2008).

65. Faeth, J. L., Valdez, P. J. & Savage, P. E. Fast Hydrothermal Liquefaction of

Nannochloropsis sp. To Produce Biocrude. Energy & Fuels 27, 1391–1398 (2013).

66. Bach, Q.-V., Sillero, M. V., Tran, K.-Q. & Skjermo, J. Fast hydrothermal liquefaction of a

Norwegian macro-alga: Screening tests. Algal Res. 6, 271–276 (2014).

67. Guo, Y., Yeh, T., Song, W., Xu, D. & Wang, S. A review of bio-oil production from

hydrothermal liquefaction of algae. Renew. Sustain. Energy Rev. 48, 776–790 (2015).

68. Akhtar, J. & Amin, N. A. S. A review on process conditions for optimum bio-oil yield in

hydrothermal liquefaction of biomass. Renew. Sustain. Energy Rev. 15, 1615–1624

(2011).

69. Ruiz, J. et al. Towards industrial products from microalgae. Energy Environ. Sci. 9, 3036–

3043 (2016).

70. Kidd, P. Astaxanthin, cell membrane nutrient with diverse clinical benefits and anti-aging

potential. Altern. Med. Rev. 16, 355–64 (2011).

71. Guerin, M., Huntley, M. E. & Olaizola, M. Haematococcus astaxanthin: applications for

human health and nutrition. Trends Biotechnol. 21, 210–216 (2003).

72. Shah, M. M. R., Liang, Y., Cheng, J. J. & Daroch, M. Astaxanthin-Producing Green

Page 111: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

96

Microalga Haematococcus pluvialis: From Single Cell to High Value Commercial

Products. Front. Plant Sci. 7, 531 (2016).

73. Wayama, M. et al. Three-Dimensional Ultrastructural Study of Oil and Astaxanthin

Accumulation during Encystment in the Green Alga Haematococcus pluvialis. PLoS One

8, e53618 (2013).

74. HAGEN, C., SIEGMUND, S. & BRAUNE, W. Ultrastructural and chemical changes in

the cell wall of Haematococcus pluvialis (Volvocales, Chlorophyta) during aplanospore

formation. Eur. J. Phycol. 37, S0967026202003669 (2002).

75. Mendes-Pinto, M. M., Raposo, M. F. J., Bowen, J., Young, A. J. & Morais, R. Evaluation

of different cell disruption processes on encysted cells of Haematococcus pluvialis: effects

on astaxanthin recovery and implications for bio-availability. J. Appl. Phycol. 13, 19–24

(2001).

76. Kim, D.-Y. et al. Cell-wall disruption and lipid/astaxanthin extraction from microalgae:

Chlorella and Haematococcus. Bioresour. Technol. 199, 300–310 (2016).

77. da Silva, R. P. F. F., Rocha-Santos, T. A. P. & Duarte, A. C. Supercritical fluid extraction

of bioactive compounds. TrAC Trends Anal. Chem. 76, 40–51 (2016).

78. Yen, H., Yang, S., Chen, C., Jesisca & Chang, J. Supercritical fluid extraction of valuable

compounds from microalgal biomass. Bioresour. Technol. 184, 291–296 (2015).

79. Nobre, B. et al. Supercritical carbon dioxide extraction of astaxanthin and other

carotenoids from the microalga Haematococcus pluvialis. Eur. Food Res. Technol. 223,

787–790 (2006).

80. Valderrama, J. O., Perrut, M. & Majewski, W. Extraction of Astaxantine and

Phycocyanine from Microalgae with Supercritical Carbon Dioxide. J. Chem. Eng. Data

48, 827–830 (2003).

81. Krichnavaruk, S., Shotipruk, A., Goto, M. & Pavasant, P. Supercritical carbon dioxide

extraction of astaxanthin from Haematococcus pluvialis with vegetable oils as co-solvent.

Bioresour. Technol. 99, 5556–5560 (2008).

82. Thana, P. et al. Response surface methodology to supercritical carbon dioxide extraction

of astaxanthin from Haematococcus pluvialis. Bioresour. Technol. 99, 3110–3115 (2008).

83. Pan, J.-L., Wang, H.-M., Chen, C.-Y. & Chang, J.-S. Extraction of astaxanthin from

Haematococcus pluvialis by supercritical carbon dioxide fluid with ethanol modifier. Eng.

Life Sci. 12, 638–647 (2012).

84. Reyes, F. A., Mendiola, J. A., Ibañez, E. & del Valle, J. M. Astaxanthin extraction from

Haematococcus pluvialis using CO2-expanded ethanol. J. Supercrit. Fluids 92, 75–83

(2014).

85. Machmudah, S., Shotipruk, A., Goto, M., Sasaki, M. & Hirose, T. Extraction of

Astaxanthin from Haematococcus p luvialis Using Supercritical CO 2 and Ethanol as

Entrainer. Ind. Eng. Chem. Res. 45, 3652–3657 (2006).

86. Elvira, K. S., i Solvas, X. C., Wootton, R. C. R. & DeMello, A. J. The past, present and

potential for microfluidic reactor technology in chemical synthesis. Nat. Chem. 5, 905–

915 (2013).

Page 112: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

97

87. Hjort, K. High-pressure microfluidics. in (eds. Gray, B. L. & Becker, H.) 9320, 93200V

(2015).

88. Hessel, V., Kralisch, D., Kockmann, N., Noël, T. & Wang, Q. Novel Process Windows for

Enabling, Accelerating, and Uplifting Flow Chemistry. ChemSusChem 6, 746–789 (2013).

89. Luther, S. K. & Braeuer, A. High-pressure microfluidics for the investigation into multi-

phase systems using the supercritical fluid extraction of emulsions (SFEE). J. Supercrit.

Fluids 65, 78–86 (2012).

90. Assmann, N., Kaiser, S. & Rudolf von Rohr, P. Supercritical extraction of vanillin in a

microfluidic device. J. Supercrit. Fluids 67, 149–154 (2012).

91. Pierobon, S. C. et al. Emerging microalgae technology: a review. Sustain. Energy Fuels 0,

1–26 (2017).

92. Mata, T. M., Martins, A. A. & Caetano, N. S. Microalgae for biodiesel production and

other applications: A review. Renew. Sustain. Energy Rev. 14, 217–232 (2010).

93. Lam, M. K. & Lee, K. T. Microalgae biofuels: A critical review of issues, problems and

the way forward. Biotechnol. Adv. 30, 673–690 (2012).

94. Chen, X. et al. Ionic liquid-assisted subcritical water promotes the extraction of lipids

from wet microalgae Scenedesmus sp. Eur. J. Lipid Sci. Technol. 117, 1192–1198 (2015).

95. Barros, A. I., Gonçalves, A. L., Simões, M. & Pires, J. C. M. Harvesting techniques

applied to microalgae: A review. Renew. Sustain. Energy Rev. 41, 1489–1500 (2015).

96. Uduman, N., Qi, Y., Danquah, M. K., Forde, G. M. & Hoadley, A. Dewatering of

microalgal cultures: A major bottleneck to algae-based fuels. J. Renew. Sustain. Energy 2,

12701 (2010).

97. DOE (U.S. Department of Energy). National Algal Biofuels Technology Review. Natl.

Algal Biofuels Technol. Rev. (2016).

98. Coons, J. E., Kalb, D. M., Dale, T. & Marrone, B. L. Getting to low-cost algal biofuels: A

monograph on conventional and cutting-edge harvesting and extraction technologies.

Algal Res. 6, 250–270 (2014).

99. Milledge, J. J. & Heaven, S. A review of the harvesting of micro-algae for biofuel

production. Rev. Environ. Sci. Bio/Technology 12, 165–178 (2013).

100. Brennan, L. & Owende, P. Biofuels from microalgae-A review of technologies for

production, processing, and extractions of biofuels and co-products. Renew. Sustain.

Energy Rev. 14, 557–577 (2010).

101. Pragya, N., Pandey, K. K. & Sahoo, P. K. A review on harvesting, oil extraction and

biofuels production technologies from microalgae. Renew. Sustain. Energy Rev. 24, 159–

171 (2013).

102. Kim, J. et al. Methods of downstream processing for the production of biodiesel from

microalgae. Biotechnol. Adv. 31, 862–876 (2013).

103. Ndikubwimana, T. et al. Flotation: A promising microalgae harvesting and dewatering

technology for biofuels production. Biotechnol. J. 11, 315–326 (2016).

104. Zhang, X., Wang, L., Sommerfeld, M. & Hu, Q. Harvesting microalgal biomass using

Page 113: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

98

magnesium coagulation-dissolved air flotation. Biomass and Bioenergy 93, 43–49 (2016).

105. Irving, T. E. & Allen, D. G. Species and material considerations in the formation and

development of microalgal biofilms. Appl. Microbiol. Biotechnol. 92, 283–294 (2011).

106. Pierobon, S. C., Ooms, M. D. & Sinton, D. Evanescent cultivation of photosynthetic

bacteria on thin waveguides. J. Micromechanics Microengineering 24, 45017 (2014).

107. Sander, K. & Murthy, G. S. Life cycle analysis of algae biodiesel. Int. J. Life Cycle

Assess. 15, 704–714 (2010).

108. Weschler, M. K., Barr, W. J., Harper, W. F. & Landis, A. E. Process energy comparison

for the production and harvesting of algal biomass as a biofuel feedstock. Bioresour.

Technol. 153, 108–115 (2014).

109. Sim, T.-S., Goh, A. & Becker, E. W. Comparison of centrifugation, dissolved air flotation

and drum filtration techniques for harvesting sewage-grown algae. Biomass 16, 51–62

(1988).

110. Molina Grima, E., Belarbi, E.-H., Acién Fernández, F. ., Robles Medina, A. & Chisti, Y.

Recovery of microalgal biomass and metabolites: process options and economics.

Biotechnol. Adv. 20, 491–515 (2003).

111. Christenson, L. & Sims, R. Production and harvesting of microalgae for wastewater

treatment, biofuels, and bioproducts. Biotechnol. Adv. 29, 686–702 (2011).

112. Chen, C.-Y., Yeh, K.-L., Aisyah, R., Lee, D.-J. & Chang, J.-S. Cultivation,

photobioreactor design and harvesting of microalgae for biodiesel production: A critical

review. Bioresour. Technol. 102, 71–81 (2011).

113. Bilad, M. R., Vandamme, D., Foubert, I., Muylaert, K. & Vankelecom, I. F. J. Harvesting

microalgal biomass using submerged microfiltration membranes. Bioresour. Technol. 111,

343–352 (2012).

114. Zhao, F. et al. Comparison of axial vibration membrane and submerged aeration

membrane in microalgae harvesting. Bioresour. Technol. 208, 178–183 (2016).

115. PRAKASH, J. et al. Microalgal Biomass Drying By a Simple Solar Device∗ . Int. J. Sol.

Energy 18, 303–311 (1997).

116. Fudholi, A., Sopian, K., Othman, M. Y. & Ruslan, M. H. Energy and exergy analyses of

solar drying system of red seaweed. Energy Build. 68, 121–129 (2014).

117. Brink, J. & Marx, S. Harvesting of Hartbeespoort Dam micro-algal biomass through sand

filtration and solar drying. Fuel 106, 67–71 (2013).

118. Show, K.-Y., Lee, D.-J., Tay, J.-H., Lee, T.-M. & Chang, J.-S. Microalgal drying and cell

disruption – Recent advances. Bioresour. Technol. 184, 258–266 (2015).

119. Ryckebosch, E., Muylaert, K., Eeckhout, M., Ruyssen, T. & Foubert, I. Influence of

Drying and Storage on Lipid and Carotenoid Stability of the Microalga Phaeodactylum

tricornutum. J. Agric. Food Chem. 59, 11063–11069 (2011).

120. Guldhe, A., Singh, B., Rawat, I., Ramluckan, K. & Bux, F. Efficacy of drying and cell

disruption techniques on lipid recovery from microalgae for biodiesel production. Fuel

128, 46–52 (2014).

Page 114: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

99

121. Vairappan, C. S., Razalie, R., Elias, U. M. & Ramachandram, T. Effects of improved post-

harvest handling on the chemical constituents and quality of carrageenan in red alga,

Kappaphycus alvarezii Doty. J. Appl. Phycol. 26, 909–916 (2014).

122. Li, H. et al. Evaluation of a biomass drying process using waste heat from process

industries: A case study. Appl. Therm. Eng. 35, 71–80 (2012).

123. Lee, J. Y., Yoo, C., Jun, S. Y., Ahn, C. Y. & Oh, H. M. Comparison of several methods

for effective lipid extraction from microalgae. Bioresour. Technol. 101, S75–S77 (2010).

124. Pasquet, V. et al. Study on the microalgal pigments extraction process: Performance of

microwave assisted extraction. Process Biochem. 46, 59–67 (2011).

125. Iqbal, J. & Theegala, C. Microwave assisted lipid extraction from microalgae using

biodiesel as co-solvent. Algal Res. 2, 34–42 (2013).

126. Pan, J. et al. Microwave-assisted extraction of lipids from microalgae using an ionic liquid

solvent [BMIM][HSO4]. Fuel 178, 49–55 (2016).

127. Mendes, R. L., Nobre, B. P., Cardoso, M. T., Pereira, A. P. & Palavra, A. F. Supercritical

carbon dioxide extraction of compounds with pharmaceutical importance from

microalgae. Inorganica Chim. Acta 356, 328–334 (2003).

128. Brunner, G. Supercritical fluids: technology and application to food processing. J. Food

Eng. 67, 21–33 (2005).

129. Jessop, P. P. G., Heldebrant, D. J. D., Li, X., Eckert, C. A. C. & Liotta, C. C. L. Green

chemistry: Reversible nonpolar-to-polar solvent. Nature 436, 1102 (2005).

130. Richter, B. E. et al. Accelerated Solvent Extraction: A Technique for Sample Preparation.

Anal. Chem. 68, 1033–1039 (1996).

131. Hejazi, M. A. et al. Selective extraction of carotenoids from the microalga Dunaliella

salina with retention of viability. Biotechnol. Bioeng. 79, 29–36 (2002).

132. Levine, R. B., Pinnarat, T. & Savage, P. E. Biodiesel Production from Wet Algal Biomass

through in Situ Lipid Hydrolysis and Supercritical Transesterification. Energy & Fuels 24,

5235–5243 (2010).

133. Umdu, E. S., Tuncer, M. & Seker, E. Transesterification of Nannochloropsis oculata

microalga’s lipid to biodiesel on Al2O3 supported CaO and MgO catalysts. Bioresour.

Technol. 100, 2828–2831 (2009).

134. Nagle, N. & Lemke, P. Production of methyl ester fuel from microalgae. Appl. Biochem.

Biotechnol. 24–25, 355–361 (1990).

135. Huerlimann, R., de Nys, R. & Heimann, K. Growth, lipid content, productivity, and fatty

acid composition of tropical microalgae for scale-up production. Biotechnol. Bioeng. 107,

245–257 (2010).

136. Tran, D.-T., Yeh, K.-L., Chen, C.-L. & Chang, J.-S. Enzymatic transesterification of

microalgal oil from Chlorella vulgaris ESP-31 for biodiesel synthesis using immobilized

Burkholderia lipase. Bioresour. Technol. 108, 119–127 (2012).

137. Lai, J.-Q., Hu, Z.-L., Wang, P.-W. & Yang, Z. Enzymatic production of microalgal

biodiesel in ionic liquid [BMIm][PF6]. Fuel 95, 329–333 (2012).

Page 115: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

100

138. Fjerbaek, L., Christensen, K. V. & Norddahl, B. A review of the current state of biodiesel

production using enzymatic transesterification. Biotechnol. Bioeng. 102, 1298–1315

(2009).

139. Smith, B., Greenwell, H. C. & Whiting, A. Catalytic upgrading of tri-glycerides and fatty

acids to transport biofuels. Energy Environ. Sci. 2, 262–271 (2009).

140. Tran, N. H. et al. Catalytic upgrading of biorefinery oil from micro-algae. Fuel 89, 265–

274 (2010).

141. Milne, T. A., Evans, R. J. & Nagle, N. Catalytic conversion of microalgae and vegetable

oils to premium gasoline, with shape-selective zeolites. Biomass 21, 219–232 (1990).

142. Yen, H.-W. et al. Microalgae-based biorefinery – From biofuels to natural products.

Bioresour. Technol. 135, 166–174 (2013).

143. Banerjee, A., Sharma, R., Chisti, Y. & Banerjee, U. C. Botryococcus braunii : A

Renewable Source of Hydrocarbons and Other Chemicals. Crit. Rev. Biotechnol. 22, 245–

279 (2002).

144. Odjadjare, E. C., Mutanda, T. & Olaniran, A. O. Potential biotechnological application of

microalgae: a critical review. Crit. Rev. Biotechnol. 37, 37–52 (2017).

145. Solimeno, A. et al. New mechanistic model to simulate microalgae growth. Algal Res. 12,

350–358 (2015).

146. Saidur, R., Abdelaziz, E. A., Demirbas, A., Hossain, M. S. & Mekhilef, S. A review on

biomass as a fuel for boilers. Renew. Sustain. Energy Rev. 15, 2262–2289 (2011).

147. Demirbaş, A. Biomass resource facilities and biomass conversion processing for fuels and

chemicals. Energy Convers. Manag. 42, 1357–1378 (2001).

148. Milledge, J. J. & Heaven, S. Methods of energy extraction from microalgal biomass: A

review. Rev. Environ. Sci. Biotechnol. 13, 301–320 (2014).

149. Kumar, G. et al. A review of thermochemical conversion of microalgal biomass for

biofuels: chemistry and processes. Green Chem. 19, 44–67 (2017).

150. Chen, W.-H., Lin, B.-J., Huang, M.-Y. & Chang, J.-S. Thermochemical conversion of

microalgal biomass into biofuels: A review. Bioresour. Technol. 184, 314–327 (2015).

151. Carpenter, D., Westover, T. L., Czernik, S. & Jablonski, W. Biomass feedstocks for

renewable fuel production: a review of the impacts of feedstock and pretreatment on the

yield and product distribution of fast pyrolysis bio-oils and vapors. Green Chem. 16, 384–

406 (2014).

152. Miao, X., Wu, Q. & Yang, C. Fast pyrolysis of microalgae to produce renewable fuels. J.

Anal. Appl. Pyrolysis 71, 855–863 (2004).

153. Soh, L. & Zimmerman, J. Biodiesel production: the potential of algal lipids extracted with

supercritical carbon dioxide. Green Chem. 13, 1422 (2011).

154. Mohsenpour, S. F. & Willoughby, N. Luminescent photobioreactor design for improved

algal growth and photosynthetic pigment production through spectral conversion of light.

Bioresour. Technol. 142, 147–153 (2013).

155. Wang, S., Jiang, X. M., Han, X. X. & Liu, J. G. Combustion Characteristics of Seaweed

Page 116: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

101

Biomass. 1. Combustion Characteristics of Enteromorpha clathrata and Sargassum natans.

Energy & Fuels 23, 5173–5178 (2009).

156. Cheng, X., Ooms, M. D. & Sinton, D. Biomass-to-biocrude on a chip via hydrothermal

liquefaction of algae. Lab Chip 16, 256–260 (2016).

157. Elliott, D. C. et al. Process development for hydrothermal liquefaction of algae feedstocks

in a continuous-flow reactor. Algal Res. 2, 445–454 (2013).

158. Lu, Y. & Savage, P. E. Supercritical water gasification of lipid-extracted hydrochar to

recover energy and nutrients. J. Supercrit. Fluids 99, 88–94 (2015).

159. Caputo, G. et al. Supercritical water gasification of microalgae and their constituents in a

continuous reactor. J. Supercrit. Fluids 118, 163–170 (2016).

160. Tiong, L., Komiyama, M., Uemura, Y. & Nguyen, T. T. Catalytic supercritical water

gasification of microalgae: Comparison of Chlorella vulgaris and Scenedesmus

quadricauda. J. Supercrit. Fluids 107, 408–413 (2016).

161. Laurens, L. M. L. et al. Acid-catalyzed algal biomass pretreatment for integrated lipid and

carbohydrate-based biofuels production. Green Chem. 17, 1145–1158 (2015).

162. Dong, T. et al. Combined algal processing: A novel integrated biorefinery process to

produce algal biofuels and bioproducts. Algal Res. 19, 316–323 (2016).

163. Ellis, J. T., Hengge, N. N., Sims, R. C. & Miller, C. D. Acetone, butanol, and ethanol

production from wastewater algae. Bioresour. Technol. 111, 491–495 (2012).

164. Chen, C. Y. et al. Microalgae-based carbohydrates for biofuel production. Biochem. Eng.

J. 78, 1–10 (2013).

165. Ho, S. H. et al. Bioethanol production using carbohydrate-rich microalgae biomass as

feedstock. Bioresour. Technol. 135, 191–198 (2013).

166. Miranda, J. R., Passarinho, P. C. & Gouveia, L. Pre-treatment optimization of

Scenedesmus obliquus microalga for bioethanol production. Bioresour. Technol. 104,

342–348 (2012).

167. Harun, R., Danquah, M. K. & Forde, G. M. Microalgal biomass as a fermentation

feedstock for bioethanol production. J. Chem. Technol. Biotechnol. 85, 199–203 (2010).

168. YEN, H. & BRUNE, D. Anaerobic co-digestion of algal sludge and waste paper to

produce methane. Bioresour. Technol. 98, 130–134 (2007).

169. Bohutskyi, P. et al. Effects of inoculum size, light intensity, and dose of anaerobic

digestion centrate on growth and productivity of Chlorella and Scenedesmus microalgae

and their poly-culture in primary and secondary wastewater. Algal Res. 19, 278–290

(2016).

170. Mahdy, A. et al. Ammonia tolerant inocula provide a good base for anaerobic digestion of

microalgae in third generation biogas process. Bioresour. Technol. 225, 272–278 (2017).

171. Henard, C. A. et al. Bioconversion of methane to lactate by an obligate methanotrophic

bacterium. Sci. Rep. 6, 21585 (2016).

172. Myung, J. et al. Long-term cultivation of a stable Methylocystis -dominated

methanotrophic enrichment enabling tailored production of poly(3-hydroxybutyrate-co-3-

Page 117: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

102

hydroxyvalerate). Bioresour. Technol. 198, 811–818 (2015).

173. Jin, J. et al. Characteristics of extracellular hydrocarbon-rich microalga Botryococcus

braunii for biofuels production: Recent advances and opportunities. Process Biochem. 51,

1866–1875 (2016).

174. Michinaka, Y. Extracellular secretion of free fatty acids by disruption of a fatty acyl-CoA

synthetase gene in Saccharomyces cerevisiae. J. Biosci. Bioeng. 95, 435–440 (2003).

175. Moheimani, N. R., Matsuura, H., Watanabe, M. M. & Borowitzka, M. A. Non-destructive

hydrocarbon extraction from Botryococcus braunii BOT-22 (race B). J. Appl. Phycol. 26,

1453–1463 (2014).

176. Bombelli, P., Müller, T., Herling, T. W., Howe, C. J. & Knowles, T. P. J. A High Power-

Density, Mediator-Free, Microfluidic Biophotovoltaic Device for Cyanobacterial Cells.

Adv. Energy Mater. 5, 1401299 (2015).

177. Rashid, N., Cui, Y.-F., Saif Ur Rehman, M. & Han, J.-I. Enhanced electricity generation

by using algae biomass and activated sludge in microbial fuel cell. Sci. Total Environ.

456–457, 91–94 (2013).

178. Powell, E. E., Mapiour, M. L., Evitts, R. W. & Hill, G. A. Growth kinetics of Chlorella

vulgaris and its use as a cathodic half cell. Bioresour. Technol. 100, 269–274 (2009).

179. Anderson, A. et al. Exploiting algal NADPH oxidase for biophotovoltaic energy. Plant

Biotechnol. J. 14, 22–28 (2016).

180. Kondaveeti, S., Choi, K. S., Kakarla, R. & Min, B. Microalgae Scenedesmus obliquus as

renewable biomass feedstock for electricity generation in microbial fuel cells (MFCs).

Front. Environ. Sci. Eng. 8, 784–791 (2014).

181. Cui, Y., Rashid, N., Hu, N., Rehman, M. S. U. & Han, J.-I. Electricity generation and

microalgae cultivation in microbial fuel cell using microalgae-enriched anode and bio-

cathode. Energy Convers. Manag. 79, 674–680 (2014).

182. Bombelli, P. et al. Quantitative analysis of the factors limiting solar power transduction by

Synechocystis sp. PCC 6803 in biological photovoltaic devices. Energy Environ. Sci. 4,

4690 (2011).

183. Samsonoff, N., Ooms, M. D. & Sinton, D. A photosynthetic-plasmonic-voltaic cell:

Excitation of photosynthetic bacteria and current collection through a plasmonic substrate.

Appl. Phys. Lett. 104, 43704 (2014).

184. Bradley, R. W., Bombelli, P., Rowden, S. J. L. & Howe, C. J. Biological photovoltaics:

intra- and extra-cellular electron transport by cyanobacteria: Figure 1. Biochem. Soc.

Trans. 40, 1302–1307 (2012).

185. Huang, H. & Yuan, X. Recent progress in the direct liquefaction of typical biomass. Prog.

Energy Combust. Sci. 49, 59–80 (2015).

186. Xu, C. & Lad, N. Production of heavy oils with high caloric values by direct liquefaction

of woody biomass in sub/near-critical water. Energy and Fuels 22, 635–642 (2008).

187. Hietala, D. C. & Savage, P. E. Reaction pathways and kinetics of cholesterol in high-

temperature water. Chem. Eng. J. 265, 129–137 (2015).

188. Biller, P., Sharma, B. K., Kunwar, B. & Ross, A. B. Hydroprocessing of bio-crude from

Page 118: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

103

continuous hydrothermal liquefaction of microalgae. Fuel 159, 197–205 (2015).

189. Miao, C., Chakraborty, M. & Chen, S. Impact of reaction conditions on the simultaneous

production of polysaccharides and bio-oil from heterotrophically grown Chlorella

sorokiniana by a unique sequential hydrothermal liquefaction process. Bioresour. Technol.

110, 617–627 (2012).

190. Heilmann, S. M. et al. Hydrothermal carbonization of microalgae. Biomass and Bioenergy

34, 875–882 (2010).

191. Mumme, J. et al. Hydrothermal carbonization of anaerobically digested maize silage.

Bioresour. Technol. 102, 9255–9260 (2011).

192. Patel, B., Guo, M., Izadpanah, A., Shah, N. & Hellgardt, K. A review on hydrothermal

pre-treatment technologies and environmental profiles of algal biomass processing.

Bioresour. Technol. 199, 288–299 (2016).

193. Kambo, H. S. & Dutta, A. Comparative evaluation of torrefaction and hydrothermal

carbonization of lignocellulosic biomass for the production of solid biofuel. Energy

Convers. Manag. 105, 746–755 (2015).

194. Praveenkumar, R., Lee, K., Lee, J. & Oh, Y.-K. Breaking dormancy: an energy-efficient

means of recovering astaxanthin from microalgae. Green Chem. 17, 1226–1234 (2015).

195. Desai, R. K., Streefland, M., Wijffels, R. H. & Eppink, M. H. M. Novel astaxanthin

extraction from Haematococcus pluvialis using cell permeabilising ionic liquids. Green

Chem. 18, 1261–1267 (2016).

196. Desai, R. K., Streefland, M., Wijffels, R. H. & Eppink, M. H. M. Novel astaxanthin

extraction from Haematococcus pluvialis using cell permeabilising ionic liquids. Green

Chem. 18, 1261–1267 (2016).

197. Cheng, X., Riordon, J., Nguyen, B., Ooms, M. D. & Sinton, D. Hydrothermal disruption

of algae cells for astaxanthin extraction. Green Chem. 19, 106–111 (2017).

198. Lee, D.-J., Chang, J.-S. & Lai, J.-Y. Microalgae–microbial fuel cell: A mini review.

Bioresour. Technol. 198, 891–895 (2015).

199. Sikarwar, V. S. et al. An overview of advances in biomass gasification. Energy Environ.

Sci. 9, 2939–2977 (2016).

200. Okabe, K. et al. Fischer–Tropsch Synthesis over Ru Catalysts by Using Syngas Derived

from Woody Biomass. Catal. Letters 128, 171–176 (2009).

201. Yang, Y., Xiang, H., Zhang, R., Zhong, B. & Li, Y. A highly active and stable Fe-Mn

catalyst for slurry Fischer–Tropsch synthesis. Catal. Today 106, 170–175 (2005).

202. Cao, C., Hu, J., Li, S., Wilcox, W. & Wang, Y. Intensified Fischer-Tropsch synthesis

process with microchannel catalytic reactors. Catal. Today 140, 149–156 (2009).

203. Hirano, A., Hon-Nami, K., Kunito, S., Hada, M. & Ogushi, Y. Temperature effect on

continuous gasification of microalgal biomass: theoretical yield of methanol production

and its energy balance. Catal. Today 45, 399–404 (1998).

204. Lu, Y., Levine, R. B. & Savage, P. E. Fatty Acids for Nutraceuticals and Biofuels from

Hydrothermal Carbonization of Microalgae. Ind. Eng. Chem. Res. 54, 4066–4071 (2015).

Page 119: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

104

205. Miodrag Belosevic, M. G. E. D. Z. S. J. R. B. Degradation of Alizarin Yellow R using UV

/ H 2 O 2 Advanced Oxidation Process. Environ. Sci. Technol. 33, 482–489 (2014).

206. Taher, H., Al-Zuhair, S., Al-Marzouqi, A. H., Haik, Y. & Farid, M. Effective extraction of

microalgae lipids from wet biomass for biodiesel production. Biomass and Bioenergy 66,

159–167 (2014).

207. Cheng, J., Huang, R., Li, T., Zhou, J. & Cen, K. Biodiesel from wet microalgae:

Extraction with hexane after the microwave-assisted transesterification of lipids.

Bioresour. Technol. 170, 69–75 (2014).

208. Dejoye Tanzi, C., Abert Vian, M. & Chemat, F. New procedure for extraction of algal

lipids from wet biomass: A green clean and scalable process. Bioresour. Technol. 134,

271–275 (2013).

209. Huo, Y.-X. et al. Conversion of proteins into biofuels by engineering nitrogen flux. Nat.

Biotechnol. 29, 346–351 (2011).

210. Li, C. & Fang, H. H. P. Fermentative Hydrogen Production From Wastewater and Solid

Wastes by Mixed Cultures. Crit. Rev. Environ. Sci. Technol. 37, 1–39 (2007).

211. Ward, A. J., Lewis, D. M. & Green, F. B. Anaerobic digestion of algae biomass: A

review. Algal Res. 5, 204–214 (2014).

212. Lakaniemi, A.-M., Tuovinen, O. H. & Puhakka, J. A. Anaerobic conversion of microalgal

biomass to sustainable energy carriers – A review. Bioresour. Technol. 135, 222–231

(2013).

213. Chidambarampadmavathy, K., Karthikeyan, O. P. & Heimann, K. Biopolymers made

from methane in bioreactors. Eng. Life Sci. 15, 689–699 (2015).

214. Zamalloa, C., Vulsteke, E., Albrecht, J. & Verstraete, W. The techno-economic potential

of renewable energy through the anaerobic digestion of microalgae. Bioresour. Technol.

102, 1149–1158 (2011).

215. Klemm, D., Heublein, B., Fink, H.-P. & Bohn, A. Cellulose: Fascinating Biopolymer and

Sustainable Raw Material. Angew. Chemie Int. Ed. 44, 3358–3393 (2005).

216. Heaven, S., Milledge, J. & Zhang, Y. Comments on ‘Anaerobic digestion of microalgae as

a necessary step to make microalgal biodiesel sustainable’. Biotechnol. Adv. 29, 164–167

(2011).

217. Levin, D. Biohydrogen production: prospects and limitations to practical application. Int.

J. Hydrogen Energy 29, 173–185 (2004).

218. Ghasemi, Y. et al. Microalgae biofuel potentials (Review). Appl. Biochem. Microbiol. 48,

126–144 (2012).

219. Benemann, J. R. Hydrogen production by microalgae. J. Appl. Phycol. 12, 291–300

(2000).

220. Daroch, M., Geng, S. & Wang, G. Recent advances in liquid biofuel production from algal

feedstocks. Appl. Energy 102, 1371–1381 (2013).

221. Radakovits, R., Jinkerson, R. E., Darzins, A. & Posewitz, M. C. Genetic Engineering of

Algae for Enhanced Biofuel Production. Eukaryot. Cell 9, 486–501 (2010).

Page 120: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

105

222. Wijffels, R. H., Kruse, O. & Hellingwerf, K. J. Potential of industrial biotechnology with

cyanobacteria and eukaryotic microalgae. Curr. Opin. Biotechnol. 24, 405–413 (2013).

223. Gaffron, H. & Rubin, J. FERMENTATIVE AND PHOTOCHEMICAL PRODUCTION

OF HYDROGEN IN ALGAE. J. Gen. Physiol. 26, 219–240 (1942).

224. Kumar, K. & Das, D. in Natural and Artificial Photosynthesis 173–215 (John Wiley &

Sons Inc., 2013). doi:10.1002/9781118659892.ch6

225. Bombelli, P. et al. Surface morphology and surface energy of anode materials influence

power outputs in a multi-channel mediatorless bio-photovoltaic (BPV) system. Phys.

Chem. Chem. Phys. 14, 12221 (2012).

226. Lea-Smith, D. J. et al. Thylakoid Terminal Oxidases Are Essential for the

Cyanobacterium Synechocystis sp. PCC 6803 to Survive Rapidly Changing Light

Intensities. PLANT Physiol. 162, 484–495 (2013).

227. Bradley, R. W., Bombelli, P., Lea-Smith, D. J. & Howe, C. J. Terminal oxidase mutants of

the cyanobacterium Synechocystis sp. PCC 6803 show increased electrogenic activity in

biological photo-voltaic systems. Phys. Chem. Chem. Phys. 15, 13611 (2013).

228. Bombelli, P. et al. Comparison of power output by rice (Oryza sativa) and an associated

weed (Echinochloa glabrescens) in vascular plant bio-photovoltaic (VP-BPV) systems.

Appl. Microbiol. Biotechnol. 97, 429–438 (2013).

229. McCormick, A. J. et al. Biophotovoltaics: oxygenic photosynthetic organisms in the world

of bioelectrochemical systems. Energy Environ. Sci. 8, 1092–1109 (2015).

230. Zhu, Y., Albrecht, K. O., Elliott, D. C., Hallen, R. T. & Jones, S. B. Development of

hydrothermal liquefaction and upgrading technologies for lipid-extracted algae conversion

to liquid fuels. Algal Res. 2, 455–464 (2013).

231. Ooms, M. D., Jeyaram, Y. & Sinton, D. Wavelength-selective plasmonics for enhanced

cultivation of microalgae. Appl. Phys. Lett. 106, 63902 (2015).

232. Jung, E. E. et al. Slab waveguide photobioreactors for microalgae based biofuel

production. Lab Chip 12, 3740 (2012).

233. Ahsan, S. S., Gumus, A., Jain, A., Angenent, L. T. & Erickson, D. Integrated hollow fiber

membranes for gas delivery into optical waveguide based photobioreactors. Bioresour.

Technol. 192, 845–849 (2015).

234. Jain, A. et al. Optimal Intensity and Biomass Density for Biofuel Production in a Thin-

Light-Path Photobioreactor. Environ. Sci. Technol. 49, 6327–6334 (2015).

235. Graham, P. J., Riordon, J. & Sinton, D. Microalgae on display: a microfluidic pixel-based

irradiance assay for photosynthetic growth. Lab Chip 15, 3116–3124 (2015).

236. Razzaq, T. & Kappe, C. O. Continuous Flow Organic Synthesis under High-

Temperature/Pressure Conditions. Chem. - An Asian J. 5, NA-NA (2010).

237. Boden, R., Ogden, S. & Hjort, K. Microdispenser With Continuous Flow and Selectable

Target Volume for Microfluidic High-Pressure Applications. J. Microelectromechanical

Syst. 23, 452–458 (2014).

238. Elsey, D., Jameson, D., Raleigh, B. & Cooney, M. J. Fluorescent measurement of

microalgal neutral lipids. J. Microbiol. Methods 68, 639–642 (2007).

Page 121: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

106

239. Ryder, A. G. in Reviews in Fluorescence 2005 169–198 (Springer US, 2005).

doi:10.1007/0-387-23690-2_8

240. Xu, D. & Savage, P. E. Characterization of biocrudes recovered with and without solvent

after hydrothermal liquefaction of algae. Algal Res. 6, 1–7 (2014).

241. Li, H. et al. Conversion efficiency and oil quality of low-lipid high-protein and high-lipid

low-protein microalgae via hydrothermal liquefaction. Bioresour. Technol. 154, 322–329

(2014).

242. Lorenz, R. T. & Cysewski, G. R. Commercial potential for Haematococcus microalgae as

a natural source of astaxanthin. Trends Biotechnol. 18, 160–167 (2000).

243. Margalith, P. Z. Production of ketocarotenoids by microalgae. Appl. Microbiol.

Biotechnol. 51, 431–438 (1999).

244. Paudel, A., Jessop, M. J., Stubbins, S. H., Champagne, P. & Jessop, P. G. Extraction of

lipids from microalgae using CO2-expanded methanol and liquid CO2. Bioresour.

Technol. 184, 286–290 (2015).

245. Panis, G. & Carreon, J. R. Commercial astaxanthin production derived by green alga

Haematococcus pluvialis: A microalgae process model and a techno-economic assessment

all through production line. Algal Res. 18, 175–190 (2016).

246. Günerken, E. et al. Cell disruption for microalgae biorefineries. Biotechnol. Adv. 33, 243–

260 (2015).

247. Teo, C. C., Tan, S. N., Yong, J. W. H., Hew, C. S. & Ong, E. S. Pressurized hot water

extraction (PHWE). J. Chromatogr. A 1217, 2484–2494 (2010).

248. Plaza, M. & Turner, C. Pressurized hot water extraction of bioactives. TrAC Trends Anal.

Chem. 71, 39–54 (2015).

249. Nazari, L., Yuan, Z., Souzanchi, S., Ray, M. B. & Xu, C. (Charles). Hydrothermal

liquefaction of woody biomass in hot-compressed water: Catalyst screening and

comprehensive characterization of bio-crude oils. Fuel 162, 74–83 (2015).

250. Zhang, C., Tang, X., Sheng, L. & Yang, X. Enhancing the performance of Co-

hydrothermal liquefaction for mixed algae strains by the Maillard reaction. Green Chem.

18, 2542–2553 (2016).

251. Déniel, M., Haarlemmer, G., Roubaud, A., Weiss-Hortala, E. & Fages, J. Bio-oil

Production from Food Processing Residues: Improving the Bio-oil Yield and Quality by

Aqueous Phase Recycle in Hydrothermal Liquefaction of Blackcurrant ( Ribes nigrum L.)

Pomace. Energy & Fuels 30, 4895–4904 (2016).

252. Jin, F. & Enomoto, H. Rapid and highly selective conversion of biomass into value-added

products in hydrothermal conditions: chemistry of acid/base-catalysed and oxidation

reactions. Energy Environ. Sci. 4, 382–397 (2011).

253. Anastas, P. & Eghbali, N. Green Chemistry: Principles and Practice. Chem. Soc. Rev. 39,

301–312 (2010).

254. Yu, X., Yang, J., Lu, H., Tu, S.-T. & Yan, J. Energy-efficient extraction of fuel from

Chlorella vulgaris by ionic liquid combined with CO2 capture. Appl. Energy 160, 648–

655 (2015).

Page 122: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

107

255. Wang, L., Yang, B., Yan, B. & Yao, X. Supercritical fluid extraction of astaxanthin from

Haematococcus pluvialis and its antioxidant potential in sunflower oil. Innov. Food Sci.

Emerg. Technol. 13, 120–127 (2012).

256. Fábregas, J., Domínguez, A., Regueiro, M., Maseda, A. & Otero, A. Optimization of

culture medium for the continuous cultivation of the microalga Haematococcus pluvialis.

Appl. Microbiol. Biotechnol. 53, 530–535 (2000).

257. Bobleter, O. Hydrothermal degradation of polymers derived from plants. Prog. Polym.

Sci. 19, 797–841 (1994).

258. Sarada, R., Vidhyavathi, R., Usha, D. & Ravishankar, G. a. An Efficient Method for

Extraction of Astaxanthin from Green Alga Haematococcus pluvialis. J. Agric. Food

Chem. 54, 7585–7588 (2006).

259. Boussiba, S., Bing, W., Yuan, J.-P., Zarka, A. & Chen, F. Changes in pigments profile in

the green alga Haeamtococcus pluvialis exposed to environmental stresses. Biotechnol.

Lett. 21, 601–604 (1999).

260. Ranga, R., Sarada, A. R., Baskaran, V. & Ravishankar, G. A. Identification of carotenoids

from green alga Haematococcus pluvialis by HPLC and LC-MS (APCI) and their

antioxidant properties. J. Microbiol. Biotechnol. 19, 1333–41 (2009).

261. Raposo, M. F. J., Morais, A. M. M. B. & Morais, R. M. S. C. Effects of spray-drying and

storage on astaxanthin content of Haematococcus pluvialis biomass. World J. Microbiol.

Biotechnol. 28, 1253–1257 (2012).

262. Kaczor, A. & Baranska, M. Structural Changes of Carotenoid Astaxanthin in a Single

Algal Cell Monitored in Situ by Raman Spectroscopy. Anal. Chem. 83, 7763–7770

(2011).

263. Chew, K. W. et al. Microalgae biorefinery: High value products perspectives. Bioresour.

Technol. 229, 53–62 (2017).

264. Leu, S. & Boussiba, S. Advances in the Production of High-Value Products by

Microalgae. Ind. Biotechnol. 10, 169–183 (2014).

265. Wang, H.-M. D., Chen, C.-C., Huynh, P. & Chang, J.-S. Exploring the potential of using

algae in cosmetics. Bioresour. Technol. 184, 355–362 (2015).

266. Venkata Subhash, G. et al. Carbon streaming in microalgae: extraction and analysis

methods for high value compounds. Bioresour. Technol. 244, 1304–1316 (2017).

267. Lee, S. Y., Cho, J. M., Chang, Y. K. & Oh, Y.-K. Cell disruption and lipid extraction for

microalgal biorefineries: A review. Bioresour. Technol. 244, 1317–1328 (2017).

268. Su, Y. et al. Progress of microalgae biofuel’s commercialization. Renew. Sustain. Energy

Rev. 74, 402–411 (2017).

269. Díaz-Reinoso, B., Moure, A., Domínguez, H. & Parajó, J. C. Supercritical CO 2

Extraction and Purification of Compounds with Antioxidant Activity. J. Agric. Food

Chem. 54, 2441–2469 (2006).

270. Reyes, F. A., Sielfeld, C. S. & del Valle, J. M. Effect of high-pressure compaction on

supercritical CO2 extraction of astaxanthin from Haematococcus pluvialis. J. Food Eng.

189, 123–134 (2016).

Page 123: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

108

271. Perrut, M. Supercritical Fluid Applications: Industrial Developments and Economic

Issues. Ind. Eng. Chem. Res. 39, 4531–4535 (2000).

272. Todd, R. & Baroutian, S. A techno-economic comparison of subcritical water,

supercritical CO 2 and organic solvent extraction of bioactives from grape marc. J. Clean.

Prod. 158, 349–358 (2017).

273. de la Fuente, J. C., Oyarzún, B., Quezada, N. & del Valle, J. M. Solubility of carotenoid

pigments (lycopene and astaxanthin) in supercritical carbon dioxide. Fluid Phase Equilib.

247, 90–95 (2006).

274. Xu, Y., Riordon, J., Cheng, X., Bao, B. & Sinton, D. The full pressure-temperature phase

envelope of a mixture in 1000 microfluidic chambers. Angew. Chemie Int. Ed. (2017).

doi:10.1002/anie.201708238

275. Schneider, G. M. Physicochemical Principles of Extraction with Supercritical Gases.

Angew. Chemie Int. Ed. English 17, 716–727 (1978).

276. Ke, J., Oag, R. M., King, P. J., George, M. W. & Poliakoff, M. Sensing the Critical Point

of High-Pressure Mixtures. Angew. Chemie Int. Ed. 43, 5192–5195 (2004).

277. Bröll, D. et al. Chemistry in Supercritical Water. Angew. Chemie Int. Ed. 38, 2998–3014

(1999).

278. Wetenhall, B., Race, J. M. & Downie, M. J. The Effect of CO2 Purity on the Development

of Pipeline Networks for Carbon Capture and Storage Schemes. Int. J. Greenh. Gas

Control 30, 197–211 (2014).

279. Shah, M. S., Tsapatsis, M. & Siepmann, J. I. Identifying Optimal Zeolitic Sorbents for

Sweetening of Highly Sour Natural Gas. Angew. Chemie Int. Ed. 55, 5938–5942 (2016).

280. Dudin, S. M., Zemenkov, Y. D. & Shabarov, A. B. Modeling the Phase Composition of

Gas Condensate in Pipelines. IOP Conf. Ser. Mater. Sci. Eng. 154, 12010 (2016).

281. Di Primio, R., Dieckmann, V. & Mills, N. PVT and phase behaviour analysis in petroleum

exploration. Org. Geochem. 29, 207–222 (1998).

282. Song, W., Fadaei, H. & Sinton, D. Determination of Dew Point Conditions for CO2 with

Impurities Using Microfluidics. Environ. Sci. Technol. 48, 3567–3574 (2014).

283. Sullivan, M. T. & Angelescu, D. E. Microfluidic Bubble Point Measurement Using

Thermal Nucleation. Energy & Fuels 30, 2655–2661 (2016).

284. Mostowfi, F., Molla, S. & Tabeling, P. Determining phase diagrams of gas–liquid systems

using a microfluidic PVT. Lab Chip 12, 4381 (2012).

285. Bao, B., Riordon, J., Mostowfi, F. & Sinton, D. Microfluidic and nanofluidic phase

behaviour characterization for industrial CO 2 , oil and gas. Lab Chip 17, 2740–2759

(2017).

286. Laval, P., Lisai, N., Salmon, J.-B. & Joanicot, M. A microfluidic device based on droplet

storage for screening solubility diagrams. Lab Chip 7, 829 (2007).

287. Shim, J. et al. Control and Measurement of the Phase Behavior of Aqueous Solutions

Using Microfluidics. J. Am. Chem. Soc. 129, 8825–8835 (2007).

288. Selimović, Š., Gobeaux, F. & Fraden, S. Mapping and manipulating temperature–

Page 124: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

109

concentration phase diagrams using microfluidics. Lab Chip 10, 1696–1699 (2010).

289. Hansen, C. L., Sommer, M. O. A. & Quake, S. R. Systematic investigation of protein

phase behavior with a microfluidic formulator. Proc. Natl. Acad. Sci. U. S. A. 101, 14431–

14436 (2004).

290. Bao, B., Riordon, J., Xu, Y., Li, H. & Sinton, D. Direct Measurement of the Fluid Phase

Diagram. Anal. Chem. 88, 6986–6989 (2016).

291. Karimi, A., Hughes, T. J., Richter, M. & May, E. F. Density Measurements of Methane +

Propane Mixtures at Temperatures between (256 and 422) K and Pressures from

(24 to 35) MPa. J. Chem. Eng. Data 61, 2782–2790 (2016).

292. Healy, D. et al. Methane/propane mixture oxidation at high pressures and at high,

intermediate and low temperatures. Combust. Flame 155, 451–461 (2008).

293. Faramawy, S., Zaki, T. & Sakr, A. A.-E. Natural gas origin, composition, and processing:

A review. J. Nat. Gas Sci. Eng. 34, 34–54 (2016).

294. Zheng, D.-Q., Ma, W.-D., Wei, R. & Guo, T.-M. Solubility study of methane, carbon

dioxide and nitrogen in ethylene glycol at elevated temperatures and pressures. Fluid

Phase Equilib. 155, 277–286 (1999).

295. Abdi, M. A., Hussain, A., Hawboldt, K. & Beronich, E. Experimental Study of Solubility

of Natural Gas Components in Aqueous Solutions of Ethylene Glycol at Low-

Temperature and High-Pressure Conditions. J. Chem. Eng. Data 52, 1741–1746 (2007).

296. Wang, L.-K., Chen, G.-J., Han, G.-H., Guo, X.-Q. & Guo, T.-M. Experimental study on

the solubility of natural gas components in water with or without hydrate inhibitor. Fluid

Phase Equilib. 207, 143–154 (2003).

297. Jou, F.-Y., Otto, F. D. & Mather, A. E. Solubility of methane in glycols at elevated

pressures. Can. J. Chem. Eng. 72, 130–133 (1994).

298. Soave, G., Gamba, S. & Pellegrini, L. A. SRK equation of state: Predicting binary

interaction parameters of hydrocarbons and related compounds. Fluid Phase Equilib. 299,

285–293 (2010).

299. Coquelet, C., Chareton, A., Valtz, A., Baba-Ahmed, A. & Richon, D. Vapor−Liquid

Equilibrium Data for the Azeotropic Difluoromethane + Propane System at Temperatures

from 294.83 to 343.26 K and Pressures up to 5.4 MPa. J. Chem. Eng. Data 48, 317–323

(2003).

Page 125: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

110

Appendices

The following appendix is a manuscript for which the applicant was co-author and has been

published in Angewandte Chemie, reprinted274 with permission from John Wiley and Sons. While

it is not to be considered part of the core contributions of this thesis, they are nevertheless

worthwhile considering since it provides additional breadth to the application of high temperature

and pressure microfluidics.

A1. The full thermodynamic phase envelope of a mixture in 1000

microfluidic chambers

Authors: Yi Xu, Jason Riordon, Xiang Cheng, Bo Bao and David Sinton*

Abstract: Knowing the thermodynamic state of complex mixtures – liquid, gas,

supercritical or two-phase – is essential to industrial chemical processes. Traditionally,

phase diagrams are compiled piecemeal from individual measurements in a pressure-

volume-temperature cell performed in series, where each point is subject to a long fluid

equilibrium time. Herein, 1,000 microfluidic chambers, each isolated by a liquid piston

and set to a different pressure and temperature combination, provide the complete

pressure-temperature phase diagram of a hydrocarbon mixture at once, including the

thermodynamic phase envelope. Measurements closely match modelled values, with a

standard deviation of 0.13 MPa between measurement and model for the dew and bubble

point lines, and a difference of 0.04 MPa and 0.25 ºC between measurement and model

for the critical point.

Long a fixture within chemistry textbooks, phase diagrams are crucial to understanding chemical

processes – phase diagrams identify how complex fluid mixtures behave under a wide range of

pressures and temperatures.275 Detailed knowledge of a multi-component fluid’s two-phase

envelope and associated key parameters – dew and bubble point lines, liquid volume lines and

critical point – is critical to understanding a variety of chemical processes,276 and is often difficult

to obtain using equation-of-state models.277 For many fluid mixtures, even small changes in

composition can have a profound impact on thermodynamic properties.278 Measuring fluid phase

Page 126: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

111

properties is routine in the oil and gas industry, where complex hydrocarbon mixtures279,280 – raw

natural gas, gas condensates and oil – are analyzed during recovery, treatment, transportation and

consumption processes. Measurement of the phase diagram is typically performed using traditional

pressure-volume-temperature (PVT) cells.281 Compiling even a sparse phase diagram in this

manner is a costly and time-consuming process, often requiring hours with costly large-scale

experimentation to determine the phase state at a single point on the pressure-temperature plane.

Recently, specialized microfluidic approaches have been applied to measure isolated phase

properties, including the dew point,22,282 bubble point,22,283,284 and liquid-to-vapor ratios.284 These

methods each use direct visual observation to determine the formation of droplets or bubbles

within a confined planar surface or microfluidic channel and benefit from short equilibrium times,

reduced sample volume requirements and precise control over pressure and temperature offered

by microfluidic systems.16,285 Approaches to date, however, involve measuring single pressure-

temperature conditions in series, and fail to exploit parallelization opportunities as recently

demonstrated in other phase-mapping microfluidics applications, for example in concentration-

concentration or temperature-concentration phase mapping of salts,286 polymers287 or protein

crystallization.287–289 Bao et al. recently demonstrated phase diagram mapping of pure CO2 by

using an array of microwells subject to different pressure-temperature conditions.290 Visualizing

the two-phase envelope of mixtures, however, was not possible due to the interconnected nature

of the device. Specifically, fluid communication between microwells precluded the resolution of

the mixture phase envelope, limiting application to pure substances. Mixtures, however, are the

norm in chemical applications and pure substances are the exception.

Herein, we report full mapping of the phase diagram of a fluid mixture with an array of 1,000

microfluidic chambers, each isolated and pressurized by liquid pistons. The liquid piston fluid

compartmentalizes the test fluid, fully isolating and pressurizing each microfluidic chamber and

enabling accurate measurement of vapor-liquid ratios therein. A temperature gradient is applied

perpendicular to the pressure gradient by using external temperature controls. Our method enables

rapid one-step measurement of a fluid mixture’s phase diagram including dew and bubble point

lines, liquid volume lines and critical point.

Page 127: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

112

Figure 1. a) Schematic of the rapid pressure-temperature phase mapping device for mixtures. Insets show enlarged

regions of the device, Liquid piston operation is demonstrated in (b) and (c), respectively. d) Corresponding phase

diagram.

Our approach is shown schematically in Figure 1. Trapped pockets of test fluid are isolated by the

piston fluid within long vertical interdigitated microfluidic chambers (2 mm long × 100 μm wide

× 15 μm tall) in a silicon-glass chip by sequentially loading first the mixture and then the piston

fluid. These microfluidic chambers are each connected at their base to dead-end horizontal

channels which are positioned parallel to each other across the device, and connect to the flow

channel at their leftmost point (see Figure S1 in the Supporting Information for full details). During

regular operation, the piston fluid is continually flowing through the device – entering at the inlet,

flowing vertically along the flow channel, and exiting at the outlet; fluid throughout the remainder

of the channel network is stagnant (Figure 1a). The pressure in each horizontal channel is fixed

based on the location of the connection point to the flow channel, and pressure measurements at

both ends. At low pressure, the liquid pistons are near the entrance, whereas at high pressure, the

liquid pistons extend deep into the chamber (Figure 1b and c). In addition, an orthogonal linear

Page 128: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

113

temperature gradient is achieved in the horizontal direction via temperature controlled copper

blocks in direct contact with the silicon. The high thermal conductivity of silicon ensures a linear

temperature gradient and precise control over local temperature conditions within the gradient-

orthogonal microfluidic chambers (see Figure S2 in the Supporting Information for temperature

calibration). The result is a pair of orthogonal, linear pressure and temperature gradients across the

device, which enable mapping of the pressure-temperature phase diagram of a fluid mixture (see

Table S1 in the Supporting Information). The temperature variation across a given chamber is

negligible, less than 1% of the full temperature range and the temperature range is tunable to the

region of interest.

The ideal piston fluid should (i) be insoluble with the test mixture (i.e. use a polar piston fluid for

a non-polar test mixture), (ii) have strong wall-wetting properties, (iii) remain liquid at all test

pressures and temperatures, (iv) have a low vapor pressure, (v) have a low viscosity and (vi)

provide sharp visual contrast with the test fluid. Here, to contain our test fluid – an 80.0% propane

+ 20.0 % methane mixture (mole fraction, Praxair Canada, Inc.) representative of thermogenic

natural gas.291–293 – we used ethylene glycol (99.99% purity, Shell Chemical, LP.), which readily

satisfies these criteria and has been deployed successfully in related applications in industry.294,295

Notably, the solubility of natural gas within ethylene glycol is low, and has been well characterized

over a wide range of temperature and pressures.294–297 . For other applications involving, for

instance, aqueous test fluids, mineral oil could be a suitable piston fluid.

The structure of our device ensures a linear gradient between inlet and outlet pressures and

temperature. The user controls the range, effectively setting the scale of both axes on the phase

diagram of interest, as well as the corresponding resolution. We first demonstrate application over

a wide range of pressures and temperatures, P = 0.89 ± 0.013 MPa – 6.75 ± 0.013 MPa and T =

20.9 ± 0.23 ºC – 89.8 ± 0.23 ºC, respectively (error bars are representative of instrument

uncertainty and fluctuations observed over time). As shown in Figure 2a, this broad, industrially

relevant pressure-temperature range was chosen to encompass a broad region of the two-phase

envelope of the mixture. A complete set of microscope images of each microfluidic chamber were

collected in sequence immediately after sample loading over the course of 3 min. These images

were processed similarly (contrast enhanced, cropped, and realigned) and for cases where two

phases were present (two-phase envelope), images were overlaid with a red or blue filter based on

phase state as determined through semi-automated analysis (see Figure S3 in Supporting

Page 129: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

114

Information image processing example and protocol). All images were positioned within a grid at

coordinates corresponding to their pressure-temperature location to form the mosaic in Figure 2a.

Enlarged regions of interest (i-vii) are shown to highlight typical test fluid within key areas of the

phase diagram. For all pressure and temperature cases, the test mixture was successfully isolated

by ethylene glycol within individual chambers. The piston extension length, and the volume of the

test fluid, is dependent on applied pressure as well as the state of the test fluid. The two-phase

envelope, highlighted here with a red/blue filter overlay, is rich in thermodynamic information.

Within the two-phase envelope, vapor bubbles (red) increase in size from left to right (low

temperature to high temperature). At low pressures, a dark hue can often be seen within the vapor

regions, indicative of the presence of a liquid film along the chamber wall. Figure 2b shows a color

map of the vapor volume percentage within the two-phase region, as obtained through image

analysis. Such analysis was performed by measuring the liquid region area, and dividing by the

total confined area. Given that the microfluidic chambers are shallow, the area ratio is taken as a

measure of the volume ratio – the role of edge “rounding” and film formation is minimal. Whereas

identification of the presence of both liquid and vapor phases is straightforward at the center of the

envelope, measurements are more challenging near the bubble point line (where bubbles are small)

and dew point line (where the liquid region is confined to a thin film). Whereas quantifying the

precise liquid content (vol. %) at these locations was not feasible, optical microscopy was fully

capable of discerning the location of the dew/bubble point boundary, which is of interest here.

Page 130: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

115

Figure 2. Wide-range pressure and temperature mapping of the phase diagram of an 80.0% propane + 20.0% methane

mixture. a) Image mosaic of an array of microfluidic chambers, each light-corrected, contrast-enhanced, cropped,

realigned and overlaid with a red or blue filter based on phase. Enlarged areas of interest are also shown. The test fluid

nearest the critical point is overlaid with a yellow filter. b) Liquid content within the two-phase envelope. c) Measured

phase envelope boundary positions, and points intersecting 25, 50 and 75 % liquid volume lines, as well as curve fits

and SRK model.

Page 131: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

116

Figure 2c shows how the liquid fill percentage data can be further analyzed to yield the

thermodynamic two-phase envelope with liquid volume lines. The red points here are plotted along

the boundary of the envelope in Figure 2b, at a temperature point half-way to the next microfluidic

chamber along the same pressure row. The error bars represent thus not only the experimental

uncertainty based on pressure and temperature measurements, but also the temperature difference

between two adjacent microfluidic chambers – our resolution limit. Determination of liquid

volume lines, however, is not bound by this resolution limit, as all liquid content data points along

a single pressure are fit to a curve, and the precise location of 25 %, 50 % and 75 % liquid volume

line intersects determined. The error bars on these liquid volume line points (green, magenta, cyan)

are based on the standard deviation of these fits, as wells as the experimental uncertainty based on

pressure and temperature measurements, as above. These positions were then fit to low-degree

polynomials to obtain dew point, bubble point and liquid volume lines. To determine the critical

point, the 50 % liquid-volume points nearing critical were fit to a first order polynomial function.

The intersection point between this line (magenta) and the phase envelope boundary (red)

corresponds to the measured critical point. The 25 % and 75 % liquid volume lines were fitted to

a low degree polynomial with a forced intersect to the critical point (further details in Table S2 in

Supporting Information).

To validate the measured phase diagram, we used a Soave-Redlich-Kwong equation of state (SRK)

model, widely used in industry to analyze the phase state of hydrocarbon mixtures,298,299 which

yielded the dashed blue curve on Figure 2c. Overall, a standard deviation of 0.13 MPa was obtained

between measurement and model for the dew and bubble point lines. The critical point was 6.07

MPa and 82.1 ºC, which is near the modeled critical point of 5.86 MPa and 82.7 ºC, respectively.

As is typical for instrumentation, our critical point resolution depends on the full scale. In the

Figure 2 example the full scale is very large (7 MPa and 90 ºC).

To demonstrate the device’s ability to zoom-in to region of interest within the phase diagram,

improve resolution, and demonstrate reproducibility, a series three consecutive mapping

experiments were performed around the critical point. Similar pressure and temperature ranges

were applied in each run, and collectively span the narrowed ranges of 5.63 ± 0.013 MPa – 6.52 ±

0.013 MPa and 68.9 ± 0.23 ºC - 88.2 ± 0.23 ºC. Figure 3a shows a mosaic of microfluidic chambers

for one of the three runs, illustrating fluid behavior near the critical point. In contrast to the Figure

2a mosaic, the zoom-in better resolves the fluid behavior near the critical point. The pressure

Page 132: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

117

difference between rows in Figure 2 is an order of magnitude smaller than in Figure 3. The phase

trends are immediately visible, such as the bubble size increasing as temperature is increased left

to right.

Figure 3. a) Narrow-range pressure and temperature mapping of the phase diagram of an 80.0 % methane + 20.0 %

propane mixture at the critical point. a) Image mosaic of an array of microfluidic chambers each light-corrected,

contrast-enhanced, cropped, realigned and overlaid with a red or blue filter based on phase state. Yellow highlights

the supercritical chamber nearest the critical point. b) Liquid content of microfluidic chambers within the two-phase

envelope. c) Measured phase envelope boundary positions, and points intersecting 25 %, 50 % and 75 % liquid volume

lines, as well as corresponding curve fits for a series of three identical experiments, each performed with the same

device. SRK model is also shown.

Figure 3b shows the measured phase envelope boundary locations and the fit envelope, 25 %, 50 %

and 75 % liquid volume line points and liquid volume lines, as well as curve fits for all three

consecutive mapping experiments combined. The experimental envelope and critical point are

compared to the SRK model. With the heightened resolution provided by the zoom-in, the

Page 133: In-situ observation and quantification of microalgae ......The direct observation of these complex reaction processes was made possible for the first time ... Sunlight-to-biomass conversion

118

measured critical point of 5.90 MPa and 81.8 ºC is now offset by 0.04 MPa and 0.25 ºC from the

model prediction. All three runs led to liquid volume and phase envelope boundary points that

were closely matched, which further indicates the reproducibility of the device.

Overall, the phase diagram mapping device provides flexibility, with the option to map the mixture

phase over a wide range of temperatures and pressures at once, or to zoom-in to a narrower region

of interest with high resolution. It is important to note that while the thermodynamic phase data

measured are time-independent, the data should be collected quickly to limit any potential for

dissolution of a mixture component within the piston fluid (see Figure S4 in Supporting

Information). Potential for solubility is greatest at high pressures. For the highest pressures, the

bubble point line can deviate from the initial measurement if left over time. In practice, we first

established a stable temperature gradient and uniform filling of the mixture in gas phase throughout.

Then we pressurized with the piston fluid, and immediately scanned the chip. The pressure and

phase data stabilized rapidly, within a few seconds, and the resulting bubble point line agreed well

with theory (Figure 2), and showed no more deviation than that the dew point line (unaffected by

solubility).

In conclusion, full thermodynamic phase envelope mapping of a fluid mixture has been

demonstrated by precisely varying temperature and pressure over 1,000 liquid piston-isolated

microfluidic chambers simultaneously. The approach is tunable, offering both wide and narrow

condition ranges while resolving the full suite of phase envelope data: dew and bubble point lines,

liquid volume lines and the critical point. Measurements closely matched expected results based

on an established equation of state model, with a standard deviation of 0.13 MPa between

measurement and model for the dew and bubble point lines, and a difference of 0.04 MPa and 0.25

ºC between measurement and model for the critical point. In stark contrast to the established state

of the art (serial point-by-point measurement) this method provides full characterization of the

phase envelope of a fluid mixture over the pressure-temperature plane of interest, informing a

broad range of chemical applications where fluid mixtures are the norm.