+ All Categories
Home > Documents > The Monoraphidium - Northern Illinois University

The Monoraphidium - Northern Illinois University

Date post: 03-Feb-2022
Category:
Upload: others
View: 4 times
Download: 0 times
Share this document with a friend
95
ABSTRACT PROPERTIES OF THE GREEN MICROALGA MONORAPHIDIUM SP. DEK19 FOR PHYCOREMEDIATION OF WASTEWATER AND BIOFUEL PRODUCTION Nicholas James Kirchner, MS Department of Biological Sciences Northern Illinois University, 2016 Gabriel Holbrook, Director The focus of this thesis was to investigate the growth of Monoraphidium sp. Dek19 in post-primary filtration and final wastewater effluent for the purposes of phycoremediation and as a potential biofuel feedstock. The algae were inoculated in wastewater effluent collected from the local wastewater treatment facility in DeKalb, Illinois and grown in 1L flasks at 10°C and 22°C. It was determined that initial population density (IPD) played an important role in the successful growth of a culture. Cultures started at an E680 greater than 0.100 could grow, but at a suboptimal rate. However, it was found that when cultures were started at a higher E680 value of ~0.60, cultures exhibited significantly shorter lag phases and therefore less time to the onset of stationary phase. Monoraphidium sp. Dek19 was shown to compete with other commonly occurring green microalgae when grown in a consortium of species. When cultures were inoculated with 20% Monoraphidium sp. Dek19 and an 80% mixture of Chlorella sp. and Ulothrix sp. in a culture grown at 10°C, Monoraphidium sp. Dek19 was able to maintain its population composition and made up ~70% of the total algal biomass. Sucrose density gradients were developed as an efficient method to separate a desired microalga from a consortium of species. Photosynthetic oxygen evolution rates of Monoraphidium sp. Dek19 were observed at 10°C and an ambient lab temperature (22-25°C) in both post-primary filtration and final wastewater effluent. Equivalent rates of oxygen evolution were observed in both media.
Transcript
Page 1: The Monoraphidium - Northern Illinois University

ABSTRACT

PROPERTIES OF THE GREEN MICROALGA MONORAPHIDIUM SP. DEK19 FOR

PHYCOREMEDIATION OF WASTEWATER AND BIOFUEL PRODUCTION

Nicholas James Kirchner, MS

Department of Biological Sciences

Northern Illinois University, 2016

Gabriel Holbrook, Director

The focus of this thesis was to investigate the growth of Monoraphidium sp. Dek19 in

post-primary filtration and final wastewater effluent for the purposes of phycoremediation and

as a potential biofuel feedstock. The algae were inoculated in wastewater effluent collected from

the local wastewater treatment facility in DeKalb, Illinois and grown in 1L flasks at 10°C and

22°C. It was determined that initial population density (IPD) played an important role in the

successful growth of a culture. Cultures started at an E680 greater than 0.100 could grow, but at

a suboptimal rate. However, it was found that when cultures were started at a higher E680 value

of ~0.60, cultures exhibited significantly shorter lag phases and therefore less time to the onset of

stationary phase. Monoraphidium sp. Dek19 was shown to compete with other commonly

occurring green microalgae when grown in a consortium of species. When cultures were

inoculated with 20% Monoraphidium sp. Dek19 and an 80% mixture of Chlorella sp. and

Ulothrix sp. in a culture grown at 10°C, Monoraphidium sp. Dek19 was able to maintain its

population composition and made up ~70% of the total algal biomass. Sucrose density gradients

were developed as an efficient method to separate a desired microalga from a consortium of

species. Photosynthetic oxygen evolution rates of Monoraphidium sp. Dek19 were observed at

10°C and an ambient lab temperature (22-25°C) in both post-primary filtration and final

wastewater effluent. Equivalent rates of oxygen evolution were observed in both media.

Page 2: The Monoraphidium - Northern Illinois University

Photosynthetic rates were higher at ambient lab temperatures than at 10°C. Monoraphidium sp.

Dek19 exhibited diminished rates of photosynthesis when tested at a higher or lower temperature

than its growth condition, indicating a possible perturbation of photosynthesis when forced to

quickly acclimate to a new temperature.

Oxygen evolved during algae growth may supplement the activated sludge process

during water treatment. A rise in the alga’s chlorophyll a:b ratios may indicate an adaptation to

cooler conditions. The algae were able to remediate polluting concentrations of nitrogen and

phosphorous to minimal levels in both media. Monoraphidium sp. Dek19 cultures were able to

rapidly accumulate lipids at the onset of stationary growth phase. This was determined through

the use of Nile red dye and images taken with a confocal microscope. It was concluded that

Monoraphidium sp. Dek19 could be used for phycoremediation of wastewater effluent either

from primary filtration or from the final settling tanks prior to effluent discharge into local

waterways. Monoraphidium sp. Dek19 could also be a potential feedstock for biofuel production

as the cells can grow within the cooler range of temperatures observed at wastewater treatment

facilities in the Midwest. This alga has lipids that can be successfully converted to biofuel.

Page 3: The Monoraphidium - Northern Illinois University

NORTHERN ILLINOIS UNIVERSITY

DEKALB, ILLINOIS

MAY 2016

PROPERTIES OF THE GREEN MICROALGA MONORAPHIDIUM SP.

DEK19 FOR PHYCOREMEDIATION OF WASTEWATER AND BIOFUEL

PRODUCTION

BY

NICHOLAS JAMES KIRCHNER

©2015 Nicholas James Kirchner

A THESIS SUBMITTED TO THE GRADUATE SCHOOL

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE

MASTER OF SCIENCE

DEPARTMENT OF BIOLOGICAL SCIENCES

Thesis Director:

Gabriel P. Holbrook

Page 4: The Monoraphidium - Northern Illinois University

ACKNOWLEDGEMENTS

I would like to thank many individuals for aiding me in the completion of my project.

First and foremost, I would like to thank Dr. Gabriel P. Holbrook for accepting me into his lab in

the fall of 2013. I am very appreciative of the opportunity he gave me and I cannot express how

grateful I am for the help he has provided me through my time here at NIU. I would like to

thank my fellow graduate students, Anthony Kephart and Adam Hage, for their advice and help.

Without them, the completion of my project would not have been completed as smoothly. I

would also like express my gratitude to Dr. W. Scott Grayburn and Dr. Neil W. Blackstone for

joining my committee and advising me on potential paths to take in my research. I also am very

appreciative of the help provided to me by Lori Bross and Dr. Kalyan Karumanchi in the

development of my Nile red protocol for Monoraphidium sp. Dek19.

I would like to thank the Department of Biological Sciences at Northern Illinois

University for providing me support for the last three semesters. I would also like to thank the

DeKalb Sanitary District for access to their facilities and use of effluent. I acknowledge the

Midwestern ASPB for providing travel support to their conferences in spring of 2014 and 2015. I

also acknowledge the Venturewell agency for providing the Holbrook lab a level 1 E-team grant

and inviting me to participate in their workshop at MIT in Boston.

I would like to thank my mother, father, sister, and fiancée. This project was difficult to

complete, but it was made easier with their emotional support.

Page 5: The Monoraphidium - Northern Illinois University

iii

TABLE OF CONTENTS

Page

LIST OF TABLES ......................................................................................................................... vi

LIST OF FIGURES ...................................................................................................................... vii

Chapter

1. INTRODUCTION ..........................................................................................................1

Monoraphidium as a Genus in the Family Selenastraceae ......................................1

Photosynthesis in Green Microalgae .......................................................................3

Lipid Biosynthesis in Algae .....................................................................................4

Nutrient Uptake by Algae ........................................................................................5

Water Treatment Regulations ..................................................................................6

Wastewater Treatment Process at the DeKalb Sanitary District .............................7

Possible Benefits of Microalgae when Introduced to Wastewater ........................10

Demand for Renewable Energy .............................................................................11

Research Goals.......................................................................................................14

2. MATERIALS AND METHODS ..................................................................................15

Identification of Algal Species...............................................................................15

Effluent Collection .................................................................................................16

Culture Setup .........................................................................................................17

Absorbance Measured at E680 ..............................................................................18

Absorption Spectra of Monoraphidium sp. Dek19 Pigments ................................19

Separation of Monoraphidium sp. Dek19 from a Mixed Consortium of Algal

Species ...................................................................................................................20

Photosynthetic Rates of Monoraphidium sp. Dek19 .............................................21

Page 6: The Monoraphidium - Northern Illinois University

iv

Chapter Page

Nutrient Depletion in Algal Cultures .....................................................................24

Nitrates .......................................................................................................24

Ammonium ................................................................................................24

Phosphates..................................................................................................25

Fluorescence Microscopy ......................................................................................25

Cell Imaging...............................................................................................25

Quantification of Lipids via a Fluorescence Spectrophotometer ...............26

Quantification of Lipids via ImageJ ..........................................................26

3. RESULTS .......................................................................................................................27

Relationship of Light Extinction to Cell Density ..................................................27

Effect of Initial Population Density (IPD) of Monoraphidium sp. Dek19 on

Growth of Batch Cultures ......................................................................................28

Species Competition at Differing Temperatures....................................................32

Sucrose Density Gradients .....................................................................................34

Photosynthetic Rates of Monoraphidium sp. Dek19 .............................................37

Photosynthetic Oxygen Evolution per Total Chlorophyll .........................37

Photosynthetic Oxygen Evolution per Million Cells .................................41

Chlorophyll Levels in Monoraphidium sp. Dek19 During Photosynthesis

Measurements ............................................................................................46

Pollutant Remediation by Growth of Monoraphidium sp. Dek19 .........................48

Initial Nutrient Levels ................................................................................48

Nutrient Uptake in Final Effluent and Post-Primary Filtration Effluent ...48

Lipid Quantification ...............................................................................................53

Nile Red Microscopy .................................................................................53

Page 7: The Monoraphidium - Northern Illinois University

v

Chapter Page

Fluorescence Spectrometry ........................................................................53

4. DISCUSSION ...............................................................................................................58

Summary ................................................................................................................58

Effect of Initial Population Density (IPD) on Growth of Cultures ........................59

Species Competition ..............................................................................................60

Sucrose Density Gradients .....................................................................................62

Photosynthetic Rates of Monoraphidium sp. Dek19 .............................................65

Phycoremediation ..................................................................................................68

Lipid Quantification ...............................................................................................70

5. CONCLUSIONS...........................................................................................................72

REFERENCES ..................................................................................................................74

Page 8: The Monoraphidium - Northern Illinois University

vi

LIST OF TABLES

Table Page

1. Effects of Initial Population Density (IPD) as Quantified by Extinction of Light at 680nm

(E680) of Monoraphidium sp. Dek19 on Peak Cell Densities/mL Measured at 680nm,

Time to Log Phase, and Time to Stationary Phase. ...........................................................31

2. Chlorophyll a:b Ratios of Monoraphidium sp. Dek19 for Cultures Grown at 10°C and

22°C in Final Effluent at 48.3 ± 5.4 µmol Photons m-2s-1 .................................................45

3. Chlorophyll a:b Ratios of Monoraphidium sp. Dek19 for Cultures Grown at 10°C and

22°C in Post-Primary Filtration Effluent at 48.3 ± 5.4 µmol Photons m-2s-1 .....................45

4. Chlorophyll per Cell Expressed as pg/Cell in Monoraphidium sp. Dek19 Cultures Grown

in Both Final and Post-Primary Filtration Effluent as well as 10°C and 25°C ..................45

5. Dilution of Stationary Growth Phase Cultures and the Effect on Light-Saturated Oxygen

Evolution Rates at 300 µmol Photons m2s-1 ......................................................................47

6. Average Initial Nutrient Levels of Wastewater Effluent from the DeKalb Sanitary District

(DSD) at Time of Collection..............................................................................................48

Page 9: The Monoraphidium - Northern Illinois University

vii

LIST OF FIGURES

Figure Page

1. Light Micrograph of Monoraphidium sp. Dek19 Using a Nikon Eclipse E-600 Light

Microscope .........................................................................................................................16

2. Aerial View of the DSD from Google ...............................................................................17

3. Flasks Containing Monoraphidium sp. Dek19 Grown in an Environmental Growth

Chamber .............................................................................................................................19

4. Hansatech Oxygen Electrode in Which Photosynthetic Rates Were Obtained .................23

5. Output of the Oxygen Electrode Signal in Logger Pro Showing O2 Evolution from

Illuminated Monoraphidium cells in 1ml Final effluent ....................................................23

6. Relationship Between Monoraphidium sp. Dek19 and Light Extinction at 680nm in Final

Effluent (N=25) ..................................................................................................................28

7. Effects of Initial Population Density (IPD) in Cells/mL (via Hemocytometer Count) of

Monoraphidium sp. Dek19 Grown at 25°C in 1L Flasks on Peak Cell Density, Time to

Log Phase, and Time to Stationary Phase ..........................................................................30

8. Effect of Growth Temperature on Algal Species Composition over Three Weeks for

Cultures Shown in Figure 9 ...............................................................................................33

9. Growth (E680) Curves of the Consortium of Species .......................................................33

10. A Mixed Microalga Culture Is Centrifuged Through a Sucrose Gradient.........................36

Page 10: The Monoraphidium - Northern Illinois University

viii

Figure Page

11. Light Saturation Curve of Photosynthetic Oxygen Evolution per Weight of Total

Chlorophyll of Monoraphidium sp. Dek19 Grown in Final Effluent ................................38

12. Light Saturation Curve of Photosynthetic Oxygen Evolution per Weight of Total

Chlorophyll of Monoraphidium sp. Dek19 Grown in Post-primary Filtration Effluent....40

13. Light Saturation Curve of Photosynthetic Oxygen Evolution per Million Cells of

Monoraphidium sp. Dek19 Grown in Final Effluent .........................................................43

14. Light Saturation Curve of Photosynthetic Oxygen Evolution per Million Cells of

Monoraphidium sp. Dek19 Grown in Post-Primary Filtration Effluent ............................44

15. Representative Nutrient Removal Graph of Nitrate in Final Wastewater Effluent ...........50

16. Representative Nutrient Removal Graph of Phosphate in Final Wastewater Effluent ......50

17. Representative Nutrient Removal Graph of Ammonium in Post-Primary Filtration

Wastewater Effluent...........................................................................................................51

18. Representative Nutrient Removal Graph of Phosphate in Post-Primary Filtration

Wastewater Effluent...........................................................................................................52

19. Nile Red Microscopy Images of Monoraphidium sp. Dek19 Grown in Final Wastewater

Effluent ..............................................................................................................................54

20. Nile Red Microscopy Images of Monoraphidium sp. Dek19 Grown in Post-Primary

Filtration Wastewater Effluent ...........................................................................................55

21. Calculated Total Cell Fluorescence (CTCF) of Monoraphidium sp. Dek19 Grown in Both

Final and Post-Primary Filtration Wastewater Effluent and Tested During Both Log and

Stationary Growth Phases ..................................................................................................56

Page 11: The Monoraphidium - Northern Illinois University

ix

Figure Page

22. Representative Data Showing an Estimation of Comparative Lipid Concentrations via a

Hitachi F2500 Fluorescence Spectrophotometer ...............................................................57

Page 12: The Monoraphidium - Northern Illinois University

1

INTRODUCTION

Monoraphidium as a Genus in the Family Selenastraceae

Monoraphidium sp. Dek19 has previously been described as a cold-tolerant algal species

native to the Midwest (Holbrook et al., 2014). Monoraphidium is a genus in the family

Selenastraceae. More specifically, the Selenastraceae are in the order Sphaeropleales. This order

contains microalgae such as Scenedesmus, Ankistrodesmus, and Selenastrum. These algae are

classified together as they are all non-motile and unicellular green microalgae. This order resides

within the division of Chlorophyta (Lewis & McCourt, 2004). These microalgae are typically

morphologically similar, having long cells that come to a point at either end or are in the shape of

a crescent moon (Fawley, Dean, Dimmer, & Fawley, 2005). The main means of reproduction in

Monoraphidium sp. is thought to be asexually through autosporulation (Krienitz, Ustinova, Friedl,

& Huss, 2001).

Monoraphidium sp. Dek19 is a cold-tolerant alga that is capable of growth at low light

intensities (Holbrook et al., 2014). It has successfully been able to grow in “final” wastewater

effluent. This effluent comes from the settling tanks just prior to chlorination. To determine the

utility of Monoraphidium sp. Dek19 in the treatment of wastewater, it should be determined in

which stages of wastewater treatment that the algae are capable of growth. However, little is known

about the viability of microalgae at different stages of treatment. Some research has been

completed with Chlorella sp. as to growth in different stages of wastewater (L. Wang et al., 2010),

but no one has yet studied the growth of Monoraphidium sp. or related species in various steps of

Page 13: The Monoraphidium - Northern Illinois University

2

water treatment. Monoraphidium sp. Dek19 may be able to further cut operating costs of

wastewater treatment facilities via nutrient remediation if the alga is capable of growth earlier in

the treatment process. When grown outdoors in a non-sterile environment such as a wastewater

treatment facility, it is quite possible that growth media will become inoculated by many types of

microalgae. It is important that the competitiveness of Monoraphidium sp. Dek19 be studied as

part of its development as an introduced species leading to accumulation of biomass and

phycoremediation of wastewater. Should an unwanted consortium of microalgae form, it could be

useful to select for microalgae that are able to perform optimally at the given growing conditions.

Some microalgae are able to take up nutrients in wastewater at a faster rate. Others may grow

better at different temperatures or at different light intensities. Further, some microalgae produce

more lipids than others. Either way, there are options to select for algae that best fit a given

situation.

There are a few ways to select for a given microalga involving density gradient cell

preparations. One way is the use of Ludox-TM, but this is toxic to the algae and would result in a

loss of viability of the algal cells (de Jonge, 1979). A second and more common way to separate

the microalgae would be through the use of the reagent Percoll (Schwinghamer, Anderson, &

Kulis, 1991; Whitelam, Lanaras, & Codd, 1983). This may be the most effective way to select for

a given microalga, but Percoll is very expensive and would not be a financially viable option at

large-scale production. An inexpensive way to separate out unwanted microalgae species is

desirable if grown at a large scale.

Page 14: The Monoraphidium - Northern Illinois University

3

A proposed way to cut operating costs would be to provide oxygenation to the activated

sludge treatment step as an alternative to pumping compressed air into the tanks, which can account

for ~60% of the energy consumption of a typical wastewater treatment plant. Monoraphidium sp.

Dek19 are capable of evolving enough oxygen to possibly supply the aeration needed for the

activated sludge treatment of wastewater. Oxygen evolution rates have previously been studied in

a variety of algae including Scenedesmus sp. and Chlorella sp. (Godos et al., 2010), but not in

Monoraphidium.

Photosynthesis in Green Microalgae

Green microalgae perform the essential photosynthetic process to convert readily available

oxidized carbon (CO2) to a reduced state of carbon in the form of carbohydrates. The simplest of

these carbohydrates is glucose, which can then be converted to structural carbohydrates such as

cellulose or energy storage such as starch. Photosynthesis occurs as two sets of reactions: light

dependent and light independent. Light-dependent reactions occur in the thylakoid membranes.

Photosystems and chlorophyll are imbedded within the thylakoid membrane. Chlorophylls are

pigments that harvest light and convert it into energy. Specifically, green microalgae contain

chlorophylls a and b. Chlorophyll a is a necessary pigment that is required for photosynthesis to

occur as it moves energized electrons to the molecules necessary for forming carbohydrates.

Chlorophyll b is a light-harvesting accessory pigment that absorbs wavelengths of light that are

not absorbed by chlorophyll a, thus increasing the efficiency of photosynthesis.

Two photosystems are necessary for light reactions within the thylakoid membrane, each

of which is surrounded by light-harvesting complexes (LHCs). The main responsibility of LHCs

Page 15: The Monoraphidium - Northern Illinois University

4

is to absorb energy from sunlight and then transmit that energy to chlorophyll. Photosystem I (PSI)

and Photosystem II (PSII) excite special chlorophyll a pigments P700 and P680, respectively,

which moves electrons across the thylakoid membrane. This movement of electrons through the

thylakoid membrane forms the electron transport chain. This process will eventually reduce

NADP+ to NADPH and create a differential in the proton concentration on either side of the

membrane, which will allow for the synthesis of ATP (Barsanti & Gualtieri, 2014). NADPH and

ATP produced in the light-dependent reactions are then used in the light-independent reactions to

reduce oxidized carbon in the form of carbon dioxide to carbohydrates. Light-independent

reactions occur outside the thylakoids in the stroma. Light-independent reactions are often referred

to as the Calvin cycle. The Calvin cycle is made up of three phases. In the first phase, CO2 is fixed

into ribulose bisphosphate (RuBP) catalyzed by the enzyme RuBisCO. This six-carbon product

then splits into two three-carbon 3-phosphoglyceric acid (3-PG) molecules. In the second phase,

3-PG is phosphorylated via ATP to form 1, 3-biphosphoglycerate, which is then reduced to

glyceraldehyde-3-phosphate (G3P) via NADPH. NADP+ and ADP remaining after this phase are

returned to the thylakoids to regenerate NADPH and ATP in the light-dependent reactions. The

third and final phase of the Calvin cycle regenerates RuBP from a five-carbon intermediate via

ATP from the light-dependent reactions (Calvin & Massini, 1952).

Lipid Biosynthesis in Algae

Lipid biosynthesis in algae occurs in the chloroplast. Microalgae fix carbon dioxide into

sugars via photosynthesis which are then processed to make acetyl-CoA. Acetyl-CoA is the

precursor for lipid synthesis. In algae, lipid biosynthesis is initiated by two enzymes, acetyl-CoA

Page 16: The Monoraphidium - Northern Illinois University

5

carboxylase (ACCase) and type-II fatty acid synthase, which allows for the carboxylation of

acetyl-CoA to form malonyl-CoA. The malonyl group is then transferred from coenzyme-A to an

acetyl carrier protein. Fatty-acid chains are formed by the addition of two-carbon chain-

lengthening reactions catalyzed by a fatty-acid synthase. Photosynthesis provides the additional

amounts of ATP, NADPH, and acetyl-CoA required by lipid biosynthesis to add on each new two

carbon chain (De Bhowmick, Koduru, & Sen, 2015; Hu et al., 2008). The lipids most desired for

the purposes of biofuel production are the triacylglycerols (TAGs). TAGs are small carbon- chain

fatty acids attached to a glycerol backbone. These fatty acids can be removed and converted into

fuel (Wahlen, Willis, & Seefeldt, 2011).

Nutrient Uptake by Algae

For algae to grow, they require inorganic nutrients, sunlight, and a temperature range in

which the algae can complete necessary metabolic functions. Two inorganic nutrients that are

necessary to the growth of algae are phosphorous and nitrogen. Algae use phosphorus in the cell

membranes, enzymes, DNA, RNA, and ATP (Barsanti & Gualtieri, 2014). Natural deposits of

phosphorous are usually in the form of phosphate, caused by erosion of rock that is exposed to

water. Algae can take up phosphate in the orthophosphate form through an active process at the

cell surface. In natural conditions, phosphorous is a limiting nutrient. Due to this, when algae are

exposed to sources of phosphorous, they quickly take up the nutrients and store excess as

polyphosphate granules. These reserves allow the algae to grow for long periods of time without

being near a new source of phosphorus (Oliver & Ganf, 2000).

Page 17: The Monoraphidium - Northern Illinois University

6

Nitrogen is used in the production of chlorophyll, enzymes, and other proteins, as well as

ATP, ADP, RNA, and DNA (Barsanti & Gualtieri, 2014). Nitrogen removal from a nutrient source

occurs by assimilation of the nitrogen into the cells. The two most common forms of nitrogen used

by algae are ammonium and nitrate. If a microalga were exposed to both forms of nitrogen, the

preferred form to use is ammonium (Oliver & Ganf, 2000). If an excess of either nutrient is

provided to a lake or stream, this can result in eutrophication of the natural waterways due to

overproduction of algae or cyanobacterial biomass. We believe it would be feasible to encourage

controlled algal growth in treatment plants and incorporate their use for phycoremediation of the

water before its discharge into the environment.

Water Treatment Regulations

In 1972, the Clean Water Act (CWA) was passed to attempt to maintain and improve the

quality of the United States’ waterways. This undertaking began with the issuing of NPDES

(National Pollutant Discharge Elimination System) permits that would limit the amount of

nutrients allowed in wastewater from industry and municipal wastewater treatment facilities. As

of 2004, there were approximately 16,000 municipal wastewater treatment facilities operating in

the United States ("Primer for Municipal Wastewater Treatment Systems," 2004). All of these

wastewater treatment facilities must hold a uniquely tailored permit to achieve water quality

standards at a given location. Locations located near waterways are especially regulated ("NPDES

Wastewater & Stormwater Permits," 2015). Punishments for not meeting the given water quality

standards can be quite severe. If a person or company negligently violates the permit, they are

subject to a fine of $2,500-$25,000 per day of violation or a prison sentence up to a year. The fines

Page 18: The Monoraphidium - Northern Illinois University

7

are even worse if the person or company knowingly violated the permits. The fines double per day

and the people involved could then be sentenced to prison for up to three years ("Construction

General Permit - Standard Permit Conditions," 2012). As the punishments for not meeting the

expectations of the NPDES permits are so severe and regulation of discharge surely will become

more stringent, it is imperative that more efficient treatments of wastewater be explored. Not only

will this benefit the environment of local waterways, but it should provide cost savings to treatment

facilities that undergo these changes.

One of DeKalb County’s wastewater treatment facilities (DeKalb Sanitary District, or

DSD) is one of the 16,000 municipal wastewater treatment facilities in operation. Like all of the

others, the DSD must stay compliant to the EPA-written NPDES permits. The major concern of

the EPA for the DSD is the quality of the south branch of the Kishwaukee River ("NPDES Permit

No. IL0023027," 2011). This permit states the monthly average and daily maximum values of

waste allowed to be discharged into the Kishwaukee River. Parameters observed include

suspended solids, pH, fecal coliform, chlorine, ammonia, phosphorous, total nitrogen, etc. To

better understand where wastewater treatment facilities can be improved, it is essential to

understand the treatment of wastewater at a given treatment facility.

Wastewater Treatment Process at the DeKalb Sanitary District

As wastewater makes its way from DeKalb County into the DSD, large objects are

separated from the wastewater via a mechanical preliminary treatment in which the water is passed

through a set of metal bar screens. This separates larger objects out of the wastewater such as

diapers, sticks, and anything else that should not have been deposited into the wastewater system.

Page 19: The Monoraphidium - Northern Illinois University

8

Next, the wastewater is pushed through a finer set of screens that remove sand, gravel, etc., from

the wastewater ("Preliminary Treatment," 2015). The next step is completed by pumping the

wastewater into one of three settling tanks in which particles are allowed to sediment. Any particles

that settled or floated on top of the water are removed from the system ("Primary Treatment,"

2015). At this point, the post-primary filtration effluent moves through to one of two different

types of secondary biological treatment. The first type of secondary biological treatment is an older

method in which the wastewater is sprayed onto rocks that are covered with microorganisms

(trickling filters). This process allows microorganisms to slowly take up nutrients in the

wastewater. After passing through the trickling filters, wastewater moves into biodiscs that mix

the microorganisms again into the wastewater to uptake more nutrients. The wastewater is then

passed into a settling tank, at which point ~95% of nutrients have been removed ("Secondary

Treatment," 2015).

The second type of wastewater treatment method, activated sludge, used at the DSD is a

newer and more efficient way to treat the municipal wastewater effluent ("Activated Sludge,"

2015). After primary filtration, wastewater is allowed to move to a tank and introduced to a

saturated batch of aerobic bacteria which is then suspended via aeration. This is a much more

efficient and faster method than the older methods of secondary treatment (13 hours for older and

6 hours for newer), but the downsides are increased expenses in maintaining constant aeration rates

to maintain the O2 at 1-3 mg/L of oxygen necessary for optimum growth of the bacteria

("Introduction to Activated Sludge Study Guide," 2010).

Page 20: The Monoraphidium - Northern Illinois University

9

The final stage of biological treatment consists of running the treated wastewater from both

the trickling filter/biodiscs and the activated sludge treatment through a bed of gravel and sand.

The intention of this is to remove any finer particles that may have made their way through the

treatment process. Last, the wastewater is put through a chemical treatment process to provide a

final disinfection. Wastewater is chlorinated by mixing it into a tank of bleach to kill any bacteria

or other microorganisms in the wastewater. Before discharge into the Kishwaukee River, the

wastewater is dechlorinated and examined to insure that nutrient levels are below daily allowed

levels by their NPDES permit ("Disinfection," 2015).

This final step of nutrient reduction is essential as sources of nitrogen and phosphorous

have been shown to cause eutrophication of waterways and potential negative side effects to human

health (Arora & Saxena, 2005; de-Bashan & Bashan, 2010; Rawat, Ranjith Kumar, Mutanda, &

Bux, 2011). Eutrophication is a direct cause of added nutrients to the waterways. This can lead to

uncontrollable algal blooms and consequent fish kills, both of which are negative outcomes to the

environment (Borsuk, Stow, & Reckhow, 2004). Studies have even shown that excess nitrate

leaching into the drinking water supplies can lead to increased chances of non-Hodgkin lymphoma

(Ward et al., 1996), bladder, and ovary cancer (Weyer et al., 2001). Additional reproductive issues

have been claimed to arise, including premature births (Bukowski, Somers, & Bryanton, 2001),

still births (Aschengrau, Zierler, & Cohen, 1989; Grant, Steel, & Isiorho, 1996), and congenital

malformations (Cedergren, Selbing, Lofman, & Kallen, 2002; Dorsch, Scragg, McMichael,

Baghurst, & Dyer, 1984).

Page 21: The Monoraphidium - Northern Illinois University

10

Possible Benefits of Microalgae When Introduced to Wastewater

Although current biological and chemical treatment of municipal wastewater facilities are

able to meet the nutrient remediation levels required by the Environmental Protection Agency, it

is assumed that more stringent requirements will be set into place. The process of building new

infrastructure, retrofitting older structure, and maintenance of the new/modified infrastructure is

quite costly and may persuade municipalities into adopting different methods to treat wastewater

("Case Studies on Implementing Low-Cost Modifications to Improve Nutrient Reduction at

Wastewater Treatment Plants," 2015). Some groups have suggested that microalgae could be

added into the current wastewater treatment processes in order to offset costs of nutrient removal

(Chinnasamy, Bhatnagar, Claxton, & Das, 2010; Hoffmann, 1998; Mallick, 2002; Olguin, 2003).

Through use of already available infrastructure, microalgae can be incorporated as an extra step in

the wastewater treatment process that may be able to meet future demands of the EPA. Chinnasamy

et al. (2010) have shown that a consortium of 15 local microalgae was able to remediate greater

than 96% of the nutrients from wastewater within 72 hours in 250mL flasks. Although these types

of experiments are currently at a small scale, there is an expectation that these approaches could

improve large-scale treatment of municipal wastewater. Another paper written by Wang et al.

(2010) showed that Chlorella sp. was able to deplete 54-95% of nitrate and phosphate over the

course of 15 days. Similar rates of nutrient removal have also been observed using Spirulina

platenis (Mezzomo et al., 2010), Chlorella sorokiniana (Ogbonna, Yoshizawa, & Tanaka, 2000),

and Botryococcus braunii (Sawayama, Minowa, Dote, & Yokoyama, 1992). Some microalgae

may even be able to provide partial or full oxygenation of the activated sludge treatment step using

only sunlight. The electricity used in aeration of the activated sludge is currently one of the major

Page 22: The Monoraphidium - Northern Illinois University

11

operating costs of the DSD and the majority of similar wastewater treatment facilities.

Scenedesmus obliquus and Chlorella sorokiniana have been shown to evolve 116-133 mg O2 L-1

day-1, which may help maintain the constant 1-3 mg O2L-1 required of the activated sludge

treatment process (Godos et al., 2010; "Introduction to Activated Sludge Study Guide," 2010).

Even though microalgae are able to take up nutrients at high rates and may possibly lower

operating costs of the activated sludge treatment process, it would be even more cost effective to

recycle the biomass accumulated for use as fertilizer, animal feed, or a source for biofuel (Mulbry,

Kondrad, Pizarro, & Kebede-Westhead, 2008).

Demand for Renewable Energy

Although the price of crude oil is the lowest it has been since 2009 ($44.24 per barrel as of

11/10/15) ("Crude Oil," 2015), fossil fuels are being consumed at an unsustainable rate. Due to an

increased need for energy due to an exponentially increasing world population and

industrialization, fossil fuels are not a secure source of fuel. In fact, fossil fuels are a finite resource

that some expect to be depleted by 2075 if the rate of use in 2002 does not change (The Colorado

River Commission of Nevada, 2012). As global energy consumption has consistently been rising

(+2.0% over 2013 and +0.9% over 2014), this depletion date has likely been moved forward ("BP

Statistical Review of World Energy June 2015," 2015). Due to these concerns, governments have

spent a large sum of money on research of renewable energy. In 2005, the Environmental

Protection Agency created the Renewable Fuel Standard (RFS) program to require certain volumes

of renewable fuels to replace a portion of fossil fuels used in transportation. One of the four

renewable fuels that were considered under RFS was biomass-based fuel, an example of which

Page 23: The Monoraphidium - Northern Illinois University

12

could be microalgae-based biofuel ("Program Overview for Renewable Fuel Standard Program,"

2005). Two years later, the RFS was further expanded by the addition of the Energy Independence

and Security Act (EISA). The EISA aimed towards reduction of the United States dependence on

petroleum-based fuels. The EISA required the introduction of at least 36 billion gallons of

renewable fuel into the fuel sold for transportation within the United States annually by 2022.

Many types of biofuel feedstock have been proposed to meet these demands. Most notably,

corn has been grown for the purposes of biofuel production. However, there are quite a few

downsides to current biofuel production. The growth of these crops is only financially viable due

to large government subsidies (de Gorter & Just, 2010). In 2000, 90% of the corn in the United

States went directly to feeding people and livestock, with the remainder converted to ethanol. By

2013, 60% went directly to feeding people and livestock while 40% went to the conversion of

ethanol. In 2013, the United States used 130 billion gallons of gasoline and 50 billion gallons of

diesel. With the current conversion rate of a bushel of corn to ~3 gallons of corn ethanol fuel, if

all of the corn grown in the United States were converted to ethanol fuel production, that would

only account for a quarter of the 130 billion gallons of fuels (of all types) consumed (Agricultural

Marketing Resource Center, 2014). Other potential sources for a biofuel feedstock would not even

be able to yield enough biodiesel to meet 50% of the required fuel for transport. Soybeans would

require 326% of the existing United States crop area, 122% for canola, 54% for coconut (palm

oil), etc. Even if those crops were able to grow year round in all climates, there would still not be

enough farmland to supply the needs of the United States (Chisti, 2007). Not only would those

crops not meet demands for biofuel, but there would no longer be any land to grow crops for animal

and human consumption.

Page 24: The Monoraphidium - Northern Illinois University

13

Microalgae circumvent many of these issues. There would no longer be a debate on

whether or not the crop is used for human consumption. Microalgae have the capability to grow in

wastewater treatment facilities and/or on 2.2-5% of the existing United States cropland to

completely satisfy 100% the public’s need for fuel (Chisti, 2007). Microalgae are capable of

producing lipids throughout their growth. These can be stored as phospholipids or triacylglycerides

(TAGs). TAGs are capable of conversion to biodiesel via transesterification reactions (Sharma,

Schuhmann, & Schenk, 2012). For this reason, it is important to identify algal species with lipids

accounting for a large percentage of dry weight. Common levels of oil per dry weight in microalgae

are 20-50%. Chlorella sp. has been shown to have 29-32% lipid per dry weight and

Nannochloropsis sp. has been shown to have 31-68% lipid per dry weight (Chisti, 2007).

Depending on the algal species, there will be variable amounts of lipids that are available for

conversion to biofuel. If the DSD were to adopt an algal-based phycoremediation step into their

current infrastructure, the facility would not only cut the operating costs of nutrient removal but

would have the benefit of being able to convert the biomass into biofuel. The sale of this biofuel

would most likely reduce the loss in capital each year and possibly allow DSD to pass on savings

to connected residents.

To many research groups, microalgae seem like the perfect feedstock for the production of

biofuel; however, some do not feel that algae is economically feasible at its current state. Many

of the species of algae currently being used in the study of algal biofuel production, most notably

Chlorella sp. and Nannochloropsis sp., require warmer temperatures of 20-30°C (Converti,

Casazza, Ortiz, Perego, & Borghi, 2009; Olofsson et al., 2012; Z. T. Wang, Ullrich, Joo,

Waffenschmidt, & Goodenough, 2009) at high light intensities comparable to natural sunlight.

Page 25: The Monoraphidium - Northern Illinois University

14

This would only allow year-round growth to occur in a small portion of the United States. Also, at

these light intensities required for growth, the algae would not be able to be grown in larger settling

tanks at wastewater treatment facilities. This would restrict where and how the microalgae are

grown for the purposes of producing biofuel. One way to bypass these issues would be to find an

algae that can be grown at a wider range of temperature and light intensity.

Research Goals

The goals of this research are to answer the following questions: A) Can Monoraphidium

sp. Dek19 be grown successfully in post-primary filtration and final wastewater effluent? B) Could

Monoraphidium sp. Dek19 be introduced as a potential biofuel feedstock if grown in wastewater

effluent? C) How competitive is Monoraphidium sp. Dek19 in a consortium of species and how

does temperature affect their competitiveness? D) How does growth at 10°C and an ambient lab

temperature (22-25°C) affect the photosynthetic properties of Monoraphidium sp. Dek19 and

would the cells be able to adapt to conditions seen year round at a midwestern wastewater treatment

facility? E) Could the use of sucrose as a density gradient be an effective preparative method of

separating Monoraphidium sp. Dek19 from other microalgae? F) Can lipids be

visualized/quantified in Monoraphidium sp. Dek19 and when would be the ideal time to harvest

the cells?

Page 26: The Monoraphidium - Northern Illinois University

15

MATERIALS AND METHODS

Identification of Algal Species

Algae were obtained from flasks in an environmental growth chamber housed in

Montgomery Hall at Northern Illinois University (NIU). The flasks were prepared by a previous

graduate student, Zachary Davidson, and contained a consortium of local microalga. These

species included Chlorella sp., Ulothrix sp., and Monoraphidium sp. Dek19. Monoraphidium sp.

Dek19 (Figure 1) was previously identified through the use of 18S rRNA gene sequencing by Dr.

W. Scott Grayburn of NIU (Holbrook et al., 2014). Chlorella sp. and Ulothrix sp. were keyed

out through utilizing an algae ID text, titled Identification, Ecology and Control of Nuisance

Freshwater Algae (St. Amand, 2012). This was then compared to online sites to confirm the

keys.

Page 27: The Monoraphidium - Northern Illinois University

16

Figure 1 - Light Micrograph of Monoraphidium sp. Dek19 Using a Nikon Eclipse E-600 Light

Microscope.

Effluent Collection

Effluent was collected multiple times over the past two years (2013- 2015) from the

DeKalb Sanitary District (41°56’37.8” N, 88°44’26.7” W) (Figure 2). On each occasion, a sump

pump was lowered into the final settling tanks and wastewater was pumped into 5-gallon

Culligan bottles or a 210-gallon polyethylene tank on the bed of a 1-ton pickup truck. For post-

primary filtration effluent, the sump pump was lowered into a channel off a settling tank earlier

in the treatment process. Centrate was collected by DeKalb Sanitary District (DSD) employees

inside 5-gallon bottles and was picked up at a later point. Immediately upon return to the lab, the

Page 28: The Monoraphidium - Northern Illinois University

17

effluent was autoclaved to remove any bacteria or zooplankton that could prove harmful to

personnel or the experiments.

Figure 2 - Aerial View of the DSD from Google. Wastewater effluent collection occurred at the

DeKalb Sanitary District (DSD).

Culture Setup

Algae was grown in wastewater effluent taken from the DeKalb Sanitary District (DSD).

Most cultures consisted of final effluent taken from the settling tank just prior to being

discharged into the Kishwaukee River. Other cultures consisted of post-primary filtration

effluent which was taken from an earlier step in the wastewater treatment process, and a few

cultures were also inoculated in centrate. Centrate is the excess wastewater after solids have been

Page 29: The Monoraphidium - Northern Illinois University

18

spun out of solution in the biocentrifuge. Algae cultures were grown in 1L or 2L flasks and 20L

buckets in the lab as well as an environmental growth chamber (Figure 3). Cultures grown in the

lab were grown in variable conditions based on experiment. All light intensity and temperature

measurements of each culture were measured with Onset HOBO data loggers and confirmed

with a photometer and thermometer. All lab cultures were grown at 23.4 ± 1.1 °C with an

aeration rate of ~1L per minute. Light intensities and light cycles varied between experiments

from 8-60 µmol photons m-2s-1 and 24:0 or 13:11 dark:light cycles (D:L). In the environmental

growth chamber, 1L and 2L cultures were grown at 48.3 ± 5.4 µmol photons m-2s-1 at 10°C in a

13:11 D:L with an aeration rate ~1L per minute. Inoculum was added to previously autoclaved

effluent wastewater. Starting cell densities were confirmed by measuring E680 and cell counts

via a hemocytometer.

Absorbance Measured at E680

Extinction of light at 680nm (E680) was collected for all algae grown via a Pharmacia

LKB Ultrospec III UV/Vis spectrophotometer. These readings are a metric that was used to

monitor growth of cultures as it reflects the absorbance of chlorophyll, the primary pigment

found in green algal cells (Griffiths, Garcin, van Hille, & Harrison, 2011). A blank of distilled

water was used for final effluent absorption readings as it was determined that final effluent

showed a minimal difference when compared to distilled water (~0.001). A blank of post-

primary filtration wastewater effluent was used for the post-primary filtration effluent absorption

readings, as the post-primary filtration effluent varied in turbidity among different collections.

Page 30: The Monoraphidium - Northern Illinois University

19

Manual cell counts were completed in triplicates using a phase microscope at 430X and a

hemocytometer.

Figure 3 - Flasks Containing Monoraphidium sp. Dek19 Grown in an Environmental Growth

Chamber.

Absorption Spectra of Monoraphidium sp. Dek19 Pigments

Absorption spectrum data was taken in two separate ways. In both, 1mL samples were

centrifuged to a pellet and the supernatant was removed. The pellet was resuspended in 1mL

analytical grade dimethyl sulfoxide (DMSO) through the use of a vortex. Samples were allowed

Page 31: The Monoraphidium - Northern Illinois University

20

to sit in the refrigerator overnight in the dark at 4°C for approximately 12 hours (Shinano,

Kawamukai, Inoue, Koike, & Tadano, 1996). For the first method, samples were read in a

spectrophotometer at 10nm intervals from 400-700nm. This allowed me to observe the pigments

presented in the spectrum through the two peaks. Consistently the observed peaks were

chlorophyll a and chlorophyll b. The second method was through the use of a Perkin-Elmer

Lambda 19 UV/Vis spectrometer. This produced an absorbance spectrum in seconds via

computer that measured absorbance of the sample from 300-800nm in 1nm intervals. With this

spectrum data, I was able to determine chlorophyll a and chlorophyll b concentrations in each

sample. Chlorophyll a was expressed in μg Chl a/mL determined via the following equation: Chl

a = 14.85 x A665 – 5.14 x A648. Likewise, chlorophyll b was expressed in μg Chl b/mL

determined via a similar equation: Chl b = 25.48 x A648– 7.36 x A665 (Shinano et al., 1996).

Separation of Monoraphidium sp. Dek19 from a Mixed Consortium of Algal Species

A variety of concentrations of sucrose solution were made and after a multitude of

experiments a gradient was found in which the Monoraphidium sp. Dek19 could be selected out

of the consortium of mixed algal species. Through the use of a peristaltic pump, sucrose was

slowly layered into Nunc 50mL conical tubes as to not accidently mix layers. The most efficient

gradient consisted of a bottom layer of 5mL of 2.5M sucrose with layers above it of 10mL of

2.0M, 10mL of 1.0M, and 15mL of 0.5M. Prior to loading the mixed microalgae culture onto the

sucrose gradient, 800mL of the given culture was centrifuged at 15,530xg in a Sorvall RC6 Plus

centrifuge for 15 minutes, in which time a pellet was formed. The supernatant was removed and

the pellet was resuspended in 20mL of final wastewater effluent; then 1mL of this concentrate

Page 32: The Monoraphidium - Northern Illinois University

21

was layered onto the top of each sucrose gradient. The gradients were then centrifuged at

1,845xg for 35 minutes in a Thermo IEC Centra GP6R. This resulted in two distinct layers of

algae within the sucrose gradient. The bottom-most layer was extracted and shown to be pure

Monoraphidium sp. Dek19. Before inoculating a new culture with this algae, the algal cells were

washed multiple times with distilled water to remove any excess sucrose on the outside of the

cells.

Photosynthetic Rates of Monoraphidium sp. Dek19

Photosynthetic rates were measured via a Hansatech Instruments oxygen electrode. This

is a delicate system and needs to be treated as such. Before each use, the central platinum

cathode and concentric silver anode of the S1 oxygen electrode was polished to remove

oxidation. Next, concentrated KCl was placed in the well of the electrode. A membrane of Rizla

cigarette paper and Teflon tape was then fitted onto the electrode to provide an even layer of the

electrolyte to both the platinum cathode and the silver anode. This was done by fitting a small

plastic O-ring over the membrane/cathode via an O-ring applicator. Next, the electrode was

covered by a cylindrical container that was cooled/heated via flow of water. The electrode was

then plugged into the electrode control box that measures oxygen diffusion through the

membrane and these numbers were then graphed via a chart recorder.

To get accurate results, an effective range of measurement was determined by first

bubbling N2 into a 1mL sample of distilled water. After a zero is determined on the chart reader,

atmosphere is pumped into the distilled water sample to determine the O2 solubility standard. A

1mL sample of algae was then placed into the machine and was zeroed with addition of N2. Light

Page 33: The Monoraphidium - Northern Illinois University

22

saturation curves were made by increasing light intensity in 3-minute increments. A TCP

2R301635K CFL (compact fluorescent light) 65 watt equivalent (16W) flood-light lamp was

used to supply light to this 1mL sample and light intensities were measured with a photometer.

Light was moved manually from approximately one meter away until a few centimeters from the

oxygen electrode. This data provides the user a graph of oxygen evolution rates shown as slope

(Figure 4). Oxygen evolution rates were then measured in two different ways. The first and most

commonly used was manually drawing tangent lines to the curve at each varying light intensity.

The second method used was by attaching the potentiometer box to a Vernier Lab Pro with the

addition of the software Logger Pro. The Logger Pro effectively transmits analog signals to a box

that is then connected to a computer via USB (Figure 5). This data was exported to a Microsoft

Excel workbook and the slopes were calculated to determine oxygen evolution rates at the

various light intensities. To determine oxygen evolution rates at different temperatures, a

constant was multiplied towards the rate of photosynthesis to express the amount of oxygen that

can be dissolved in 1mL of pure water. The values were 0.359 µmol/mL for 10°C and 0.272

µmol/mL for 25°C ("Oxygen Solubility in Fresh and Sea Water," 2015).

Page 34: The Monoraphidium - Northern Illinois University

23

Figure 4 – Hansatech Oxygen Electrode in Which Photosynthetic Rates Were Obtained.

Figure 5- Output of the Oxygen Electrode Signal in Logger Pro Showing O2 Evolution from

Illuminated Monoraphidium Cells in 1ml Final Effluent.

Atmosphere

Nitrogen

Oxygen Evolution

Respiration

Page 35: The Monoraphidium - Northern Illinois University

24

Nutrient Depletion in Algal Cultures

Nitrates

The nitrates were measured via the colorimetric Szechrome NAS nitrate assay

(Polysciences, 2007). In triplicates, 0.1mL of sample supernatant was added to 0.9mL of

Szechrome reagent. The samples were vortexed and allowed to sit for 30 minutes. The sample’s

E570 nm value was measured on a spectrophotometer (Friedmann & Kiebler, 1980). These

readings were then compared to a standard curve to convert to mg/L. Standard curves were made

up from NaNO3 consisting of 0, 3.125, 6.25, 12.5, 25, 50, 75, and 100 mg/L concentrations.

These values were then corrected to account for solely NO3-. Nitrates were also measured using

an Oakton Ion 700 nitrate electrode. Algae was centrifuged into a pellet and the supernatant was

extracted and tested that day and/or saved at -20°C for later testing. This was done by adding

0.2mL of an ionic strength adjuster (2M (NH4)2SO4) to 10mL of sample and then slowly stirring

the probe. Values on the probe were given in parts per million.

Ammonium

Ammonium samples were measured using an Oakton Ion 700 ammonium probe. This

was done by adding 0.2mL of an ionic strength adjuster (5M NaCl) to 10mL of sample and then

slowly stirring the probe. Values on the probe were given in parts per million. This was then

converted to mg/L, as the two units are congruent. Standards were made up from NH4Cl in 0, 10,

25, 50, 75, and 100mg/L concentrations. These values were then corrected to account for solely

NH4+

ions.

Page 36: The Monoraphidium - Northern Illinois University

25

Phosphates

The phosphates were measured via a colorimetric method. Hanna Instruments 93713-0

(K2S

2O

7) was added to 10mL of sample supernatant and vortexed until completely suspended.

After 3 minutes, samples were measured on a spectrophotometer (at 610nm). Readings were then

compared to a standard curve to convert to mg/L. Standards were made up from KH2PO4 in 0,

3.125, 6.25, 12.5, 25, and 50mg/L concentrations. These values were then corrected to account

for solely PO43- ions.

Fluorescence Microscopy

Cell Imaging

Nile red dye was utilized to investigate the lipid content of Monoraphidium sp. Dek19

cells (Bono, Ahner, & Kirby, 2013; Castro, Larson, Panilaitis, & Kaplan, 2005). Nile red

undergoes a spectral blueshift of fluorescence in nonpolar environments resulting in a

fluorescence maximum in the 500-600nm range (Rumin et al., 2015). Cells were counted

manually via a hemocytometer and samples were diluted to achieve 6x105 cells per mL. Samples

were placed into darkened microcentrifuge tubes to avoid bleaching of the samples when

exposed to light. Samples were taken the day of the experiment and centrifuged at 15,000xg for

15 minutes to create a pellet of cells. The supernatant was removed and discarded. Fifty

microliters of 20% DMSO was added to each sample and vortexed to completely resuspend the

cells. Ten microliters of 0.5ug mL-1 Nile red dye was added to each sample. Samples were

placed in a heat block at 50°C for 10 minutes with vortexing of samples at the halfway point.

Samples were then imaged on a Zeiss LSM 5 Pascal confocal laser scanning microscope.

Page 37: The Monoraphidium - Northern Illinois University

26

Quantification of Lipids via a Fluorescence Spectrophotometer

Methods were similar to cell imaging with exception that samples were diluted to achieve

1.5x104 cells per mL. The cells were sonicated in a Branson 3510 sonicator for 30 minutes to

break apart the algal cells before quantifying fluorescence at 562nm using a Hitachi F-2500

fluorescence spectrophotometer.

Quantification of Lipids via ImageJ

Cell images of Monoraphidium sp. Dek19 from the microscopy lab were imported into

the free software, ImageJ 1.49, on Windows. Cells were selected using the freeform option to

highlight the perimeter of each cell. In the Analyze menu, the button “set measurements” was

chosen and the terms “area,” “integrated density,” and “mean grey value” were checked. Next,

the “measure” button was selected from the Analyze menu, which brought up a popup showing

values for each cell. Next, an area on the image housing no cells was selected and used as a

blank to the fluorescence. This step was done at least three times per selection to insure a proper

blank. This information was then translated to Excel where a simple equation for calculated total

cell fluorescence (CTCF) was determined. Calculated total cell fluorescence is determined by

finding the product of the area of an individual algal cell and the mean fluorescence of the

background readings. This value is then subtracted from the integrated density value (Fitzpatrick,

2014).

Page 38: The Monoraphidium - Northern Illinois University

27

RESULTS

Relationship of Light Extinction to Cell Density

Cell density of Monoraphidium sp. Dek19 in 25 cultures was obtained at inoculation and

during growth of the culture via manual cell counts on a hemocytometer through a Bausch and

Lomb low-power compound microscope (430x). These counts were then compared to the light

extinction (absorbance) values at 680nm, as absorption spectra completed on a UV/Vis showed a

peak at approximately 680nm. This correlation was used to estimate cell densities (cells/mL) in

situations when a manual cell count was not possible due to time constraints. The R2 value was

found to be 0.9399, which indicates that ~94% of the variance in cell number can be explained

by the light extinction values (Figure 6). Factors that may account for the unexplained ~6% of

variance include cell debris within a given sample and/or variation in the chlorophyll content

resulting from changes in growth phase of the culture. This correlation was determined to be

sufficient to quickly estimate cell density of a culture.

Page 39: The Monoraphidium - Northern Illinois University

28

Figure 6 - Relationship Between Monoraphidium sp. Dek19 and Light Extinction at 680nm in

Final Effluent (N=25). The equation defining the line of fit is y=0.2502x+ 0.0112 with an R² of

0.9399.

Effect of Initial Population Density (IPD) of Monoraphidium sp. Dek19 on Growth of Batch

Cultures

Initial population density (IPD) was observed, as previous research on Chlorella sp.

suggested that lower IPD resulted in a lower overall biomass as well as a decreased percent of

lipids per dry weight (Chen, Wang, Liu, & Gao, 2012). In an attempt to determine the most

efficient IPD with which to inoculate Monoraphidium sp. Dek19 cultures, six different ranges of

IPD were investigated in 45 cultures grown at 25°C at 48.3 ± 5.4 µmol photons m-2s-1. Cultures

were inoculated in 1L flasks with log phase growth algae that had been grown at 10°C to maintain

Page 40: The Monoraphidium - Northern Illinois University

29

cell composition at these six IPD ranges: 0-2.96x105 cells/mL (E680 of 0.00-0.100), N=8;

2.96x105–4.63x105 cells/mL (E680 of 0.10-0.15), N=9; 4.63x105–6.29x105 cells/mL (E680 of

0.15-0.20), N=9; 6.29x105-1.296x106 cells/mL (E680 of 0.20-0.40), N=6; 1.296x106-1.963x106

cells/mL (E680 of 0.40-0.60), N=5; 1.963x106-3.296x106 cells/mL (E680 of 0.60-1.00), N=7

(Figure 7 and Table 1). A Tukey HSD (honest significant difference) post hoc test with an α=0.05

was run to determine any statistical difference in average peak cell density/absorbance, average

time until culture reached the log growth phase, and average time until the culture reached

stationary growth phase. This data was compiled after observing the growth of flasks from

inoculation until post-stationary phase culture collapse. Afterwards, it was decided to run the post

hoc test to show any statistical significance based on the patterns that were visualized during the

growth of the cultures.

The results indicated that cultures started at a lower cell density (0-2.96x105 cells/mL) had

a significantly lower peak cell density/peak absorbance over the course of growth. A lower overall

cell density will result in a smaller yield of overall biomass, which in the case of creating biofuel

will result in significantly lower biofuel yields. For the purposes of nutrient remediation, a lower

peak cell density will most likely result in a lesser percentage of nutrients used by the microalga.

All other IPD inoculation ranges were not significantly different from each other for peak cell

density/absorbance, indicating that if cultures are inoculated with greater than 2.96x105 cells (or

E680 of 0.100), cultures should reach approximately the same peak cell density over time. Results

also indicated that cultures starting at a lower cell density (0-2.96x105 cells/mL) had a significantly

longer lag growth phase which resulted in a later occurring log growth phase. The mean average

hours to log phase declined as cell density was increased but due to the standard deviations, the

Page 41: The Monoraphidium - Northern Illinois University

30

Figure 7 - Effects of Initial Population Density (IPD) in Cells/mL (via Hemocytometer Count) of

Monoraphidium sp. Dek19 Grown at 25°C in 1L Flasks on Peak Cell Density, Time to Log Phase,

and Time to Stationary Phase. Significantly different values are denoted by differing letters. Light

intensity of cultures was 48.3 ± 5.4 µmol photons m-2s-1 with an aeration rate of 1L

atmosphere/minute.

Page 42: The Monoraphidium - Northern Illinois University

31

only statistically significant difference was seen in cultures started at a lower cell density (0-

2.96x105 cells/mL). Cultures inoculated with a denser inoculum (1.963x106-3.296x106 cells/mL)

resulted in significantly lower lag phases which resulted in a significantly quicker onset of

stationary phase.

Table 1 – Effects of Initial Population Density (IPD) as Quantified by Extinction of Light at

680nm (E680) of Monoraphidium sp. Dek19 on Peak Cell Densities/mL measured at 680nm,

Time to Log Phase, and Time to Stationary Phase.

Data comes from the flask sets grown for Fig 6. Range of Cell Dry Weight was derived from an

equation by Adam Hage (E680=40.986(CDW) - 0.0611, R2=0.9942).

The mean average time to stationary phase showed no significant difference between

cultures inoculated with a cell density lower than 1.296 x 106 cells. There is a general decrease in

average hours to stationary phase as initial cell density is increased, but the trend is not statistically

significant. Although this is not significant, the general downwards trend may indicate that

cultures started with a higher cell density have longer log phases relative to the given culture’s

lifespan. This in turn may result in a quicker depletion of nutrients and a greater overall

accumulation of biomass (see Table 1) (Hage unpublished 2015).

Light Extinction at

680nm

Average Time to

Log Phase (Hours)

Average Time to

Stationary Phase

(Hours)

Average Peak Light

Extinction at 680nm

Range of Cell

Dry Weight/15ml

(g)

0.00-0.10 128.86 ± 37.98 391.60 ± 55.30 0.59 ± 0.14 0.010-0.016

0.10-0.15 73.43 ± 21.16 479.29 ± 149.56 1.16 ± 0.26 0.021-0.033

0.15-0.20 62.11 ± 34.20 470.60 ± 79.90 1.04 ± 0.24 0.018-0.030

0.20-0.40 48.17 ± 22.51 384.83 ± 63.92 1.24 ± 0.16 0.025-0.033

0.40-0.60 44.25 ± 17.56 266.00 ± 99.88 1.30 ± 0.16 0.026-0.034

0.60-1.00 24.33 ± 2.93 187.33 ± 38.12 1.33 ± 0.17 0.027-0.035

Page 43: The Monoraphidium - Northern Illinois University

32

Species Competition at Differing Temperature

This experiment was completed to determine any possible temperature-related

advantages that Monoraphidium sp. Dek19 may have over other species when grown at both

10°C and 22°C. According to a Microsoft Excel document obtained from the DSD, between the

years 2005- 2015 wastewater effluent existed at a temperature lower than 14°C for 47.64% of the

year with a low temperature of 8.65°C and a high temperature of 20.98°C. If Monoraphidium sp.

Dek19 were able to outcompete other alga at these colder midwestern temperatures, it is possible

that it would become an efficient source of biofeedstock for future biofuel production.

Species composition was determined through the use of manual cell counts for flasks at

10°C and 22°C. The average percentage of each species in a culture was calculated for both of

these flasks (Figure 8). Cultures initially consisted of three groups of phytoplankton:

Monoraphidium sp. Dek19, Chlorella sp., and Ulothrix sp. Inoculation of these open-pond-

growth flasks resulted in ~80% Chlorella sp., ~20% Monoraphidium sp. Dek19, and < 5%

Ulothrix sp. When grown at 22°C, Chlorella sp. was able to outcompete the other two

phytoplankton and quickly overtake the culture, reducing Monoraphidium sp. Dek19 to minimal

levels (<5%). However, when grown at 10°C, Monoraphidium sp. Dek19 was able to maintain

its initial culture composition. Growth curves indicate that the culture grown at 22°C was

entering stationary growth phase at the 500 hour mark, whereas the culture grown at 10°C was

still in the midst of its log growth phase at the 500 hour mark (Figure 9).

Page 44: The Monoraphidium - Northern Illinois University

33

Figure 8 - Effect of Growth Temperature on Algal Species Composition over Three Weeks for

Cultures Shown in Figure 9. Monoraphidium sp. Dek19 (●) are able to maintain their percentage

of the culture composition in the 10°C flask (B) but are outcompeted and pushed to extinction by

Chlorella sp. (■) when grown at 22°C (A). Ulothrix sp. (▲) was seen in minimal levels. Data

was obtained from cell counts of each species using a hemocytometer.

Figure 9 – Growth (E680) Curves of the Consortium of Species. The mixed 1L culture grown at

22°C (A) is entering stationary growth phase at the 500 hour mark, whereas the culture grown at

10°C (B) is still amidst log growth phase at the 500 hour mark.

A B

A B

Page 45: The Monoraphidium - Northern Illinois University

34

Even though Chlorella sp. may be more numerous at cultures grown at 22°C, the overall

volume they maintain is low. Monoraphidium sp. Dek19 (41-51x 2-3 µm) are much larger then

Chlorella sp. (2-10 µm in diameter) and are approximately 9.4 fold greater in volume. This

means that Monoraphidium sp. Dek19 at 20% of the total cells accounts for 70% of the biomass

and is largely responsible for extinction of light at 680nm (E680) in Figure 9B. This shows that

Monoraphidium sp. Dek19 could potentially be used as a biofuel feedstock when grown in

colder temperatures. This experiment helped identify a possible selection method for

Monoraphidium sp. Dek19. All future inoculum would be grown at 10°C to allow

Monoraphidium sp. Dek19 to outcompete other algae that do not grow well at this temperature.

Sucrose Density Gradients

The idea behind the sucrose density gradients was to determine a cheap and yet efficient

method of separating a particular algal species from the rest of the algal species in a consortium.

A naturally occurring consortium of species similar in composition to that seen in Figures 8 and

9 was centrifuged through a sucrose gradient. After centrifugation the microalgae formed two

distinct layers at the interface of the 2.0 and 2.5M layers of the gradient. The uppermost layer

was green in color while the thin bottom layer was yellow. Fractions (1mL) were carefully

pumped from the top of the sucrose gradient via a peristaltic pump and the previously stated

layers were observed under a low-power microscope at 430X. After observation, it was

determined that the top green layer was solely Monoraphidium sp. Dek19, whereas the bottom

Page 46: The Monoraphidium - Northern Illinois University

35

yellow layer was a mixed consortium of Chlorella sp., Ulothrix sp., and a minute portion of

Monoraphidium sp. Dek19 (Figure 10). The density of sucrose in water at 2.0M was determined

to be ~1.685g/mL and the 2.5M was determined to be 1.856g/mL. This indicates that the density

of Monoraphidium sp. Dek19 must be close to the density of 2.0M sucrose and the Chlorella sp.

must have a density closer to 2.5M. The top green layer of Monoraphidium sp. Dek19 cells could

then be washed with distilled water multiple times and used to inoculate a new culture. Cells

were washed via vortexing the samples in distilled water and centrifuging to remove the

supernatant. Scenedesmus sp. has been shown to have a higher concentration of lipids at

stationary phase when compared to the cells observed during log growth phase (Gardner et al.,

2012). This can potentially result in an increase in buoyancy of the algal cells due to a decrease

in their density associated with larger lipid and/or starch reserves (Smith & Manoylov, 2013).

Page 47: The Monoraphidium - Northern Illinois University

36

Figure 10 - A Mixed Microalga Culture Is Centrifuged Through a Sucrose Gradient (A). Two

distinct layers are formed at the interface of the 2.0 and 2.5M layers. The top green layer when

pumped out with a peristaltic pump is solely Monoraphidium sp. Dek19 (B). The smaller bottom

yellow layer is a mixed consortium of Chlorella sp., Ulothrix sp., and very few Monoraphidium

sp. Dek19 cells (C).

Page 48: The Monoraphidium - Northern Illinois University

37

Photosynthetic Rates of Monoraphidium sp. Dek19

Photosynthetic Oxygen Evolution per Total Chlorophyll

An oxygen electrode was used to measure differences in photosynthetic rates caused by

growth temperature, growth phase, and medium of a given culture. Two temperatures (10°C and

25°C) were observed as these are the annual minimum and maximum temperatures observed in

wastewater effluent at the DeKalb Sanitary District (DSD). The growth phase of the culture was

observed to detect any changes in photosynthetic competence as the cultures aged. It is possible

that older stationary phase cells may divert newly fixed carbon away from growth and into

harvestable lipids. Monoraphidium sp. Dek19 was grown as a monoculture in both final effluent

and post-primary filtration effluent in 10°C and ~25°C (ambient lab temperature). Photosynthetic

oxygen evolution rates of these cultures were observed during both the log growth and stationary

growth phases at the temperature at which growth occurred. Also, photosynthetic rates were

observed at 25°C for cultures grown at 10°C and at 10°C for cultures grown at 25°C.

In Figure 11A, cultures were grown in final effluent at ~25°C and tested at 25°C. Oxygen

evolution peaked at an average mean peak of 159.7µmols O2/mg chl/hr during log growth phase

at a light saturation level of 200 µmols photons m-2s-1. During stationary phase, max

photosynthetic rates decreased by an approximate factor of four and light saturated at a lower

light intensity. Samples from this culture were also tested at 10°C to observe how well the

Monoraphidium sp. Dek19 cells can react to quick changes in temperature (Figure 11B). As

expected, reduced rates of photosynthesis were observed in both log growth and stationary

growth phase. The light saturation levels remained unchanged.

Page 49: The Monoraphidium - Northern Illinois University

38

Figure 11 - Light Saturation Curve of Photosynthetic Oxygen Evolution per Weight of Total

Chlorophyll of Monoraphidium sp. Dek19 Grown in Final Effluent. Log phase and stationary

phase cultures were tested in these conditions. Log growth was defined as having an increase in

E680 of at least 0.100. Log phase cultures were examined between days four and ten at an E680

of 0.4-0.8. Stationary phase was defined as having an increase in E680 of less than 0.100 or a

consecutive decrease in E680 for at least two days. Stationary phase cultures were examined

after Day 10 at an E680 of 1.0-1.4. (A): Grown at an ambient lab temperature (~25°C) tested at

25°C. (B): Grown at an ambient lab temperature (~25°C) tested at 10°C. (C): Grown at 10°C

tested at 10°C. (D): Grown at 10°C tested at 25°C.

A

C

B

D

Page 50: The Monoraphidium - Northern Illinois University

39

Samples were then taken from cultures grown at 10°C and tested at 10°C (Figure 11C).

Oxygen evolution peaked at an average mean peak of 86.6 µmols O2/mg chl/hr during log

growth phase at a light saturation level of 60 µmols photons m-2s-1. During stationary phase, max

photosynthetic rates decreased by an approximate factor of four. This mimicked the changes

between log and stationary growth phases observed in Figures 11A and 11B. However, when

cultures grown at 10°C were tested at 25°C, the Monoraphidium sp. Dek19 cells responded in an

unexpected way. The expected result was that the algal cells would have seen a dramatic increase

in photosynthetic activity when provided warmer temperatures; instead, the cultures showed a

reduced rate of oxygen evolution than when tested at 10°C (Figure 11D).

Cultures grown in post-primary filtration effluent (Figure 12) exhibited similar rates of

photosynthesis as observed in final effluent (Figure 11). In Figure 12A, log growth cells

demonstrated an almost identical oxygen evolution when compared to the cultures in Figure 11A

(164.7 µmols O2/mg chl/hr in post-primary filtration effluent vs 159.7µmols O2/mg chl/hr in

final wastewater effluent). This trend continued for the comparison of Figure 11B to Figure 12B.

Figures 12C and 12D showed lowered rates of photosynthetic activity during log growth phase

when compared to Figures 11C and 11D. The opposite occurred when comparing the effect of

post-primary filtration effluent on the photosynthetic activity of stationary growth phase cells

grown at 10°C. Cultures that were tested at temperatures opposite of their growth temperature

(10°C tested at 25°C or 25°C tested at 10°C) showed reduced oxygen evolution rates than when

tested at the temperature of growth. Again, the results for post-primary filtration effluent (Figure

Page 51: The Monoraphidium - Northern Illinois University

40

Figure 12 - Light Saturation Curve of Photosynthetic Oxygen Evolution per Weight of Total

Chlorophyll of Monoraphidium sp. Dek19 Grown in Post-Primary Filtration Effluent. Log phase

and stationary phase cultures were tested in these conditions. Log and stationary growth phases

were equivalent to those described in Figure 11. (A): Grown at an ambient lab temperature

(~25°C) tested at 25°C. (B): Grown at an ambient lab temperature (~25°C) tested at 10°C. (C):

Grown at 10°C tested at 10°C. (D): Grown at 10°C tested at 25°C.

D C

B A

Page 52: The Monoraphidium - Northern Illinois University

41

12) showed many similarities to the results for Monoraphidium sp. Dek19 grown in final

wastewater effluent (Figure 11).

The comparison between photosynthetic rates observed as oxygen evolution per weight

of total chlorophyll indicated that there is a significant difference in photosynthetic rates between

log growth and stationary growth phase cultures. However, some of this difference is accounted

for due to self-shading effects on the microalgae during stationary growth phase. The comparison

between final wastewater effluent (Figure 11) and post-primary filtration effluent (Figure 12)

showed that there were no significant differences in photosynthetic rates between the media. At

each temperature and in each medium, light saturation occurred at 60-80 µmols photons m-2s-1

within stationary cultures, whereas light saturated at levels greater than 150 µmols photons m-2s-1

in cultures within log growth.

Photosynthetic Oxygen Evolution per Million Cells

Monoraphidium sp. Dek19 was grown as a monoculture in both final effluent and post-

primary filtration effluent in 10°C and ~25°C (ambient lab temperature). Photosynthetic oxygen

evolution rates of these cultures were observed during both the log growth and stationary growth

phases at the temperature at which growth occurred. Also, photosynthetic rates were observed at

25°C for cultures grown at 10°C and at 10°C for cultures grown at 25°C. Oxygen evolution was

observed on the basis of million cells because chlorophyll levels per cell can change throughout

growth of a culture. The algae were the same cultures observed in Figures 11 and 12, corrected

for cell count.

Page 53: The Monoraphidium - Northern Illinois University

42

In Figure 13A, a three-fold change was observed between stationary growth phase cells

and cells measured in log growth phase. In comparison, the change observed when measuring

per weight of chlorophyll reflected a four-fold difference (Figure 11A). This difference observed

is most likely due to a lower concentration of chlorophyll per cell in log growth when compared

to stationary growth phase cultures. Similar reactions to quick changes in temperature were

observed when photosynthetic rates were expressed per million cells. Figure 13B showed a two-

fold increase in photosynthetic rates between stationary and log phase growth cells. This was

compared to three- to four-fold increase in Figure 11B. A two-fold difference was also observed

in Figures 13C and 13D, compared to a three- to four-fold increase in Figures 11C and 11D. In

Figure 13, there is less variation in photosynthetic rates at lower light intensities. This indicates

that light saturation could be affected by self-shading of the cultures in stationary growth phase.

Light saturation rates were the same as those observed in Figure 11, as the samples were the

same.

Similar differences between oxygen evolution measured per weight of total chlorophyll

and per million cells were observed in post-primary filtration effluent (Figure 14). In each graph,

log and stationary growth were different by a smaller factor than what was observed in Figure

12. When observing changes to quick temperature changes, the trend was identical to Figure 12.

The results show that there are apparent differences in relative oxygen evolution rates for log

growth and stationary growth cells. When a comparison was made between total weight of

chlorophyll and per million cells, a noticeable difference was observed. This likely occurs due to

different chlorophyll levels per cell throughout the growth of a culture (Tables 2, 3, and 4). As a

culture ages, total chlorophyll levels per cell increase.

Page 54: The Monoraphidium - Northern Illinois University

43

Figure 13 - Light Saturation Curve of Photosynthetic Oxygen Evolution per Million Cells of

Monoraphidium sp. Dek19 Grown in Final Effluent. Log phase and stationary phase cultures

were tested in these conditions. Log phase and stationary phase cultures were tested in these

conditions. Log and stationary growth phases were equivalent to those described in Figure 11.

(A): Grown at an ambient lab temperature (~25°C) tested at 25°C. (B): Grown at an ambient lab

temperature (~25°C) tested at 10°C. (C): Grown at 10°C tested at 10°C. (D): Grown at 10°C

tested at 25°C.

D C

B A

Page 55: The Monoraphidium - Northern Illinois University

44

Figure 14 - Light Saturation Curve of Photosynthetic Oxygen Evolution per Million Cells of

Monoraphidium sp. Dek19 Grown in Post-primary Filtration Effluent Log phase and stationary

phase cultures were tested in these conditions. Log phase and stationary phase cultures were

tested in these conditions. Log and stationary growth phases were equivalent to those described

in Figure 11. (A): Grown at an ambient lab temperature (~25°C) tested at 25°C. (B): Grown at an

ambient lab temperature (~25°C) tested at 10°C. (C): Grown at 10°C tested at 10°C. (D): Grown

at 10°C tested at 25°C.

D C

B A

Page 56: The Monoraphidium - Northern Illinois University

45

Table 2 – Chlorophyll a:b Ratios of Monoraphidium sp. Dek19 for Cultures Grown at 10°C and

22°C in Final Effluent at 48.3 ± 5.4 µmol Photons m-2s-1. Log 10°C Stationary 10°C Log 22°C Stationary 22°C

Chlorophyll a (µg/mL) 3.43 ± 0.76 5.45 ± 0.14 6.93 ± 0.69 18.00 ± 2.24

Chlorophyll b (µg/mL) 1.99 ± 0.50 2.59 ± 0.19 4.13 ± 0.65 4.83 ± 0.77

Chlorophyll a:b ratio 1.76 ± 0.26 2.12 ± 0.22 1.70 ± 0.19 4.09 ± 1.33

Log growth was defined as having an increase in E680 of at least 0.100. Log phase cultures were

examined between days four and ten. Stationary phase was defined as having an increase in E680

of less than 0.100 or a consecutive decrease in E680 for at least two days. Stationary phase

cultures were examined after Day 10.

Table 3 – Chlorophyll a:b Ratios of Monoraphidium sp. Dek19 for Cultures Grown at 10°C and

22°C in Post-Primary Filtration Effluent at 48.3 ± 5.4 µmol Photons m-2s-1. Log 10°C Stationary 10°C Log 22°C Stationary 22°C

Chlorophyll a (µg/mL) 3.05 ± 0.19 5.36 ± 0.28 5.57 ± 0.73 10.11 ± 0.18

Chlorophyll b (µg/mL) 1.14 ± 0.17 2.22 ± 0.66 1.80 ± 0.17 2.69 ± 0.26

Chlorophyll a:b ratio 2.77 ± 0.62 2.99 ± 0.12 2.71 ± 0.67 3.76 ± 0.20

Table 4 – Chlorophyll per Cell Expressed as pg/Cell in Monoraphidium sp. Dek19 Cultures

Grown in Both Final and Post-Primary Filtration Effluent as Well as 10°C and 25°C. Growth Conditions Log Growth (pg/cell) Stationary Phase (pg/cell)

Final Grown at 25°C 1.71 ± 0.61 2.98 ± 1.53

Final Grown at 10°C 2.04 ± 0.02 3.08 ± 0.50

Post-Primary Filtration Grown at 25°C 1.39 ± 0.31 2.13 ± 0.93

Post- Primary Filtration Grown at 10°C 1.91 ± 0.23 2.88 ± 0.84

Total chlorophyll levels observed in the samples ranged from 1.75-13.25 µg/mL.

Page 57: The Monoraphidium - Northern Illinois University

46

Chlorophyll Levels in Monoraphidium sp. Dek19 During Photosynthesis Measurements

Samples were saved after the oxygen electrode photosynthesis experiments to identify

levels of chlorophyll a and chlorophyll b within each sample. It has been shown in

Chlamydomonas sp. and spinach that there is a relationship between the chlorophyll a/b ratio and

the photosystem II antenna size (Anderson & Melis, 1983; Perrine, Negi, & Sayre, 2012). A

small ratio of chlorophyll a to chlorophyll b indicates that the photosystem II antenna has

increased in size to maximize capture of light at high and low light intensities. A large ratio of

chlorophyll a to chlorophyll b indicates that photosystem I has reduced the size of its antenna to

prevent cell shading and increase the penetration of light into the water column (Perrine et al.,

2012).

Samples of Monoraphidium sp. Dek19 taken from cultures grown at 10°C overall had

lower levels of both chlorophyll a and chlorophyll b when compared to cultures grown at an

ambient lab temperature (22-25°C). However, chlorophyll a and b did still show an increase in

weight in the cultures grown at 10°C, as chlorophyll should be proportional to the number of

cells in culture samples. Cultures grown in final effluent had a greater total weight of chlorophyll

per mL when compared to cultures grown in post-primary filtration effluent. Although there was

a difference in overall weight, when observing the ratios of chlorophyll a to chlorophyll b

between temperature and growth phase there is a general trend that the ratio of chlorophyll a to

chlorophyll b does increase. When observing differences between temperatures of growth, algal

cultures grown at 22°C had considerably greater stationary growth chlorophyll a to b ratios than

cultures grown at 10°C (see Tables 2 and 3). After observing weight of chlorophyll per cell, it

Page 58: The Monoraphidium - Northern Illinois University

47

was noted that cultures grown at 10°C had on average more chlorophyll per cell at both log

growth and stationary growth phases. Algae grown in final effluent had on average more

chlorophyll per cell than the algae grown in post-primary filtration effluent at both log growth

and stationary growth phases. Between type of inoculum and temperature, stationary growth

algal cultures had more chlorophyll per cell than the same cultures tested at log growth phase

(see Table 4).

Samples were taken from a stationary phase culture grown at an ambient lab temperature

(22-25°C) to observe whether or not self-shading could be a factor in determining oxygen

evolution rates. To observe this, samples were taken from the same flask and diluted.

Photosynthetic rates were observed in duplicate measurements. Preliminary results show that

there is some effect of self-shading in stationary growth cultures. When under light saturation,

samples showed the highest levels of Oxygen evolution at the quarter dilution. There was not

much difference between the quarter and half dilutions, but there was a large difference in

photosynthetic rates of the quarter and half dilutions when compared to an undiluted sample

(Table 5).

Table 5 – Dilution of Stationary Growth Phase Cultures and the Effect on Light-Saturated

Oxygen Evolution Rates at 300 µmol Photons m2s-1. Dilution Factor Total Chlorophyll (µg/mL) E680 Oxygen Evolution (µmols O2/mg chl/hr)

1 15.44 1.620 51.33

0.5 7.75 0.810 96.18

0.25 4.20 0.405 105.56

Preliminary data of stationary growth phase culture of Monoraphidium sp. Dek19 diluted in half

and a quarter to show effects of self-shading. Data are means of two oxygen evolution

measurement taken at each dilution.

Page 59: The Monoraphidium - Northern Illinois University

48

Pollutant Remediation by Growth of Monoraphidium sp. Dek19

Initial Nutrient Levels

Final and post-primary filtration wastewater effluent taken from the DSD was tested for

the presence of eutrophication-causing pollutants (nutrients). Specifically, both types of

wastewater effluent were tested for nitrate, ammonium, and phosphate (Table 6). Similar levels

of phosphate were observed in both final and post-primary filtration effluent. However, the main

nitrogen source observed differed between the two wastewater effluents. In final effluent, a

majority of the nitrogen existed in the oxidized form (NO3-) with little to no ammonium (NH4

+).

In wastewater samples taken earlier in the treatment process, the majority of nitrogen existed in

the non-oxidized form of nitrogen (NH4+) with little to no nitrate (NO3

-) (Table 6).

Table 6 - Average Initial Nutrient Levels of Wastewater Effluent from the DeKalb Sanitary

District (DSD) at Time of Collection. Nutrient Final Effluent (mg/L) Post-primary filtration Effluent (mg/L)

Nitrate (NO3-) 37.37 ± 11.00 None Detected

Ammonium (NH4+) None Detected 5.06 ± 1.19

Phosphate (PO43-) 8.88 ± 2.70 7.67 ± 0.58

N=20 for final effluent, N=10 for post-primary filtration effluent. Starting nitrate levels ranged

from 18.04-52.61 mg/L. Ammonium levels ranged from 3.40-7.15 mg/L. Phosphate levels

ranged from 6.98-8.46mg/L.

Nutrient Uptake in Final Effluent and Post-Primary Filtration Effluent

Monoraphidium sp. Dek19 grown in final wastewater effluent was able to deplete both

nitrate and phosphate to minimal levels as an inverse relationship to algal growth within a two-

Page 60: The Monoraphidium - Northern Illinois University

49

week growing period. Previous research has also shown this to be true in mesocosm size 50-

gallon pools (Holbrook et al., 2014). Monoraphidium sp. Dek19 grown in a 1L flask can deplete

both nitrate and phosphate to minimal levels over the course of 13 days. The nitrate

concentration was reduced from ~40mg/L to approximately ~4 mg/L in this timespan (Figure

15). The phosphate concentration was reduced from ~14 mg/L to ~2 mg/L in the 13-day period

(Figure 16). Similar results were observed when growth of Monoraphidium sp. Dek19 occurred

in post-primary filtration effluent. Representative data showed that ammonium and phosphate

were depleted to minimal levels over the same time period at the same growth conditions. At

both temperatures, ammonium was reduced from ~7 mg/L to less than 1 mg/L (Figure 17).

Phosphate was reduced from ~11 mg/L to ~2 mg/L (Figure 18). The depletion in nutrients

coincides with the predicted inverse relationship between growth of a culture and nutrient

remediation.

In Figures 14 and 15, Monoraphidium sp. Dek19 grew to a higher E680 at 22°C than cells

grown at 10°C. At 10°C, cells grew slower than cells grown at 22°C. Cultures grown at 10°C

were only able to deplete two thirds of the phosphate in a 1L culture. In Figures 16 and 17,

Monoraphidium sp. Dek19 was able to grow to a higher E680 at 10°C than at 22°C in post-

primary filtration effluent after a slightly longer lag phase. Ammonium was depleted at both

temperatures, but the Monoraphidium sp. Dek19 cells did not deplete all of the phosphate even

when the cells grew to a higher E680 than cells grown at 22°C.

Page 61: The Monoraphidium - Northern Illinois University

50

Figure 15 – Representative Nutrient Removal Graph of Nitrate in Final Wastewater Effluent.

Growth of Monoraphidium sp. Dek19 (as E680 values) in final wastewater effluent at both 10°C

and 22°C (left axis). Corresponding levels of nitrate are also shown (right axis).

Figure 16 – Representative Nutrient Removal Graph of Phosphate in Final Wastewater Effluent.

Growth of Monoraphidium sp. Dek19 (as E680 values) in final wastewater effluent at both 10°C

and 22°C (left axis). Corresponding levels of phosphate are also shown (right axis).

0 50 100 150 200 250 300

05101520253035404550

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

0 24 48 72 96 120 144 168 192 216 240 264 288 312

Nit

rate

NO

3-(m

g/L)

Ab

sorb

ance

at

68

0n

m

Time (Hours)

Growth at 10°C Growth at 22°C 10°C Nitrate 22°C NItrate

0 50 100 150 200 250 300

0

3

6

9

12

15

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

0 24 48 72 96 120 144 168 192 216 240 264 288 312

PO

43

-(m

g/L)

Ab

sorb

ance

at

68

0n

m

Time (Hours)

Growth at 10°C Growth at 22°C 10°C Phosphate 22°C Phosphate

Page 62: The Monoraphidium - Northern Illinois University

51

Figure 17 – Representative Nutrient Removal Graph of Ammonium in Post-Primary Filtration

Wastewater Effluent. Growth of Monoraphidium sp. Dek19 (as E680 values) in post-primary

filtration effluent at both 10°C and 22°C (left axis). Corresponding levels of ammonium are also

shown (right axis).

0 50 100 150 200 250 300

0

1

2

3

4

5

6

7

8

9

10

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

0 24 48 72 96 120 144 168 192 216 240 264 288 312

NH

4+

(mg/

L)

Ab

sorb

ance

at

68

0n

m

Time (Hours)

Growth at 10°C Growth at 22°C 10°C Ammonium 22°C Ammonium

Page 63: The Monoraphidium - Northern Illinois University

52

Figure 18 – Representative Nutrient Removal Graph of Phosphate in Post-Primary Filtration

Wastewater Effluent. Growth of Monoraphidium sp. Dek19 (as E680 values) in post-primary

filtration effluent at both 10°C and 22°C (left axis). Corresponding levels of phosphate are also

shown (right axis).

0

2

4

6

8

10

12

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

0 24 48 72 96 120 144 168 192 216 240 264 288 312

PO

43

-(m

g/L)

Ab

sorb

ance

at

68

0n

m

Time (Hours)

Growth at 10°C Growth at 22°C 10°C Phosphate 22°C Phosphate

Page 64: The Monoraphidium - Northern Illinois University

53

Lipid Quantification

Nile Red Microscopy

Monoraphidium sp. Dek19 cells stained with Nile red dye were analyzed using a confocal

microscope during log and stationary growth phases for both final effluent (Figure 19) and for

post-primary filtration effluent (Figure 20). After using ImageJ to determine fluorescence of

individual cells at 568nm in the pictures taken, calculated total cell fluorescence was determined.

The results of this study indicated that there is no significant difference in cell fluorescence

between final and post-primary effluent at each growth stage. However, it was found that there is

a significant difference in fluorescence between log growth and stationary growth (Figure 21).

To reaffirm these statements, a Tukey HSD (honest significant difference) post hoc test with

α=0.05 was run to determine any statistical difference.

Fluorescence Spectrometry

Fluorescence spectrometry was utilized as an alternate method to determine fluorescence

of intracellular lipids stained by the Nile red dye. Output from the fluorescence

spectrophotometer indicated that the Monoraphidium sp. Dek19 cells read a higher absorbance

after sonication in a water bath than cells that were not sonicated. Cells were sonicated to disrupt

the cell wall to better measure intracellular lipid quantities. Results indicated a peak fluorescence

at a wavelength of 560nm (Figure 22).

Page 65: The Monoraphidium - Northern Illinois University

54

Figure 19 - Nile Red Microscopy Images of Monoraphidium sp. Dek19 Grown in Final

Wastewater Effluent. Images were taken with a confocal microscope. Image A depicts a bright

field image of Monoraphidium sp. Dek19. Image B depicts a fluorescent control image. Image C

depicts a bright field image of Monoraphidium sp. Dek19 during log growth. Image D depicts a

fluorescent image of lipids within Monoraphidium sp. Dek19 during log growth. Image E depicts

a bright field image of Monoraphidium sp. Dek19 during stationary growth. Image F depicts a

fluorescent image of lipids within Monoraphidium sp. Dek19 during stationary growth.

(Excitation= 530 nm, Emission= 568nm)

A

F E

D C

B

Page 66: The Monoraphidium - Northern Illinois University

55

Figure 20 - Nile Red Microscopy Images of Monoraphidium sp. Dek19 Grown in Post-Primary

Filtration Wastewater Effluent. Images were taken with a confocal microscope. Image A depicts

a bright field image of Monoraphidium sp. Dek19. Image B depicts a fluorescent control image.

Image C depicts a bright field image of Monoraphidium sp. Dek19 during log growth. Image D

depicts a fluorescent image of lipids within Monoraphidium sp. Dek19 during log growth. Image

E depicts a bright field image of Monoraphidium sp. Dek19 during stationary growth. Image F

depicts a fluorescent image of lipids within Monoraphidium sp. Dek19 during stationary growth.

(Excitation= 530 nm, Emission= 568nm)

A B

C D

E F

Page 67: The Monoraphidium - Northern Illinois University

56

Figure 21 - Calculated Total Cell Fluorescence (CTCF) of Monoraphidium sp. Dek19 Grown in

Both Final and Post-Primary Filtration Wastewater Effluent and Tested During Both Log and

Stationary Growth Phases. N=15 flasks for each treatment, significantly different values are

denoted by differing letters.

a

b

a

b

0

5000

10000

15000

20000

25000

Log Growth Stationary Growth

Flu

ore

scen

ce (

Arb

itra

ry u

nit

s)

Post-primary Filtration Final

Page 68: The Monoraphidium - Northern Illinois University

57

Figure 22 - Representative Data Showing an Estimation of Comparative Lipid Concentrations

via a Hitachi F2500 Fluorescence Spectrophotometer. Lipids in Monoraphidium sp. Dek19 cells

during stationary phase stained with Nile-red dye. Max fluorescence occurs at 560nm. Control

cells were Monoraphidium sp. Dek19 cells taken from the same culture and stained with Nile red

dye without sonication.

540 590 640 690 740 790

-20

80

180

280

380

480

580

680

-20

80

180

280

380

480

580

680

540 590 640 690 740 790

Flu

ore

scen

ce (

arb

itra

ry u

nit

s)

Wavelength (nm)

Control Nile Red Sonication Control Sonication Nile Red

Page 69: The Monoraphidium - Northern Illinois University

58

DISCUSSION

Summary

Monoraphidium sp. Dek19, a locally isolated green alga, was grown in both final and

post-primary filtration wastewater effluent taken from the DeKalb Sanitary District (DSD).

Monoraphidium sp. Dek19 was grown in both effluents to determine if the cells would be viable

earlier on in the wastewater treatment process. These media are very similar when it comes to

levels of phosphate, but they store nitrogen differently. Earlier on in the treatment process, the

nitrogen source is stored as ammonium, but as the wastewater is treated, nitrifying bacteria

convert nitrogen to its oxidized form of nitrate (see Table 6). Previous studies have shown that

Chlorella sp. was able to grow successfully in ammonium-based media (Hein et al., 1995;

Petrovič & Simonič, 2015; L. Wang et al., 2010). Not only could growth in an earlier state help

offset costs of pollutant remediation, but the algae could possibly be used as a source of oxygen

in activated sludge growth (Kiepper, 2013). This could potentially cut costs at wastewater

treatment facilities.

Monoraphidium sp. Dek19 was grown in both lab conditions (22-25°) and in an

environmental growth chamber (10°C) at a light intensity of approximately 50 µmol photons

m2s-1 in a 13:11 light/dark cycle. Monoraphidium sp. Dek19 was able to withstand competition

from other microalgae when grown at 10°C (see Figures 7 and 9). When grown at both lab

conditions and in the environment chamber, oxygen evolution rates were collected. It was

Page 70: The Monoraphidium - Northern Illinois University

59

determined that log phase growth had higher rates of oxygen evolution than stationary phase

growth and oxygen evolution rates were higher at 25°C than at 10°C (see Figures 10-13). When

grown in both post-primary filtration and final wastewater effluent, Monoraphidium sp. Dek19

was able to remove nitrate, phosphate, and ammonium in the wastewater to minimal levels (see

Figures 14-17). Samples taken during different growth phases of Monoraphidium sp. Dek19

were stained with Nile red dye and observations were made as to the quantity of lipids. It was

determined that algae in stationary growth phase had a significantly higher quantity of

intracellular lipids than samples tested in log growth phase (see Figures 18-20).

Effect of Initial Population Density (IPD) on Growth of Cultures

Initial population density (IPD) has been shown to exert an impact on the success

(measured in peak cell density/E680) of an algal culture (Chen et al., 2012; Yu et al., 2012). Yu

et al. (2012) have shown that cultures started at a high IPD (12x106 cells/mL) in 300mL of a BG-

11 medium have the capability to increase ~twenty-fold over a two-week span in

Monoraphidium sp. FXY-10. When grown in a 1L flask in wastewater effluent, Monoraphidium

sp. Dek19 (IPD of 1.963x106 cells/mL) are able to double the cell density per mL over the course

of ~10 days at an ambient lab temperature (22-25°C) (see Figure 9). This indicates that cultures

started at a higher IPD than used in this thesis may be able to achieve greater overall biomass

accumulation. Another study has shown similar results in cultures of Nannochloropsis sp. in

wastewater. When started at a high IPD, cultures were able to have a higher overall biomass than

cultures started at lower IPD; however, the algal cells had lower rates of biomass productivity.

This was thought to occur due to possible self-shading within the culture (Chen et al., 2012). An

Page 71: The Monoraphidium - Northern Illinois University

60

interesting concept to combat self-shading of cultures was incorporated into an experiment in

Brazil with the use of a continuous-growth culture setup of Nannochloropsis oculata. In this

setup, cell density is kept high, and as a culture grows, biomass is settled and removed for

conversion to biofuel (Olofsson et al., 2012). Through this process Olofsson et al. (2012) were

able to remove any negative factors resulting from self-shading and showed that equivalent

levels of lipids could be found year round in an outdoor setting in Portugal with temperatures of

~10-25°C. Although cultures started at higher IPD have been shown to accumulate greater

quantities of biomass, studies have shown that increased biomass does not necessarily mean

higher quantities of lipids. Chen et al. (2012) found that when starting at a moderate IPD (similar

to the IPD of 0.463-1.296x106 cells/mL), cultures have the highest rates of biomass productivity,

highest rates of nutrient uptake, and the most lipid per weight. When cultures of Chaetoceros

muelleri were started at low IPD, they were more susceptible to the negative effects of predation

and photoinhibition (Goksan, Durmaz, & Gokpinar, 2001). A culture that is started at an

extremely low IPD (see Figure 7 and Table 1) will reach stationary growth at a low cell density,

rendering the culture unsatisfactory in terms of biomass accumulation and pollutant remediation.

Initial population density is an important factor to consider when growing algae, as the biomass

production and lipid accumulation is severely affected by initial cell density. Further studies of

Monoraphidium sp. Dek19 should be completed to explore development of continuous-growth

cultures and lipid loading properties affected by IPD.

Species Competition

Competition among algal species was examined at different temperatures to determine

Page 72: The Monoraphidium - Northern Illinois University

61

the competitiveness of Monoraphidium sp. Dek19 when compared to other species. Cultures

were inoculated with an existing natural mixture of species that were allowed to grow in

autoclaved effluent in the greenhouse. The algae were determined to be Chlorella sp., Ulothrix

sp. and Monoraphidium sp. Dek19. Chlorella sp. has shown to be resilient to changes in growth

media and nutrient level but has an optimal growth of 30°C at 260 µmol photons m-2s-1

(Converti et al., 2009; L. Wang et al., 2010). Ulothrix sp., on the other hand, does not tolerate

change in conditions very well. Ulothrix sp. have been shown to demonstrate a competitive edge

at a pH of 3-5 (Niyogi, McKnight, & Lewis, 1999). Collected wastewater effluent from the

DeKalb Sanitary District tends to exist at a pH range of 8-10. Previous research completed in the

Holbrook lab has showed that Monoraphidium sp. Dek19 were capable of growth at 10°C and

light intensity of 40 µmol photons m-2s-1 (Holbrook et al., 2014). Because of this, cultures were

grown at low temperature and low light to determine if Monoraphidium sp. Dek19 displayed a

competitive advantage.

Cultures were determined to be ~80% Chlorella sp., ~20% Monoraphidium sp. Dek19,

and <5% Ulothrix sp. via manual cell counts (see Figures 7 and 9). At such a low cell density, it

was expected that Monoraphidium sp. Dek19 would not outcompete the other two species

because the IPD was too low. Indeed, that may have been the case at 22°C. Chlorella sp. was

able to overtake the culture at the ~80 hour mark. At an optimal growth temperature, Chlorella

sp. was able to grow at an exponential rate, which would account for the spike in growth on

Figure 9. However, when provided optimal growing conditions, Monoraphidium sp. Dek19 was

able to “hold its own” and maintain a ~20% species composition. Even though 20% does not

Page 73: The Monoraphidium - Northern Illinois University

62

seem like much of the culture, Monoraphidium sp. Dek19 takes up a majority of the biomass due

to the relatively large cell volume when compared to Chlorella sp. This may imply that

Monoraphidium sp. Dek19 would be able to outcompete other species if started at a higher IPD.

If Monoraphidium sp. Dek19 were to be used as a biofuel feedstock in the future, it may be

recommended to have it be an alternating algal crop based on season. In the winter when

wastewater effluent is approximately 14°C and the length of day is shorter (lower light on

average), Monoraphidium sp. Dek19 could be grown successfully as a monoculture. When the

temperatures increase in spring through summer, algae that favors the increased temperature and

light intensity could then be grown (Park, Whitney, Kozera, O'Leary, & McGinn, 2015).

Park et al. (2015) grew 11 different strains of algae in wastewater effluent taken from a

Nova Scotia wastewater treatment facility and determined whether species were more productive

(measured by lipid synthesis and nutrient removal) in 10°C or 22°C. Their group determined that

Chlorella vulgaris was capable of growth in both the cold and warmer temperatures. If

Monoraphidium sp. Dek19 were to be grown at a wastewater treatment facility, it would be a

viable option as a rotating algal crop because it is able to withstand competition from other algal

species. Another recommendation for large-scale algal growth may be to pre-load log phase

growth Monoraphidium sp. Dek19 with nutrients into a consortium of species. Without the initial

competition for nutrients, it is quite possible Monoraphidium sp. Dek19 would fare even better in

a consortium of species.

Sucrose Density Gradients

In a lab setting, it is sometimes necessary to isolate a desired algal species. This can be

Page 74: The Monoraphidium - Northern Illinois University

63

done through the use of sucrose density gradients (see Figure 10). This was a protocol developed

from scratch and arrived at after empirical trials. A discontinuous gradient was found to work the

best. This is a cost-efficient method to separate algae by their density. The benefits of using

sucrose as a density gradient is that it is cheap, with current prices in November 2015 of ~$40 for

500g from a scientific retailer and it is even cheaper if bought at a grocery store. The downsides

of using sucrose as a density gradient is the risk of bacteria taking over a young culture and

extracted algae must be washed multiple times in order to maintain viability of cells. Other ways

have been proven to separate algae but are more expensive options. One way to separate

microalgae is through the use of Percoll. Both Schwinghamer et al. (1991) and Whitelam et al.

(1983) used Percoll to separate algae out of a culture. The perks to the use of Percoll is that it

separates algae based on their density and algae can be easily extracted and used immediately to

inoculate a culture. The downside of Percoll is that it is expensive, with current prices in

November 2015 of ~$40 per 25mL. A third option of using colloidal silica Ludox-TM had been

used in the past to separate marine microalgae but has since been stopped as it is toxic to aquatic

organisms and therefore the algae could not be used to inoculate a new culture (de Jonge, 1979).

Therefore, sucrose gradients would allow separation and purification of larger volumes of

purified Monoraphidium sp. Dek19 to start enriched cultures.

Multiple attempts were made to find the correct sucrose density gradient for the purpose

of separating Monoraphidium sp. Dek19 from a consortium of algal species. First attempts at

sucrose gradients consisted of continuous gradients formed via a peristaltic pump and a gradient

former. This resulted in a scattered separation of algae throughout the column. Obviously, this

Page 75: The Monoraphidium - Northern Illinois University

64

would not be a useful way to separate algal species. The optimum gradient was eventually

discovered by trial and error. When a consortium of microalga were separated with sucrose

gradients, two distinct layers were formed at the interface of the 2.0M and 2.5M layers. The top

layer consisted of Monoraphidium sp. Dek19 and the bottom layer was a mixture of species

consisting of ~95% Chlorella sp. and the remainder a combination of Ulothrix sp. and

Monoraphidium sp. Dek19. This may have occurred because the separated cultures were in

stationary phase. If Monoraphidium sp. Dek19 showed an increase in lipids, this could result in a

potential increase in buoyancy of the algal cells (Smith & Manoylov, 2013). For example, Smith

and Manoylov (2013) showed that when cultures of Thalassisosira sp. 1 reached stationary

phase, cells became buoyant to move up in the water column. This was attributed to the larger

size of the cell and low surface-to-volume ratio. This may be what is occurring in the sucrose

gradients, as Monoraphidium sp. Dek19 cells have a greater volume than Chlorella sp. cells.

Chlorella sp. have been shown to exhibit a volume of 33µm3 (Reynolds, 1984). The approximate

volume of Monoraphidium sp. Dek19 is larger, exhibiting an approximate volume of 128-

360µm3. This increase in volume may be the reason that the layer of Monoraphidium sp. Dek19

was located above that of Chlorella sp. Another reason that the layer of Monoraphidium sp.

Dek19 was located above the layer of Chlorella sp. may be due to the lipid content of each alga.

As lipids are less dense than water, it would be assumed that the algae with the higher

concentration of lipids should move through the sucrose gradient the least distance. This could

indicate that Monoraphidium sp. Dek19 has a higher lipid content per volume than Chlorella sp.

Page 76: The Monoraphidium - Northern Illinois University

65

Photosynthetic rates of Monoraphidium sp. Dek19 were observed in both final and post-

primary filtration wastewater effluent. This was of interest for a few reasons: 1) microalgae may

have the ability to offset electrical costs involved with oxygenation of activated sludge.

Activated sludge is the treatment step in which wastewater is combined with nitrifying bacteria

that are kept in suspension through an input of air. The necessary oxygenation to obtain optimal

production from microorganisms within the activated sludge is 1.0-3.0 mg/L ("Introduction to

Activated Sludge Study Guide," 2010). 2) Light saturation levels could be used to determine

maximum light intensity required to maintain optimal oxygenation rate in the different media

and at different stages of growth. 3) Optimal growing conditions at both 10°C and 22°C could be

determined in order to make a case for Monoraphidium sp. Dek19 as a rotating algal crop in

colder climates.

Monoraphidium sp. Dek19 showed equivalent rates of photosynthesis in both post-

primary filtration and final effluent, indicating that there are no noticeable inhibitory effects of

different nitrogen species on the photosynthetic performance of the microalga. This shows that

both ammonium and nitrate can be used as a nitrogen source for Monoraphidium sp. Dek19.

Peak photosynthetic rates of Monoraphidium sp. Dek19 were observed in cultures grown at an

ambient lab temperature (22-25°C; 159.7µmols O2/mg chl/hr) during log phase growth (see

Figures 10 and 11). As the molecular weight of O2 is equal to 32.0g, 1µmol of O2 evolved is

equal to 0.032mg of O2 evolved. This indicates that at peak rates ~5.11mg of O2 will be evolved

per mg of total chlorophyll per hour. Total chlorophyll per mL ranged between 1.75-13.25

µg/mL; therefore, O2 levels evolved at peak photosynthetic rates would be in a range of 8.94 –

67.71 mg/L/hr, well above the 1.0-3.0 mg/L of constant aeration needed for the activated sludge

Photosynthetic Rates of Monoraphidium sp. Dek19

Page 77: The Monoraphidium - Northern Illinois University

66

treatment process. The lowest rates observed were in cultures grown at 10°C (18.4 µmols O2/mg

chl/hr) during stationary phase growth. This makes sense as the metabolic processes (in this case

photosynthesis) are slowed down by a factor of ~2-3 due to the Q10 value (Davison, 1991). Even

at this low rate of O2 evolution, O2 levels would be within the necessary constant aeration of 1.0-

3.0 mg/L oxygenation at an O2 evolution rate of 1.03-7.80 mg/L/hr. Although the rates are lower

at stationary growth, this could simply be due to self-shading of the cultures. In conditions such

as stationary growth phase, algae will self-shade and therefore provide skewed results for

photosynthetic rates. As seen in Table 5, a dilution experiment was set up to provide an idea to

the effect of shading on decreased photosynthetic rates. It was found that Monoraphidium sp.

Dek19 cultures at stationary phase may be subject to the diminished oxygen evolution rates

observed with self-shading. Other studies have shown that when the media becomes nitrogen

depleted, Phaeodactylum tricornutum can divert photosynthetically fixed carbon toward lipid

production rather than towards structural carbohydrates (Levitan et al., 2015).

Light saturation occurred at 200 µmol photons m-2s-1 for log phase growth cultures grown

at an ambient lab temperature (22-25°C) and at 60-100 µmol photons m-2s-1 for log phase growth

cultures grown at 10°C. These values were decreased to 40-60 µmol photons m-2s-1 for

stationary growth cultures grown at an ambient lab temperature (22-25°C) and 80 µmol photons

m-2s-1 for stationary growth cultures grown at 10°C. This indicates the need for a lower light

intensity as cultures age, perhaps because they adapt to self-shading. It has been shown that

when irradiance is greater than the level of light saturation, there are negative effects on growth

of the algae (Powles, 1984; B. Smith et al., 1990; Vasilikiotis & Melis, 1994). Powles (1984)

observed a decreased rate in photosystem II activity and oxygen evolution rates due to damaged

Page 78: The Monoraphidium - Northern Illinois University

67

photosystem II reaction centers in the thylakoids when the light intensity was greater than the

level of light saturation. Likewise, Smith et al. (1990) observed photoinhibition of algae at high

light intensities which resulted in a shorter chlorophyll antenna. Also, Vasilikiotis and Melis

(1994) found that up to 80% of photosystem II would be damaged by a high light intensity.

These studies show the negative effects of an overabundance of irradiance on algae, but in

deeper water this is less likely to occur. It is easier to see the negative effects of excess light

when testing in a 1mL-capacity oxygen electrode. This is an aspect of the project that would

need to be further explored if grown in larger batches. Monoraphidium sp. Dek19 exhibit similar

photosynthetic rates to other microalgae.

In an older study completed by Lloyd et al. (1977), photosynthetic rates of Chlorella sp.,

Anabaena sp., and Navicula sp. grown at 25°C were observed. They found Chlorella sp. to have

a photosynthetic oxygen evolution rate of 100 µmols O2/mg chl/hr at a light saturation of 200

µmol photons m-2s-1. Anabaena sp. was found to have a photosynthetic oxygen evolution rate of

170 µmols O2/mg chl/hr at a light saturation of 240 µmol photons m-2s-1. Navicula sp. was found

to have a photosynthetic oxygen evolution rate of 200 µmols O2/mg chl/hr at a light saturation of

200 µmol photons m-2s-1 (Lloyd, Canvin, & Culver, 1977). As Monoraphidium sp. Dek19 has

similar rates of photosynthesis at light intensities equivalent to the Lloyd et al. (1977) study, it

could possibly be used as a native algal crop for wastewater treatment facilities in the Midwest.

Samples taken from the photosynthetic experiments were examined for chlorophyll a and

chlorophyll b (see Tables 2 and 3). Chlorophyll a:b ratios were found to increase over the course

of growth of a culture. This may be related to a decrease in total nitrogen available to the algae.

Page 79: The Monoraphidium - Northern Illinois University

68

A study completed by Kitajima and Hogan (2003) showed increased ratios of chlorophyll a to

chlorophyll b when nitrogen was depleted in woody seedlings. Although algae and woody

seedlings are quite different, I believe this relationship holds true for Monoraphidium sp. Dek19

as sources of nitrogen are depleted during stationary growth (see Figures 14 and 16). Others

suggest that larger chlorophyll a:b ratios should result in a more photosynthetically efficient alga.

When Perrine et al. (2012) genetically engineered Chlamydomonas reinhardtii to have reduced

chlorophyll b levels, they saw a two-fold increase in the photosynthetic rate of the algae as well

as an increase in overall growth rates at light saturation due to a more efficient coupling of

photon capture in the light-harvesting complexes and electron transfer (Perrine et al., 2012). As

chlorophyll b levels remain low when grown in 10°C and 22°C, this may suggest that

Monoraphidium sp. Dek19 are photosynthetically adapted to perform at a colder temperature

such as 10°C because carbon metabolism is slower at 10°C than 25°C and not so much ATP is

used, causing a lower demand on electron transport reactions. This would also corroborate data

from nutrient remediation, as growth curves are similar (see Figures 14-16).

Phycoremediation

Monoraphidium sp. Dek19 has shown the capability of depleting concentrations of nitrate,

ammonium, and phosphate in both final and post-primary filtration wastewater effluent to minimal

levels over the course of approximately two weeks of growth (see Figures 13-16). This depletion

was observed at both 10°C and ambient lab temperature (22-25°C). Similar rates of nutrient

remediation were observed at both temperatures, which may indicate Monoraphidium sp. Dek19

could be a viable candidate for growth in the colder midwestern winters.

Page 80: The Monoraphidium - Northern Illinois University

69

Nitrogen and phosphorous sources need to be removed from wastewater effluent before

being discharged into waterways. This is a major concern not only for the health of humans but

also for the health of the waterway itself. High levels of nitrates, phosphates, and ammonium have

been shown to cause eutrophication of waterways (Smolders, Lucassen, Bobbink, Roelofs, &

Lamers, 2010; Turner & Rabalais, 1994). Eutrophication leads to a decreased water quality that

can have negative effects on the environment, such as hypoxia of water and fish kills (Borsuk et

al., 2004). Monoraphidium sp. Dek19 have the possibility to aid in nutrient remediation at

wastewater treatment facilities. Some algae have been reported to accumulate polyphosphate

granules. However, in the case of Monoraphidium sp. Dek19, the ultra-structure of the cell would

need to be examined to confirm that this also happens in this species. The treatment of wastewater

is also currently being studied using Scenedesmus sp. AMDD and Chlorella vulgaris (Dickinson,

Whitney, & McGinn, 2013; Fathi, Azooz, & Al-Fredan, 2013). As the EPA sets stricter limits on

effluent levels of nitrogen and phosphorus, the expenses of treating wastewater increase. For some

wastewater treatment facilities, new infrastructure must be built or older equipment must be

retrofitted to meet the new demands. For example, if a wastewater treatment facility chose to

upgrade its facilities to the modified Bardenpho (four-stage) process, it could cost the facility up

to $1,293,524 in construction costs with an operating cost of $162,169 per year ("Biological

Nutrient Removal Processes and Costs," 2007). This would meet the effluent levels for both

nitrogen and phosphorous, but it comes at a steep financial cost. With increasingly stringent EPA

legislation, the growth of algae as an additional step may lessen the chance of fines as well as

provide a lower operating cost for municipalities in the future. To cut costs even further, the

Page 81: The Monoraphidium - Northern Illinois University

70

biomass from the Monoraphidium sp. Dek19 could be used or sold for the purposes of creating

biofuel.

Lipid Quantification

Monoraphidium sp. Dek19 was grown in post-primary filtration (see Figure 20) and final

(see Figure 19) wastewater effluent at an ambient lab temperature (22-25°C). Samples were taken

during log growth and stationary growth phases and were stained with Nile red dye to determine

the presence of lipids. Fluorescence images taken on a confocal microscope were then analyzed in

ImageJ to determine calculated total cell fluorescence (CTCF). No significant difference was

found between post-primary filtration and final wastewater effluent (see Figure 21). However, a

statistically significant difference was found between growth stages. Fluorescence of lipids was

much stronger in stationary growth phase than in log growth phase, indicating that lipids are stored

at an increased rate during stationary growth. Log growth was defined as having an increase in

E680 of at least 0.100. Log phase cultures were examined between Days 4 and 10 at an E680 of

0.4-0.8. Stationary phase was defined as having an increase in E680 of less than 0.100 or a

consecutive decrease in E680 for at least two days. Days after inoculation depended upon each

culture, but stationary-phase flasks were usually examined between 14-21 days post-inoculation.

Studies have shown that when nutrients become limited, there is a change in the allocation

of carbon within green algae from structural carbohydrates used in growth towards neutral lipids

(Hu, 2004). As nutrients are depleted to minimal levels during stationary growth phase (see Figures

14-17), this would suggest that Monoraphidium sp. Dek19 diverts carbon towards lipid production

rather than structural carbohydrates at the end of a culture’s growth because of the exponential

Page 82: The Monoraphidium - Northern Illinois University

71

increase in lipids forming compared to the rate of new cell growth. Converti et al. (2009) and Ren

et al. (2013) have shown that when nitrogen sources are limited, Nannochloropsis sp. (Converti et

al., 2009) and Scenedesmus sp. (Gardner et al., 2012; Ren, Liu, Ma, Zhao, & Ren, 2013) show an

increase in lipid content. Another study shows that when sources of phosphorous are limited,

Ankistrodesmus falcatus shows an increase in lipid content (Kilham, Kreeger, Goulden, & Lynn,

1997). When compared to these studies, there is good evidence to suggest Monoraphidium sp.

Dek19 behaves in a similar manner. If Monoraphidium sp. Dek19 were to be used as a feedstock

for the purposes of generating biofuel, information as to when to harvest the algal cells would be

useful. The optimum yield of lipids would occur if the algal cells are harvested during early

stationary phase rather than log growth. Future experiments should observe the effect of

temperature on lipid content. If an algal species is cold tolerant, they are less likely to suffer

diminished lipid levels at colder temperatures (Park et al., 2015).

Page 83: The Monoraphidium - Northern Illinois University

72

CONCLUSIONS

Monoraphidium sp. Dek19 may be a good alga species for phycoremediation of polluting

levels of nitrogen and phosphorous at wastewater treatment facilities. Monoraphidium sp.

Dek19 was able to remove excess sources of nitrogen (ammonium and nitrate) and

phosphorous (phosphate) from both media.

Another implementation of the microalga would be as a potential biofuel feedstock. Algal

cells were found to be capable of successful growth in both post-primary filtration and final

wastewater effluent. Municipal wastewater is a free nutrient source that is currently unused,

resulting in both the financial benefit of low-cost growth media and the environmental

benefit of mitigating levels of pollutants released into the environment.

Monoraphidium sp. Dek19 was shown to be able to compete with a consortium of species

in wastewater at 10°C. Cells may have been able to compete with the other alga at 22°C if

started at a higher IPD.

Monoraphidium sp. Dek19 was found to have comparable photosynthetic rates to model

green microalgae like Chlorella. The former microalga might be used as a source of

oxygenation in the expensive step of activated sludge treatment. Even if Monoraphidium

sp. Dek19 were only grown later in final wastewater effluent, algae could be harvested and

the oxygenated water could be recycled back into the activated sludge treatment step.

Page 84: The Monoraphidium - Northern Illinois University

73

Chlorophyll a:b ratios suggest that Monoraphidium sp. Dek19 can acclimate to growth in

colder environments. If Monoraphidium sp. Dek19 were grown in a wastewater treatment

facility, it may be utilized as an alternating algal crop in which Monoraphidium sp. Dek19

is cultured in the cold midwestern winter months and a warm-climate adapted alga is grown

in the warm midwestern summer months.

Sucrose density gradients can be used as an inexpensive laboratory method to separate

Monoraphidium sp. Dek19 from other microalgae.

When nutrient depletion occurs, Monoraphidium sp. Dek19 increase production of lipids,

indicating peak harvesting during or after stationary growth.

Page 85: The Monoraphidium - Northern Illinois University

74

REFERENCES

Activated Sludge. (2015). from http://dekalbsd.com/the-sewer-plant/activated-sludge/

Agricultural Marketing Resource Center. (2014). Ethanol Usage Projections & Corn Balance

Sheet. from http://www.extension.iastate.edu/agdm/crops/outlook/cornbalancesheet.pdf

Amand, A. S. (2012). Identification, Ecology and Control of Nuisance Freshwater Algae.

Anderson, J. M., & Melis, A. (1983). Localization of different photosystems in separate regions

of chloroplast membranes. Proc Natl Acad Sci U S A, 80(3), 745-749.

Arora, A., & Saxena, S. (2005). Cultivation of Azolla microphylla biomass on secondary-treated

Delhi municipal effluents. Biomass and Bioenergy, 29(1), 60-64. doi:

http://dx.doi.org/10.1016/j.biombioe.2005.02.002

Aschengrau, A., Zierler, S., & Cohen, A. (1989). Quality of community drinking water and the

occurrence of spontaneous abortion. Arch Environ Health, 44(5), 283-290. doi:

10.1080/00039896.1989.9935895

Barsanti, L., & Gualtieri, P. (2014). Algae: Anatomy, Biochemistry, and Biotechnology (2 ed.).

Boca Raton, FL: CRC Press.

Biological Nutrient Removal Processes and Costs. (2007). from

http://www2.epa.gov/sites/production/files/documents/criteria_nutrient_bioremoval.pdf

Bono Jr, M. S., Ahner, B. A., & Kirby, B. J. (2013). Detection of algal lipid accumulation due to

nitrogen limitation via dielectric spectroscopy of Chlamydomonas reinhardtii

suspensions in a coaxial transmission line sample cell. Bioresource Technology, 143,

623-631. doi: http://dx.doi.org/10.1016/j.biortech.2013.06.040

Page 86: The Monoraphidium - Northern Illinois University

75

Borsuk, M. E., Stow, C. A., & Reckhow, K. H. (2004). A Bayesian network of eutrophication

models for synthesis, prediction, and uncertainty analysis. Ecological Modelling, 173(2–

3), 219-239. doi: http://dx.doi.org/10.1016/j.ecolmodel.2003.08.020

BP Statistical Review of World Energy June 2015. (2015). from

https://www.bp.com/content/dam/bp/pdf/energy-economics/statistical-review-2015/bp-

statistical-review-of-world-energy-2015-full-report.pdf

Calvin, M., & Massini, P. (1952). The path of carbon in photosynthesis. XX. The steady state.

Experientia, 8(12), 445-457.

Case Studies on Implementing Low-Cost Modifications to Improve Nutrient Reduction at

Wastewater Treatment Plants. (2015). from

http://www2.epa.gov/sites/production/files/2015-

08/documents/case_studies_on_implementing_low-

cost_modification_to_improve_potw_nutrient_reduction-combined_508_-_august.pdf

Castro, G. R., Larson, B. K., Panilaitis, B., & Kaplan, D. L. (2005). Emulsan quantitation by Nile

red quenching fluorescence assay. Appl Microbiol Biotechnol, 67(6), 767-770. doi:

10.1007/s00253-004-1849-2

Cedergren, M. I., Selbing, A. J., Lofman, O., & Kallen, B. A. (2002). Chlorination byproducts

and nitrate in drinking water and risk for congenital cardiac defects. Environ Res, 89(2),

124-130.

Chen, Y., Wang, J., Liu, T., & Gao, L. (2012). Effects of initial population density (IPD) on

growth and lipid composition of Nannochloropsis sp. Journal of Applied Phycology,

24(6), 1623-1627. doi: 10.1007/s10811-012-9825-1

Chinnasamy, S., Bhatnagar, A., Claxton, R., & Das, K. C. (2010). Biomass and bioenergy

production potential of microalgae consortium in open and closed bioreactors using

untreated carpet industry effluent as growth medium. Bioresour Technol, 101(17), 6751-

6760. doi: 10.1016/j.biortech.2010.03.094

Chisti, Y. (2007). Biodiesel from microalgae. Biotechnol Adv, 25(3), 294-306. doi:

10.1016/j.biotechadv.2007.02.001

Page 87: The Monoraphidium - Northern Illinois University

76

The Colorado River Commission of Nevada. (2002). World Fossil Fuel Reserves and Projected

Depletion. from http://crc.nv.gov/docs/world%20fossil%20reserves.pdf

Construction General Permit - Standard Permit Conditions. (2012). from

http://water.epa.gov/polwaste/npdes/stormwater/upload/cgp2012_finalpermit.pdf

Converti, A., Casazza, A., Ortiz, E., Perego, P., & Borghi, M. D. (2009). Effect of temperature

and nitrogen concentration on the growth and lipid content of Nannochloropsis oculata

and Chlorella vulgaris for biodiesel production. Chemical Engineering and Processing:

Process Intensification, 48(6), 1146-1151.

Crude Oil. (2015). WTI (NYMEX) Price. from http://www.nasdaq.com/markets

/crude-oil.aspx?timeframe=10y

Davison, I. R. (1991). Environmental effects on algal photosynthesis: Temperature. Journal of

Phycology, 27(1), 2-8. doi: 10.1111/j.0022-3646.1991.00002.x

de-Bashan, L. E., & Bashan, Y. (2010). Immobilized microalgae for removing pollutants:

Review of practical aspects. Bioresour Technol, 101(6), 1611-1627. doi:

10.1016/j.biortech.2009.09.043

De Bhowmick, G., Koduru, L., & Sen, R. (2015). Metabolic pathway engineering towards

enhancing microalgal lipid biosynthesis for biofuel application—A review. Renewable

and Sustainable Energy Reviews, 50, 1239-1253. doi:

http://dx.doi.org/10.1016/j.rser.2015.04.131

de Gorter, H., & Just, D. (2010). The social costs and benefits of biofuels: The intersection of

environmental, energy and agricultural policy. Applied Economic Perspectives and

Policy, 32(1), 4-32.

de Jonge, V. N. (1979). Quantitative separation of benthic diatoms from sediments using density

gradient centrifugation in the colloidal silica Ludox-TM. Marine Biology, 51(3), 267-278.

doi: 10.1007/BF00386807

Page 88: The Monoraphidium - Northern Illinois University

77

Dickinson, K. E., Whitney, C. G., & McGinn, P. J. (2013). Nutrient remediation rates in

municipal wastewater and their effect on biochemical composition of the microalga

Scenedesmus sp. AMDD. Algal Research, 2(2), 127-134. doi:

http://dx.doi.org/10.1016/j.algal.2013.01.009

Disinfection. (2015). from http://dekalbsd.com/the-sewer-plant/disinfection/

Dorsch, M. M., Scragg, R. K., McMichael, A. J., Baghurst, P. A., & Dyer, K. F. (1984).

Congenital malformations and maternal drinking water supply in rural South Australia: A

case-control study. Am J Epidemiol, 119(4), 473-486.

Fathi, A., Azooz, M., & Al-Fredan, M. (2013). Phycoremediation and the potential of sustainable

algal biofuel production using wastewaster. American Journal of Applied Sciences, 10(2),

189-194. doi: 10.3844/ajassp.2013.189.194

Fawley, M., Dean, M., Dimmer, S., & Fawley, K. (2005). Evaluating the morphospecies concept

in the Selenastraceae (Chlorophyceae, Chlorophyta). Journal of Phycology, 42, 142-154.

Fitzpatrick, M. (2014). Measuring cell fluorescence using ImageJ. from

http://theolb.readthedocs.org/en/latest/imaging/measuring-cell-fluorescence-using-

imagej.html

Friedmann, E. I., & Kiebler, A. P. (1980). Nitrogen economy of endolithic microbial

communities in hot and cold desert. Microbial Ecology, 6, 95-108.

Gardner, R., Cooksey, K., Mus, F., Macur, R., Moll, K., Eustance, E., . . . Peyton, B. (2012). Use

of sodium bicarbonate to stimulate triacylglycerol accumulation in the chlorophyte

Scenedesmus sp. and the diatom Phaeodactylum tricornutum. Journal of Applied

Phycology, 24(5), 1311-1320. doi: 10.1007/s10811-011-9782-0

Godos, I. d., Vargas, V. A., Blanco, S., González, M. C. G., Soto, R., García-Encina, P. A., . . .

Muñoz, R. (2010). A comparative evaluation of microalgae for the degradation of

piggery wastewater under photosynthetic oxygenation. Bioresource Technology, 101(14),

5150-5158. doi: http://dx.doi.org/10.1016/j.biortech.2010.02.010

Page 89: The Monoraphidium - Northern Illinois University

78

Goksan, T., Durmaz, Y., & Gokpinar, S. (2001). Effects of light path lengths and initial culture

density on the cultivation of Chaetoceros muelleri. Aquaculture, 217(1), 431-436.

Grant, W., Steel, G., & Isiorho, S. (1996). Spontaneous abortions possibly related to ingestion of

nitrate contaminated well water. MMWR Morb Mortal Wkly, 45, 569-572.

Griffiths, M. J., Garcin, C., van Hille, R. P., & Harrison, S. T. L. (2011). Interference by pigment

in the estimation of microalgal biomass concentration by optical density. Journal of

Microbiological Methods, 85(2), 119-123. doi:

http://dx.doi.org/10.1016/j.mimet.2011.02.005

Hein, M, Pedersen, M., & Sand-Jensen, K. (1995). Size-dependent nitrogen uptake in micro- and

macroalgae. Marine Ecology Progress Series, 118, 247-253.

Hoffmann, J. (1998). Wastewater treatment with suspended and nonsuspended algae. Journal of

Phycology, 34, 757-763.

Holbrook, G. P., Davidson, Z., Tatara, R. A., Ziemer, N. L., Rosentrater, K. A., & Grayburn, W.

S. (2014). Use of the microalga Monoraphidium sp. grown in wastewater as a feedstock

for biodiesel: Cultivation and fuel characteristics. Applied Energy, 131, 386-393.

Hu, Q., Sommerfeld, M., Jarvis, E., Ghirardi, M., Posewitz, M., Seibert, M., & Darzins, A.

(2008). Microalgal triacylglycerols as feedstocks for biofuel production: Perspectives and

advances. Plant J, 54(4), 621-639. doi: 10.1111/j.1365-313X.2008.03492.x

Hu, Q. (2004). Environmental effects on cell composition. Handbook of Microalgal Culture:

Biotechnology and Applied Phycology, 83-93. Blackwell Publishing Ltd.

Introduction to Activated Sludge Study Guide. (2010). from

http://dnr.wi.gov/regulations/opcert/documents/wwsgactsludgeintro.pdf

Kiepper, B. (2013). Microalgae Utilization in Wastewater Treatment. In The University of

Georgia Cooperative Extension (Ed.). Bulletin 1419.

Page 90: The Monoraphidium - Northern Illinois University

79

Kilham, S., Kreeger, D., Goulden, C., & Lynn, S. (1997). Effects of nutrient limitation on

biochemical constituents of Ankistrodesmus falcatus. Freshwater Biology, 38(3), 591-

596. doi: 10.1046/j.1365-2427.1997.00231.x

Kitajima, K., & Hogan, K. P. (2003). Increases of chlorophyll a/b ratios during acclimation of

tropical woody seedlings to nitrogen limitation and high light. Plant Cell Environ, 26(6),

857-865.

Krienitz, L., Ustinova, I., Friedl, T., & Huss, V. A. R. (2001). Traditional generic concepts

versus 18S rRNA gene phylogeny in the green algal family Selenastraceae

(Chlorophyceae, Chlorophyta). Journal of Phycology, 37(5), 852-865. doi:

10.1046/j.1529-8817.2001.01004.x

Levitan, O., Dinamarca, J., Zelzion, E., Lun, D. S., Guerra, L. T., Kim, M. K., . . . Falkowski, P.

G. (2015). Remodeling of intermediate metabolism in the diatom Phaeodactylum

tricornutum under nitrogen stress. Proc Natl Acad Sci U S A, 112(2), 412-417. doi:

10.1073/pnas.1419818112

Lewis, L., & McCourt, R. (2004). Green algae and the origin of land plants. American Journal of

Botany, 91(10), 1535-1556.

Lloyd, N. D., Canvin, D. T., & Culver, D. A. (1977). Photosynthesis and photorespiration in

algae. Plant Physiol, 59(5), 936-940.

Mallick, N. (2002). Biotechnological potential of immobilized algae for wastewater N, P and

metal removal: A review. Biometals, 15(4), 377-390. doi: 10.1023/A:1020238520948

Mezzomo, N., Saggiorato, A. G., Siebert, R., Tatsch, P. O., Lago, M. C., Hemkemeier, M., . . .

Colla, L. M. (2010). Cultivation of microalgae Spirulina platensis (Arthrospira platensis)

from biological treatment of swine wastewater. Food Science and Technology

(Campinas), 30, 173-178.

Mulbry, W., Kondrad, S., Pizarro, C., & Kebede-Westhead, E. (2008). Treatment of dairy

manure effluent using freshwater algae: Algal productivity and recovery of manure

nutrients using pilot-scale algal turf scrubbers. Bioresource Technology, 99(17), 8137-

8142. doi: http://dx.doi.org/10.1016/j.biortech.2008.03.073

Page 91: The Monoraphidium - Northern Illinois University

80

Niyogi, D. K., McKnight, D. M., & Lewis, W. M. (1999). Influences of water and substrate

quality for periphyton in a montane stream affected by acid mine drainage. Limnology

and Oceanography, 44(3), 804-809. doi: 10.4319/lo.1999.44.3_part_2.0804

NPDES Permit No. IL0023027. (2011).

NPDES Wastewater & Stormwater Permits. (2015). from

http://www3.epa.gov/region9/water/npdes/

Ogbonna, J., Yoshizawa, H., & Tanaka, H. (2000). Treatment of high strength organic

wastewater by a mixed culture of photosynthetic microorganisms. Journal of Applied

Phycology, 12(3-5), 277-284. doi: 10.1023/A:1008188311681

Olguin, E. J. (2003). Phycoremediation: Key issues for cost-effective nutrient removal processes.

Biotechnol Adv, 22(1-2), 81-91.

Oliver, R., & Ganf, G. (2000). The ecology of cyanobacteria: Their diversity in time and space.

Dordrecht: Kluwer Academic.

Olofsson, M., Lamela, T., Nilsson, E., Berge, J. P., Pino, V. d., Uronen, P., & Legrand, C.

(2012). Seasonal variation of lipids and fatty acids of the microalgae Nannochloropsis

oculata grown in outdoor large-scale photobioreactors. Energies, 5, 1577-1592. doi:

10.3390/en5051577

Oxygen Solubility in Fresh and Sea Water. (2015). from www.engineeringtoolbox.com

Park, K. C., Whitney, C. G. E., Kozera, C., O'Leary, S. J. B., & McGinn, P. J. (2015). Seasonal

isolation of microalgae from municipal wastewater for remediation and biofuel

applications. Journal of Applied Microbiology, 119(1), 76-87. doi: 10.1111/jam.12818

Perrine, Z., Negi, S., & Sayre, R. T. (2012). Optimization of photosynthetic light energy

utilization by microalgae. Algal Research, 1(2), 134-142. doi:

http://dx.doi.org/10.1016/j.algal.2012.07.002

Page 92: The Monoraphidium - Northern Illinois University

81

Petrovič, A., & Simonič, M. (2015). The effect of carbon source on nitrate and ammonium

removal from drinking water by immobilised Chlorella sorokiniana. International

Journal of Environmental Science and Technology, 12(10), 3175-3188. doi:

10.1007/s13762-014-0747-0

Polysciences, Inc. (2007). Szechrome reagents for nitrate and nitrite analysis. from

http://www.polysciences.com/skin/frontend/default/polysciences/pdf/239.pdf

Powles, S. B. (1984). Photoinhibition of photosynthesis induced by visible light. Annual Review

of Plant Physiology, 35(1), 15-44. doi: 10.1146/annurev.pp.35.060184.000311

Preliminary Treatment. (2015). from http://dekalbsd.com/the-sewer-plant/preliminary-treatment/

Primary Treatment. (2015). from http://dekalbsd.com/the-sewer-plant/primary-treatment/

Primer for Municipal Wastewater Treatment Systems. (2004). EPA 832-R-04-001, from

http://www3.epa.gov/npdes/pubs/primer.pdf

Program Overview for Renewable Fuel Standard Program. (2005). from

http://www2.epa.gov/renewable-fuel-standard-program/program-overview-renewable-

fuel-standard-program

Rawat, I., Ranjith Kumar, R., Mutanda, T., & Bux, F. (2011). Dual role of microalgae:

Phycoremediation of domestic wastewater and biomass production for sustainable

biofuels production. Applied Energy, 88(10), 3411-3424. doi:

http://dx.doi.org/10.1016/j.apenergy.2010.11.025

Ren, H. Y., Liu, B.-F., Ma, C., Zhao, L., & Ren, N.-Q. (2013). A new lipid-rich microalga

Scenedesmus sp. strain R-16 isolated using Nile red staining: effects of carbon and

nitrogen sources and initial pH on the biomass and lipid production. Biotechnology for

Biofuels(6), 143.

Page 93: The Monoraphidium - Northern Illinois University

82

Reynolds, C. S. (1984). The ecology of freshwater phytoplankton / C.S. Reynolds. New York:

Cambridge University Press.

Rumin, J., Bonnefond, H., Saint-Jean, B., Rouxel, C., Sciandra, A., Bernard, O., . . . Bougaran,

G. (2015). The use of fluorescent Nile red and BODIPY for lipid measurement in

microalgae. Biotechnology for Biofuels, 8(1), 1-16. doi: 10.1186/s13068-015-0220-4

Sawayama, S., Minowa, T., Dote, Y., & Yokoyama, S. (1992). Growth of the hydrocarbon-rich

microalga Botryococcus braunii in secondarily treated sewage. Appl Microbiol

Biotechnol, 38(1), 135-138. doi: 10.1007/BF00169433

Schwinghamer, P., Anderson, D., & Kulis, D. (1991). Seperation and concentration of living

dinoflagellate resting cysts from marine sediments via density-gradient centrifugation.

Limnology and Oceanography, 36(3), 588-592.

Secondary Treatment. (2015). from http://dekalbsd.com/the-sewer-plant/secondary-treatment/

Sharma, K., Schuhmann, H., & Schenk, P. (2012). High lipid induction in microalgae for

biodiesel production. Energies, 5, 1532-1553. doi: 10.3390/en5051532

Shinano, T., Kawamukai, T., Inoue, M., Koike, T., & Tadano, T. (1996). Dimethylsulfoxide

method for the extraction of chlorophylls a and b from the leaves of wheat, field bean,

dwarf bamboo, and oak. Photosynthetica, 32, 409-415.

Smith, B., & Manoylov, K. M. (2013). The effects of variable light and lipids on the water

column distribution and interactions of phytoplankton. Plant, 1(2), 16-24. doi:

10.11648/j.plant.20130102.12

Smith, B., Morrissey, P. J., Guenther, J. E., Nemson, J. A., Harrison, M. A., Allen, J. F., &

Melis, A. (1990). Response of the photosynthetic apparatus in Dunaliella salina (green

algae) to irradiance stress. Plant Physiol, 93(4), 1433-1440.

Page 94: The Monoraphidium - Northern Illinois University

83

Smolders, A. P., Lucassen, E. H. E. T., Bobbink, R., Roelofs, J. M., & Lamers, L. M. (2010).

How nitrate leaching from agricultural lands provokes phosphate eutrophication in

groundwater fed wetlands: The sulphur bridge. Biogeochemistry, 98(1-3), 1-7. doi:

10.1007/s10533-009-9387-8

St. Amand, A. (2012). Identification, ecology and control of nuisance freshwater algae in Illinois

lakes and streams. Illinois Lake Management Society.

Turner, R. E., & Rabalais, N. N. (1994). Coastal eutrophication near the Mississippi River delta.

Nature, 368(6472), 619-621.

Vasilikiotis, C., & Melis, A. (1994). Photosystem II reaction center damage and repair cycle:

Chloroplast acclimation strategy to irradiance stress. Proc Natl Acad Sci U S A, 91(15),

7222-7226.

Wahlen, B. D., Willis, R. M., & Seefeldt, L. C. (2011). Biodiesel production by simultaneous

extraction and conversion of total lipids from microalgae, cyanobacteria, and wild mixed-

cultures. Bioresource Technology, 102(3), 2724-2730. doi:

http://dx.doi.org/10.1016/j.biortech.2010.11.026

Wang, L., Min, M., Li, Y., Chen, P., Chen, Y., Liu, Y., . . . Ruan, R. (2010). Cultivation of green

algae Chlorella sp. in different wastewaters from municipal wastewater treatment plant.

Applied Biochemistry and Biotechnology, 162(4), 1174-1186. doi: 10.1007/s12010-009-

8866-7

Wang, Z. T., Ullrich, N., Joo, S., Waffenschmidt, S., & Goodenough, U. (2009). Algal lipid

bodies: Stress induction, purification, and biochemical characterization in wild-type and

starchless Chlamydomonas reinhardtii. Eukaryotic Cell, 8(12), 1856-1868. doi:

10.1128/EC.00272-09

Ward, M. H., Mark, S. D., Cantor, K. P., Weisenburger, D. D., Correa-Villasenor, A., & Zahm,

S. H. (1996). Drinking water nitrate and the risk of non-Hodgkin's lymphoma.

Epidemiology, 7(5), 465-471.

Page 95: The Monoraphidium - Northern Illinois University

84

Weyer, P. J., Cerhan, J. R., Kross, B. C., Hallberg, G. R., Kantamneni, J., Breuer, G., . . . Lynch,

C. F. (2001). Municipal drinking water nitrate level and cancer risk in older women: The

Iowa Women's Health Study. Epidemiology, 12(3), 327-338.

Whitelam, G., Lanaras, T., & Codd, G. (1983). Rapid separation of microalgae by density

gradient centrifugation in Percoll. British Phycological Journal, 18, 23-28.

Yu, X., Zhao, P., He, C., Li, J., Tang, X., Zhou, J., & Huang, Z. (2012). Isolation of a novel

strain of Monoraphidium sp. and characterization of its potential application as biodiesel

feedstock. Bioresour Technol, 121, 256-262. doi: 10.1016/j.biortech.2012.07.002


Recommended